Anda di halaman 1dari 1193

This PDF is available from The National Academies Press at http://www.nap.edu/catalog.php?

record_id=11536

Twenty-Fifth Symposium on Naval Hydrodynamics

ISBN
978-0-309-10104-2

Office of Naval Research, National Research Council, Institute for Ocean


Technology, Newfoundland, Memorial University of Newfoundland

1192 pages
web only
2005

Visit the National Academies Press online and register for...


Instant access to free PDF downloads of titles from the
NATIONAL ACADEMY OF SCIENCES
NATIONAL ACADEMY OF ENGINEERING
INSTITUTE OF MEDICINE
NATIONAL RESEARCH COUNCIL

10% off print titles


Custom notification of new releases in your field of interest
Special offers and discounts

Distribution, posting, or copying of this PDF is strictly prohibited without written permission of the National Academies Press.
Unless otherwise indicated, all materials in this PDF are copyrighted by the National Academy of Sciences.
Request reprint permission for this book

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Twenty-Fifth Symposium on

NAVAL HYDRODYNAMICS

Wave Hydrodynamics
Propulsor Hydrodynamics
Ships and Propulsion in Ice
Hydrodynamics of Fast or Unconventional Ships
Viscous Ship Hydrodynamics
Hydrodynamics of Underwater Vehicles
Wake Dynamics
Fluid Dynamics in the Naval Context
Cavitation and Bubbly Flows
Nonlinear Wave-Induced Motions and Loads
Frontier Experimental Techniques
Maneuvering
Hydrodynamics in Ship Design and Optimization
Hydrostructural Acoustics

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Twenty-Fifth Symposium on

NAVAL HYDRODYNAMICS

Wave Hydrodynamics
Propulsor Hydrodynamics
Ships and Propulsion in Ice
Hydrodynamics of Fast or Unconventional Ships
Viscous Ship Hydrodynamics
Hydrodynamics of Underwater Vehicles
Wake Dynamics
Fluid Dynamics in the Naval Context
Cavitation and Bubbly Flows
Nonlinear Wave-Induced Motions and Loads
Frontier Experimental Techniques
Maneuvering
Hydrodynamics in Ship Design and Optimization
Hydrostructural Acoustics

Sponsored Jointly by
Office of Naval Research
National Research Council, Institute for Ocean Technology, Newfoundland
Memorial University of Newfoundland
Naval Studies Board

THE NATIONAL ACADEMIES PRESS


Washington, D.C.
www.nap.edu

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

THE NATIONAL ACADEMIES PRESS

500 Fifth Street, N.W.

Washington, DC 20001

NOTICE: The project that is the subject of this proceedings was approved by the Governing Board of the National
Research Council, whose members are drawn from the councils of the National Academy of Sciences, the National
Academy of Engineering, and the Institute of Medicine.
A portion of the work done to prepare this document was performed under Department of the Navy Contract N0001402-I-0563 issued by the Office of Naval Research under contract authority NR 201-124. However, the content does not
necessarily reflect the position or the policy of the Department of the Navy or the government, and no official
endorsement should be inferred.
This work also relates to Department of the Navy Grant N00014-02-1-1007 issued by the Office of Naval Research
International Field Office. The United States Government has at least a royalty-free, nonexclusive, and irrevocable
license throughout the world for government purposes to publish, translate, reproduce, deliver, perform, and dispose of
all or any of this work, and to authorize others so to do.
Copies available from:
Naval Studies Board
National Academies
500 Fifth Street, N.W., Room WS904
Washington, DC 20001
Additional copies of this report are available from the National Academies Press, 500 Fifth Street, N.W., Lockbox 285,
Washington, DC 20055; (800) 624-6242 or (202) 334-3313 (in the Washington metropolitan area); Internet,
http://www.nap.edu
The proceedings are also available online at the National Academies Presss Web site at <http://www.nap.edu>.
Copyright 2005 by the National Academy of Sciences. All rights reserved.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

The National Academy of Sciences is a private, nonprofit, self-perpetuating society of distinguished scholars
engaged in scientific and engineering research, dedicated to the furtherance of science and technology and to their
use for the general welfare. Upon the authority of the charter granted to it by the Congress in 1863, the Academy
has a mandate that requires it to advise the federal government on scientific and technical matters. Dr. Bruce M.
Alberts is president of the National Academy of Sciences.
The National Academy of Engineering was established in 1964, under the charter of the National Academy of
Sciences, as a parallel organization of outstanding engineers. It is autonomous in its administration and in the
selection of its members, sharing with the National Academy of Sciences the responsibility for advising the federal
government. The National Academy of Engineering also sponsors engineering programs aimed at meeting national
needs, encourages education and research, and recognizes the superior achievements of engineers. Dr. Wm. A.
Wulf is president of the National Academy of Engineering.
The Institute of Medicine was established in 1970 by the National Academy of Sciences to secure the services of
eminent members of appropriate professions in the examination of policy matters pertaining to the health of the
public. The Institute acts under the responsibility given to the National Academy of Sciences by its congressional
charter to be an adviser to the federal government and, upon its own initiative, to identify issues of medical care,
research, and education. Dr. Harvey V. Fineberg is president of the Institute of Medicine.
The National Research Council was organized by the National Academy of Sciences in 1916 to associate the
broad community of science and technology with the Academys purposes of furthering knowledge and advising the
federal government. Functioning in accordance with general policies determined by the Academy, the Council has
become the principal operating agency of both the National Academy of Sciences and the National Academy of
Engineering in providing services to the government, the public, and the scientific and engineering communities.
The Council is administered jointly by both Academies and the Institute of Medicine. Dr. Bruce M. Alberts and Dr.
Wm. A. Wulf are chair and vice chair, respectively, of the National Research Council.
www.national-academies.org

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Attendees at the Twenty-Fifth Symposium on Naval Hydrodynamics, St. Johns, Newfoundland, August 8-13, 2004.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Naval Studies Board


JOHN F. EGAN, Nashua, New Hampshire, Chair
MIRIAM E. JOHN, Sandia National Laboratories, Vice Chair
ARTHUR B. BAGGEROER, Massachusetts Institute of Technology
JOHN D. CHRISTIE, LMI
ANTONIO L. ELIAS, Orbital Sciences Corporation
BRIG CHIP ELLIOTT, BBN Technologies
KERRIE L. HOLLEY, IBM Global Services
JOHN W. HUTCHINSON, Harvard University
HARRY W. JENKINS, JR., ITT Industries
DAVID V. KALBAUGH, Applied Physics Laboratory, Johns Hopkins University
ANNETTE J. KRYGIEL, Great Falls, Virginia
THOMAS V. McNAMARA, Charles Stark Draper Laboratory
L. DAVID MONTAGUE, Menlo Park, California
WILLIAM B. MORGAN, Rockville, Maryland
JOHN H. MOXLEY III, Korn/Ferry International
JOHN S. QUILTY, Oakton, Virginia
NILS R. SANDELL, JR., BAE Systems
WILLIAM D. SMITH, Fayetteville, Pennsylvania
JOHN P. STENBIT, Oakton, Virginia
RICHARD L. WADE, Risk Management Sciences
DAVID A. WHELAN, The Boeing Company
CINDY WILLIAMS, Massachusetts Institute of Technology
ELIHU ZIMET, National Defense University
Navy Liaison Representatives
RADM SAMUEL J. LOCKLEAR III, USN, Office of the Chief of Naval Operations, N81
RADM JAY M. COHEN, USN, Office of the Chief of Naval Operations, N091
Marine Corps Liaison Representative
LTGEN JAMES N. MATTIS, USMC, Commanding General, Marine Corps Combat
Development Command
Staff
CHARLES F. DRAPER, Acting Director (as of July 12, 2003)

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

MAURIZIO LANDRINI
March 2, 1963June 26, 2003
Dr. Maurizio Landrini was tragically killed in a motorcycle accident in Rome, Italy, on June 26, 2003. He is
survived by his wife Sara and son Lorenzo, who was born after the accident. Dr. Landrini was an outstanding
marine hydrodynamics researcher who had been selected as the 2003 Georg Weinblum Lecturer. He was born on
March 2, 1963, and earned his Ph.D. degree in theoretical and applied mechanics at the University of Rome. Except
for short periods as a visiting researcher at the Ocean Engineering Laboratory, University of California, Santa
Barbara, and at the Department of Marine Technology, Norwegian University of Science and Technology, he
worked his entire career at INSEAN, the Italian Ship Model Basin. Dr. Landrini had deep knowledge in numerical
and experimental techniques for free-surface waves, fluid-structure interactions, and ship hydrodynamics. He was
manager of the towing tank activities at INSEAN. He was a member of the International Ship and Offshore
Structures Congress Committee on Loads and of the International Towing Tank Conference Maneuvering
Committee. He authored or co-authored over 80 journal articles and conference papers. An enthusiastic and

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

innovative researcher, he was a personal friend to many in the field. He will be greatly missed in the marine
hydrodynamics community.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Foreword

The Twenty-Fifth Symposium on Naval


Hydrodynamics was held in St. Johns,
Newfoundland and Labrador, Canada, from August 8
to 13, 2004. This international symposium was
organized jointly by the Office of Naval Research
(Mechanics and Energy Conversion S&T Division),
the National Research Council of the National
Academies (Naval Studies Board), the National
Research Council, Canada, and Memorial University
of Newfoundland. This biennial symposium
promotes the technical exchange of naval
hydrodynamics research developments of common
interest to all the countries of the world. The forum
encourages both formal and informal discussion of
the presented papers, and the occasion provides an
opportunity for direct communication between
international peers.
Some 165 participants from 21 countries attended
the symposium. The attendees represented a mix of
experience and expertise, from students to
established researchers of international renown. The
breadth of the presentations also permitted the
attendees to learn of the latest developments in fields
of naval hydrodynamics outside those of their own
expertise. Seventy-six papers were presented in
fourteen topical areas, including wave
hydrodynamics, propulsor hydrodynamics, ships and
propulsion in ice, hydrodynamics of fast or
unconventional ships, viscous ship hydrodynamics,
hydrodynamics of underwater vehicles, wake
dynamics, fluid dynamics in the naval context,
cavitation and bubbly flows, nonlinear wave-induced
motions and loads, frontier experimental techniques,
maneuvering, hydro-dynamics in ship design and
optimization, and hydrostructural acoustics. These
topical areas were chosen for this meeting because of
the recent advances made in them and their
importance to the overall field of naval
hydrodynamics. Significant advances presented in
the papers included the continuing increases in
fidelity and resolution in viscous computations from
steady RANS, time-dependent RANS, and detachededdy and large-eddy simulation; increased detail in
the computation of the free surface around ships,
including wave breaking, bubble, and droplet
formation; impressive modeling

of ships in ice; new optics-based instrumentation at


both model and full-scale application; simulation of
bubble dynamics in vortical flow and resulting sound
pressure; a trend toward surface-capturing techniques
such as VOF and level set; and more mature and
realistic shape optimization. These examples
illustrate the timeliness and quality of the work
presented and its importance to the field of naval
hydrodynamics.
The symposium featured invited lectures each
morning. These lectures were presented by S. Jones,
G. Chahine, K. Kijima, and R. Leighton, who
covered, respectively, topics of ships and propulsion
in ice, cavitation, maneuvering in shallow water, and
ship wavebreaking. These presentations by
prominent experts set the pace for the sessions that
followed them.
The success of this symposium is the result of hard
work on the part of many people. There was, of
course, the Organizing and Paper Selection
Committee consisting of Dr. Ki-Han Kim, Dr.
Ronald Joslin, and myself (Office of Naval
Research), Dr. Charles Draper (Naval Studies
Board), Dr. Emilio Campana (INSEAN), Dr. Arthur
Reed (David Taylor Model Basin), Prof. Robert Beck
(Journal of Ship Research), Prof. Choung-Mook Lee
(Pohang University), and Mr. David Murdey
(National Research Council of Canada). The
contribution of this committee was certainly the
cornerstone for the success of the symposium.
However, the administrative preparation and
execution, and the production of these proceedings,
would not have been possible without the support of
Ms. Susan Campbell and the staff of the Naval
Studies Board of the National Research Council and
of Ms. Joanne Harris of Memorial University, who
was instrumental in all of the detailed organizing in
St. Johns. Special appreciation is extended to Ms.
Jennifer McDonald from my office for handling the
abstract collection, the Organizing Committee
meetings and minutes, numerous requests for
information, and the compilation of the discussion
sections.
L. Patrick Purtell
Office of Naval Research

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Contents
OPENING REMARKS
John Leggat, CEO Defense Research and Development Canada
Stephen Lubard, Technical Director, Office of Naval Research
Axel Meisen, President and Vice Chancellor, Memorial University of Newfoundland
William Morgan, Naval Studies Board, National Research Council of the United States
Michael Reymont, National Research Council, Canada
KEYNOTE SPEAKERS
Wave Breaking
R. Leighton (General Dynamics-AIS, USA), paper not available
Ships in IceA Review
S. Jones (Institute for Ocean Technology, National Research Council, Canada)
Nuclei Effects on Cavitation Inception and Noise
G. Chahine (Dynaflow, Inc., USA)
Ship Maneuverability in Shallow Water
K. Kijima (Kyushu University, Japan)
TECHNICAL SESSIONS
Wave Hydrodynamics
Numerical Simulations of Breaking Waves Around an Advancing Ship by an Unstructured NS Solver
T. Hino (National Maritime Research Institute, Japan)
Numerical Simulations of Breaking Wave Around a Wedge
R. Broglia, A. Di Mascio, and R. Muscari (INSEAN, Italian Ship Model Basin, Italy)
A BEM-Level Set Domain Decomposition for Violent Two-Phase Flows in Ship Hydrodynamics
G. Colicchio and M. Greco (INSEAN, Italian Ship Model Basin, Italy) and
O. Faltinsen (Centre for Ships and Ocean Structures, Norwegian University of Science and Technology,
Norway)
The Numerical Simulation of Ship Waves Using Cartesian Grid Methods with Adaptive Mesh Refinement
D. Dommermuth,1 M. Sussman,2 R. Beck,3 T. OShea,1 D. Wyatt,1 K. Olson,4 and P. MacNeice5 (1Science
Applications International Corporation, 2Florida State University, 3University of Michigan, 4University of
Maryland at Baltimore County, 5Drexel University, USA)
Experimental Measurements of the Surface of a Breaking Bow Wave
A. Karion, T. Fu, J. Rice, D. Walker, and D. Furey (Naval Surface Warfare Center, Carderock Division, USA)
Experimental Study of the Bow Wave of the R/V Athena I
T. Fu, A. Karion, J. Rice, and D. Walker (Naval Surface Warfare Center, Carderock Division, USA)
An Experimental Investigation of Breaking Bow Waves Simulated with a 2D+T Technique
M. Shakeri, X. Liu, S. Goll, and J. Duncan (University of Maryland, USA)

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Field Measurements of Bow Spray Droplets


T. Sur and K. Chevalier (Science Applications International Corporation, USA)
Propulsor Hydrodynamics
Computational Design of Trans-Cavitating Propellers and Experimental Evaluation of Their Performance
Y. Ukon, T. Kudo, and J. Fujisawa (National Maritime Research Institute, Japan) and
N. Sasaki (Sumitomo Heavy Industry, Japan)
A Vorticity-Based Propulsor Turbulent Inflow Model
S. Huyer and D. Beal (Naval Undersea Warfare Center, Newport Division, USA)
Toward High-Fidelity Prediction of Tip-Vortex Around Lifting SurfacesWhat Does It Take?
S.-E. Kim and S.H. Rhee (Fluent Incorporated, USA)
Advanced Design, Analysis, and Testing of Waterjet Pumps
T. Michael and C. Chesnakas (Naval Surface Warfare Center, Carderock Division, USA)
Propeller Performance at Extreme Off Design Conditions
S. Jessup, C. Chesnakas, D. Fry, M. Donnelly, S. Black, and J. Park (Naval Surface Warfare Center, Carderock
Division, USA)
Numerical and Experimental Investigation of the Hub Vortex Flow of a Marine Propeller
M. Abdel-Maksoud (Duisburg-Essen University, Germany), K. Hellwig (Potsdam Model Basin, Germany), and
J. Blaurock (Consultant, Hamburg, Germany)
A B-Spline Based Higher Order Panel Method for Analysis of Steady Flow Around Marine Propellers
C.-S. Lee and G.-D. Kim (Chungnam National University, Korea), and J. Kerwin (Massachusetts Institute of
Technology, USA)
Prediction of Performance of Ducted and Podded Propellers
S. Kinnas, H. Lee, H. Gu, and A. Gupta (The University of Texas at Austin, USA)
Ships and Propulsion in Ice
Experimental Uncertainty Analysis for Ship Model Testing in the Ice Tank
A. Derradji-Aouat (National Research Council, Institute for Ocean Technology, Canada)
Preliminary Modeling of Ship Maneuvering in Ice
M. Lau and A. Derradji-Aouat (Institute for Ocean Technology, National Research Council, Canada)
Hydrodynamics of Fast or Unconventional Ships
Experimental and Numerical Study of Semi-Displacement Mono-Hull and Catamaran in Calm Water and Incident
Waves
C. Lugni, A. Colagrossi, and M. Landrini (Istituto Nazionale per Studi ed Esperienze di Architettura Navale,
Italy) and O. Faltinsen (Centre for Ships and Ocean Structures, Norwegian University of Science and
Technology, Norway)
Environmental Wave Generation of High-Speed Marine Vessels
L. Doctors (The University of New South Wales, Australia) and G. Zilman (Tel-Aviv University, Israel)
An Effective Scaling Device for Model Testing of Air Cushion Vehicles in a Laboratory
K. Thiagarajan (The University of Western Australia, Australia)

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Viscous Ship Hydrodynamics


Large Eddy Simulation of the Viscous Flow Around a Ship Hull Including the Free-Surface
E. Lillberg and U. Svennberg (The Swedish Defense Research Agency, FOI, Sweden)
Computation of Free-Surface Viscous Flows at Model and Full Scale by a Steady Iterative Approach
H. Raven, A. van der Ploeg, and B. Starke (Maritime Research Institute, The Netherlands)
The Use of Detached-Eddy Simulation in Ship Hydrodynamics
R. Pattenden, S. Turnock, and N. Bressloff (University of Southampton, United Kingdom)
Numerical Study on Turbulent Flow Around Ship Models Using a Large-Eddy Simulation Technique
H.-H. Chun, J.-C. Park, H.-J. Choi, H.-S. Yoon, and D.-H. Kang (Pusan National University, Korea)
An Investigation of Propeller Inflow for Naval Surface Combatants
J. Gorski, R. Miller, and R. Coleman (Naval Surface Warfare Center, Carderock Division, USA)
The Wakes of Idealized Propeller Shafts with Cross-Flow and Rotation
D. Hally (Defence R&D CanadaAtlantic, Canada)
Complementary RANS Equations for Viscous Flow Computations
K. Kim, A. Sirviente, and R. Beck (University of Michigan, USA)
Experimental Study of the Flow Field Around a Rolling Ship Model
M. Felli, F. Di Felice, and C. Lugni (INSEAN, Istituto Nazionale per Studi ed Esperienze di Architettura
Navale, Italy)
An Investigation of Viscous Roll Damping Through the Application of Particle-Image Velocimetry
R. Bishop, P. Atsavapranee, S. Percival, J. Shan, and A. Engle (Naval Surface Warfare Center, Carderock
Division, USA)
Towing-Tank Tests for Surface Combatant for Free Roll Decay and Coupled Pitch and Heave Motions
M. Irvine, J. Longo, and F. Stern (University of Iowa, USA)
Viscous Flow Simulation Past a Ship in Waves Using the SWENSE Approach
R. Luquet, L. Gentaz, P. Ferrant, and B. Alessandrini (Ecole Centrale de Nantes, France)
Computation of Three-Dimensional Free-Surface Flows with an Automatic Adaptive Mesh Refinement and
Coarsening Strategy
A. Hay, P. Queutey, and M. Visonneau (Ecole Centrale de Nantes, France)
Hydrodynamics of Underwater Vehicles
Experimental Measurements for CFD Validation of the Flow About a Submarine Model (ONR Body-1)
P. Atsavapranee, T. Forlini, D. Furey, J. Hamilton, S. Percival, and C.-H. Sung (Naval Surface Warfare Center,
Carderock Division, USA)
Investigation of the Turbulent Boundary Layer Flow on a Microfilament Towed Array
D. Furey,1 K. Cipolla,2 and P. Atsavapranee1 (1Naval Surface Warfare Center, Carderock Division, USA, 2Naval
Undersea Warfare Center, Newport Division, USA)
Internal Wave Generation by a Horizontally Moving Sphere at Low Froude Number
J. Rottman (Science Applications International Corporation, USA), D. Broutman (Computational Physics, Inc.,
USA), and G. Spedding and P. Meunier (University of Southern California, USA)

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Large Eddy Simulation of the Viscous Flow Around Submarine Hulls


R. Bensow,1 T. Persson,1 C. Fureby,1,2 U. Svennberg,2 and N. Alin2 (1Chalmers University of Technology,
Sweden, 2Swedish Defense Research Agency, FOI, Sweden)
Improvements to AUV Control Surface Hydrodynamic Modeling for Use in Control
P. Ostafichuk, S. Calisal, and D. Cherchas (University of British Columbia, Canada)
Submarine Rising Stability: Quasi-Steady Theory and Unsteady Effects
G. Watt (Defence Research and Development Canada Atlantic, Canada) and H.-J. Bohlmann
(Howaldtswerke-Deutsche Werft, Germany)
Wake Dynamics
Validation of a Quasi-Potential Flow Model for the Analysis of Marine Propellers Wake
L. Greco, F. Salvatore, and F. Di Felice (INSEAN, Italian Ship Model Basin, Italy)
Experimental Analysis of the Wake From a Dynamic Positioning Thruster
S. El Lababidy and N. Bose (Memorial University of Newfoundland, Canada),
P. Liu (National Research Council, Institute for Ocean Technology, Canada), and
F. Di Felice, M. Felli, and F. Pereira (Istituto Nazionale per Studi ed Esperienze di Architettura Navale, Italy)
LDA Measurements in the Wake of the Propelled KCS Model and Its Use to Validate CFD Calculations
L. Lbke and K.-P. Mach (Potsdam Model Basin, Germany)
Comparison of Detailed Simulations of a Turbulent Ship Wake on a Straight and Circular Track
I. Yavuz, Z. Cehreli, and I. Celik (West Virginia University, USA)
Hydrodynamic Wakes of Surface Penetrating Structures
A. Brandt, S. Scorpio, E. Ericson, and C. Schemm (The Johns Hopkins University Applied Physics Laboratory,
USA)
Fluid Dynamics in the Naval Context
A Pseudo-Spectral Method for Non-Linear Wave Hydrodynamics
W. Choi and C. Kent (University of Michigan, USA)
A Theoretical Framework for Marine Applications of Fluid Dynamics
J. Pawlowski (TRDC Inc., Canada)
Air Entrainment Induced by the Impact of a Planar Translating Jet on a Flat Free Surface
A. Iafrati and E. Campana (Italian Ship Model Basin, Italy) and
R. Gomez Ledesma, K. Kiger, and J. Duncan (University of Maryland, USA)
Cavitation and Bubbly Flows
Experimental Investigation of a Cavitating Propeller in Non-Uniform Inflow
F. Pereira and F. Di Felice (INSEAN Propulsion and Cavitation Group, Italy) and M. Soave (CEIMMItalian
Navy Cavitation Tunnel, Italy)
Cavitation Inception in Co-Flow Nozzles
R. Meyer, F. Zajaczkowski, W. Straka, and E. Paterson (Pennsylvania State University, USA)
Numerical Study of Cavitation Inception Due to Vortex/Vortex Interaction in a Ducted Propulsor
C.-T. Hsiao and G. Chahine (Dynaflow, Inc., USA)

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Bubble Drag Reduction at Large Scales and High Reynolds Numbers


W. Sanders,1 J. Cho,1 E. Winkel,1 E. Ivy,1 R. Etter,2 D. Dowling,1 M. Perlin,1 and S. Ceccio1 (1University of
Michigan, USA, 2Naval Surface Warfare Center, Carderock Division, USA)
Tip Vortex Cavitation Inception Study Using the Surface Averaged Pressure (SAP) Model Combined with a Bubble
Splitting Model
J.-K. Choi, C.-T. Hsiao, and G. Chahine (Dynaflow, Inc., USA)
Hydroelastic Modeling for Surface-Piercing Propellers
Y. Young (Princeton University, USA)
Effect of Operational Conditions on the Cavitation Inception Speed of Naval Propellers
T. van Terwisga,1,2 D. Noble,3 R. vant Veer,1 F. Assenberg,1 B. McNeice,4 and P.van Terwisga5 (1Maritime
Research Institute, The Netherlands, 2Delft University of Technology, The Netherlands, 3Defence R&D Canada
Atlantic, Canada, 4Royal Australian Navy, Australia, 5Royal Netherlands Navy, The Netherlands)
Nonlinear Wave-Induced Motions and Loads
Roll Motion in a Nonlinear Pseudo-Spectral Ship Motion Model
R.-Q. Lin and B. Campbell (Naval Surface Warfare Center, Carderock Division, USA)
Validation of Ship Motion Predictions with Sea Trials Data for a Naval Destroyer in Multidirectional Seas
K. McTaggart and D. Stredulinsky (Defence R&D Canada Atlantic, Canada)
A Single-Phase Level Set Method with Application to Breaking Waves and Forward Speed Diffraction Problem
R. Wilson,1 P. Carrica,1, M. Hyman,2 and F. Stern1 (1University of Iowa, USA, 2Naval Surface Warfare
Center, Coastal Systems Station, USA)
A Numerical Study of Nonlinear Diffraction Loads on Floating Bodies Due to Extreme Transient Waves
J. Kim (American Bureau of Shipping, USA), J. Kyoung (Korea Research Institute of Ships and Ocean
Engineering, Korea), K. Bai (Seoul National University, Korea), and R. Ertekin (University of Hawaii at
Manoa, USA)
Prediction of Slamming Loads for Ship Structural Design Using Potential Flow and RANSE Codes
O. el Moctar, A. Brehm, and T. Schellin (Germanischer Lloyd, Germany)
Numerical and Experimental Analysis of Bow Flare Slamming on a Ro-Ro Vessel in Oblique Waves
O. Hermundstad and T. Moan (Centre for Ships and Ocean Structures, Norwegian University of Science and
Technology, Norway)
On the Estimation of Torsional Loads Acting on a Large-Container Ship
R. Miyake, T. Zhu, and A. Kumano (Nippon Kaiji Kyokai, Japan)
H2O: Hierarchical Hydrodynamic Optimization
C. Yang and R. Lhner (George Mason University, USA)
Frontier Experimental Techniques
Quantitative Characterization of the Free-Surface Around Surface Ships
J. Rice, D. Walker, T. Fu, A. Karion, and T. Ratcliffe (Naval Surface Warfare Center, Carderock Division,
USA)

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Maneuvering
Unsteady RANS Simulation of a Maneuvering Ship Hull
A. Di Mascio, R. Broglia, and R. Muscari (Istituto Nazionale per Studi ed Esperienze di Architettura Navale,
Italy) and R. Dattola (Italian Navy General Staff, Italy)
Validation of Forces, Moments and Stability Derivatives of a Maneuvering Series-58 Bare Hull
C.-H. Sung, B. Rhee, and I.-Y. Koh (Naval Surface Warfare Center, Carderock Division, USA)
Hydrodynamics in Ship Design and Optimization
Comparison and Validation of CFD Based Local Optimization Methods for Surface Combatant Bow
E. Campana and D. Peri (INSEAN, Italian Ship Model Basin), Y. Tahara (Osaka Prefecture University, Japan),
and F. Stern (University of Iowa, USA)
Parametric Optimization of SWAT-Hull Forms by a Viscous-Inviscid Free Surface Method Driven by a Differential
Evolution Algorithm
S. Brizzolara (University of Genova, Italy)
Theoretical Hull Form Optimization for Fine Higher-Speed Ships
K.-S. Min Hyundai (Heavy Industries, Korea) and Y.-S. Lee, S.-H. Kang, and B.-W. Han (Hyundai Maritime
Research Institute, Korea)
Hull Form Optimization Using a Free Surface RANSE Solver
E. Jacquin, Q. Derbanne, D. Bellvre, and S. Cordier (Bassin dessais des carnes, France), B. Alessandrini,
(Ecole Centrale de Nantes, France), and Y. Roux (K-Epsilon, France)
Hydrostructural Acoustics
A Physics-Based Simulation Methodology for Predicting Hydrofoil Singing
E. Paterson, J. Poremba, L. Peltier, and S. Hambric (Pennsylvania State University, USA)
Characterizing and Attenuating the Large-Scale Oscillations Downstream of Shallow Cavities Covered by a
Perforated-Lid
S. Jordan (Naval Undersea Warfare Center, Newport Division, USA)
Hull Excitation by Fluctuating and Rotating Acoustic Sources at the Propeller
O. Spivack,1 R. Kinns,2 and N. Peake1 (1University of Cambridge, United Kingdom, 2RKAcoustics, Scotland)

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

List of Attendees
M. Islam
Memorial University of Newfoundland

AUSTRALIA
L.J. Doctors
University of New South Wales
K. Thiagarajan
University of Western Australia
J. Xia
Australian Maritime College

J. Kennedy
Defence Research Development Canada (Atlantic)
V. Klaptocz
University of British Columbia
M. Lau
Institute for Ocean Technology, NRCC

AUSTRIA
G. Strasser
Vienna Model Basin

F. Lin
Martec Limited
P.-I. Liu
Institute for Ocean Technology, NRCC

BRAZIL
J. Sales, Jr.
Cidade Universitaria

C. Moores
Department of National Defence
D.C. Murdey
National Research Council, Canada

BULGARIA
K.Yossifov
Bulgarian Ship Hydrodynamics Centre

D.J. Noble
Defence Research Development Canada (Atlantic)
B. Parsons
Institute for Ocean Technology, NRCC

CANADA
N. Bose
Memorial University of Newfoundland
S. Calisal
University of British Columbia
A. Derradji-Aouat
Institute for Ocean Technology, NRCC
S. El Lababidy
Memorial University of Newfoundland
D. Hally
Defence Research Development Canada (Atlantic)
M. He
Memorial University of Newfoundland
T.C. Humphrey
Consultant

S. Jones
Institute for Ocean Technology, NRCC

J. Pawlowski
TRDC, Inc.
W. Qiu
Memorial University of Newfoundland
S. Sarkar
Memorial University of Newfoundland
D. Spencer
Oceanic Consulting Corporation
R. Taylor
Memorial University of Newfoundland
D. Vyselaar
University of British Columbia
D. Walker
Oceanic Consulting Corporation

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

J. Wang
Memorial University of Newfoundland

L. Lubke
Schiffbau-Versuchsanstalt Potsdam

G. Watt
Defence Research Development Canada (Atlantic)

M. Mehmel
Schiffbau-Versuchsanstalt Potsdam

CHINA

ISRAEL

Y.-S. Wu
China Ship Scientific Research Center

G. Zilman
Tel-Aviv University

K. Yan
China Ship Scientific Research Center

ITALY

FINLAND

G. Caprino
CETENA

T. Kukkanen
VTT Industrial Systems

S. Brizzolara
University of Genova

FRANCE

R. Broglia
Istituto Nazionale per Studi ed Esperienze di
Architettura Navale

B. Alessandrini
Ecole Centrale de Nantes

U.P. Bulgarelli
Istituto Nazionale per Studi ed Esperienze di
Architettura Navale

S. Cordier
Bassin dEssais des Carnes

E. Campana
Istituto Nazionale per Studi ed Esperienze di
Architettura Navale

P. Ferrant
Ecole Centrale de Nantes

G. Colicchio
Istituto Nazionale per Studi ed Esperienze di
Architettura Navale

E. Jacquin
Bassin dEssais des Carnes
R. Luquet
Ecole Centrale de Nantes

R. Dattola
Italian Navy

M. Visonneau
Ecole Centrale de Nantes

F. Di Felice
Istituto Nazionale per Studi ed Esperienze di
Architettura Navale

GERMANY
A. Di Mascio
Istituto Nazionale per Studi ed Esperienze di
Architettura Navale

M. Abdel-Maksoud
University of Duisburg

M. Felli
Istituto Nazionale per Studi ed Esperienze di
Architettura Navale

J. Blaurock
Consultant
O. el Moctar
Germanischer Lloyd
J. Friesch
Hamburgische Shiffbau-Versuchsanstalt

A. Iafrati
Istituto Nazionale per Studi ed Esperienze di
Architettura Navale

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

C. Lugni
Istituto Nazionale per Studi ed Esperienze di
Architettura Navale

KOREA
B.J. Chang
Hyundai Heavy Industries Co., Ltd.

F.J. Pereira
Istituto Nazionale per Studi ed Esperienze di
Architettura Navale

H.-H. Chun
Pusan National University

G. Pisi
Istituto Nazionale per Studi ed Esperienze di
Architettura Navale

C.-G. Kang
Korean Research Institute of Ships and Ocean
Engineering

F. Pistani
Istituto Nazionale per Studi ed Esperienze di
Architettura Navale

C.-S. Lee
Chungham National University

F. Salvatore
Istituto Nazionale per Studi ed Esperienze di
Architettura Navale
M. Soave
Centro Esperienze Idrodinamiche Marina Militare

Y.-S. Lee
Hyundai Heavy Industries Co., Ltd.
J.-C. Park
Pusan National University
W.G. Park
Pusan National University
D.-J. Yum
Hyundai Heavy Industries Co., Ltd.

JAPAN
T. Hino
National Maritime Research Institute

THE NETHERLANDS
K. Kijima
Kyushu University

H.C. Raven
Maritime Research Institute

Y. Kodama
National Maritime Research Institute
A. Masuko
Ishikawajima-Harima Heavy Industries Co., Ltd.

A. van der Ploeg


Maritime Research Institute
T. Van Terwisga
Maritime Research Institute

R. Miyake
Nippon Kaiji Kyokai
NORWAY
H. Miyazaki
National Maritime Research Institute
H. Narita
Office of Naval Research, Asia

O.A. Hermundstad
Marintek
K. Holden
Marintek

Y. Tahara
Osaka Prefecture University
POLAND
N. Toki
Mitsubishi Heavy Industries, Ltd.

L. Wilczynski
Centrum Techniki Okretowej

Y. Ukon
National Maritime Research Institute

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

D. Broutman
Computational Physics, Inc.

RUSSIA
A. Pustoshny
Krylov Shipbuilding Research Institute

B. Campbell
Naval Surface Warfare Center, Carderock
S.L. Ceccio
University of Michigan

SWEDEN
R. Bensow
Chalmers University of Technology
M. Liefvendahl
Swedish Defence Research Agency
T. Persson
Chalmers University of Technology
W. van Berlekom
SSPA Sweden AB

I.B. Celik
University of West Virginia
G.L. Chahine
Dynaflow, Inc.
C. Chesnakas
Naval Surface Warfare Center, Carderock
J.-K. Choi
Dynaflow, Inc.
D. Cusanelli
Naval Surface Warfare Center, Carderock

TAIWAN
S.-K. Chou
United Ship Design and Development Center

F. DeBord
BMT Scientific Marine Services
D. Dommermuth
Science Applications International Corporation

UNITED KINGDOM
M. Atlar
University of Newcastle

J. Dreyer
Applied Research Laboratory, Pennsylvania State
University

C. Jennings
DST International

J. Duncan
University of Maryland

R. Kinns
RKAcoustics
M. Renilson
QinetiQ

J. Eaton
Applied Research Laboratory, Pennsylvania State
University

UNITED STATES

A. Engle
Naval Surface Warfare Center, Carderock

P. Atsavapranee
Naval Surface Warfare Center, Carderock

R. Etter
Naval Surface Warfare Center, Carderock

R.F. Beck
University of Michigan

T. Fu
Naval Surface Warfare Center, Carderock

R. Bishop
Naval Surface Warfare Center, Carderock

D.A. Furey
Naval Surface Warfare Center, Carderock

A. Brandt
Applied Physics Laboratory, Johns Hopkins
University

J.J. Gorski
Naval Surface Warfare Center, Carderock

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

C.-T. Hsiao
Dynaflow, Inc.

W. Morgan
Rockville, Maryland

S. Huyer
Naval Undersea Warfare Center, Newport

T. Nguyen
Naval Surface Warfare Center, Panama City

M. Hyman
Coastal Systems Station, Panama City

E. Paterson
Applied Research Laboratory, Pennsylvania State
University

M. Irvine
Naval Surface Warfare Center, Carderock
S. Jessup
Naval Surface Warfare Center, Carderock
S. Jordan
Naval Undersea Warfare Center, Newport
R.D. Joslin
Office of Naval Research

L. Patrick Purtell
Office of Naval Research
A.M. Reed
Naval Surface Warfare Center, Carderock
J. Rice
Naval Surface Warfare Center, Carderock
J. Rottman
Science Applications International Corporation

A. Karion
Naval Surface Warfare Center, Carderock
J. Kim
American Bureau of Shipping

C. Schemm
Applied Physics Laboratory, Johns Hopkins
University

K.-H. Kim
Naval Surface Warfare Center, Carderock

S. Scorpio
Applied Physics Laboratory, Johns Hopkins
University

S.-E. Kim
Fluent, Inc.

J. Shan
Rutgers University

S. Kinnas
University of Texas at Austin

A. Sirviente
University of Michigan

I.-Y. Koh
Naval Surface Warfare Center, Carderock

J. Slutsky
Naval Surface Warfare Center, Carderock

R. Lahey
Rensselaer Polytechnic Institute

F. Stern
University of Iowa

R. Leighton
General Dynamics

C.-H. Sung
Naval Surface Warfare Center, Carderock

R.-Q. Lin
Naval Surface Warfare Center, Carderock

T. Sur
Science Applications International Corporation

W.M. Lin
Science Applications International Corporation

G. Wilkie
Lockheed Martin Maritime Systems and Sensors

S. Lubard
Office of Naval Research

R. Wilson
University of Iowa

T. Michael
Naval Surface Warfare Center, Carderock

D. Wyatt
Science Applications International Corporation

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

C. Yang
George Mason University
S. Yim
Oregon State University
Y.L. Young
Princeton University

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DRDC Hydrodynamics Research


Past to Present
Dr. John Leggat
CEO Defence Research and Development Canada

Defence Research and


Development Canada

Recherche et dveloppement
pour la dfense Canada

Copyright National Academy of Sciences. All rights reserved.

Canada

Twenty-Fifth Symposium on Naval Hydrodynamics

Propeller Hydrodynamics
Navies have always had a requirement for high performance
propellers, and since the 1940s they have also been been
concerned with quiet propellers:
DRDC started with the hardest problems first - developing
supercavitating propellers for hydrofoil ships (1970s)

Defence R&D Canada Atlantic R & D pour la dfense Canada Atlantique

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

This work led to novel methods for understanding


propeller cavitation

Defence R&D Canada Atlantic R & D pour la dfense Canada Atlantique

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

and on to todays methods of Computational Fluid


Dynamics and computer visualization

Defence R&D Canada Atlantic R & D pour la dfense Canada Atlantique

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Ship Hydrodynamics
Navy ships have a requirement to operate effectively and efficiently in
extreme sea conditions
This has led to continuous improvement of methods to predict motions and
loads
2D strip theory codes have been around for decades and are still used today
Technology has moved on to 3D panel methods and time domain codes
including nonlinear effects

Defence R&D Canada Atlantic R & D pour la dfense Canada Atlantique

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Hydrodynamic prediction codes have been supported


through extensive validation via model tests and
full scale trials

Defence R&D Canada Atlantic R & D pour la dfense Canada Atlantique

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Hydrodynamics codes are now being used for real


time modeling and simulation applications with
DRDCs Virtual Ship

Defence R&D Canada Atlantic R & D pour la dfense Canada Atlantique

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Submarine Hydrodynamics
As with surface ships, requirements are for submarines to
achieve effective hydrodynamic performance and
establish safe maneuvering limitation diagrams
Unlike ships, most of this effort has been undertaken by
navies
Work in this area is a complex mix of analytical and
experimental development

Defence R&D Canada Atlantic R & D pour la dfense Canada Atlantique

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Wind tunnel testing at


NRC Institute for
Aerospace Research in
Ottawa

Marine Dynamic Test Facility


at NRC Institute for Ocean
Technology in St. Johns
Defence R&D Canada Atlantic R & D pour la dfense Canada Atlantique

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

where flow
characteristics
representative of
extreme maneuvers
are obtained.

These are useful for


validating our CFD
capability
Defence R&D Canada Atlantic R & D pour la dfense Canada Atlantique

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Our CFD methods can predict the complex flow


around submarines for moderate flow incidence.
Extreme incidence and unsteady flows are the future
challenge.

Defence R&D Canada Atlantic R & D pour la dfense Canada Atlantique

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Future Technology Challenges


Continued validation of numerical predictions with model
tests and full-scale trials
Generation of effective computational models and meshes
with minimal required labour (e.g., CFD and FE meshes)
Simulations capable of running faster than real time
Interoperable simulation components (e.g., propeller CFD
coupled with ship motion)

Defence R&D Canada Atlantic R & D pour la dfense Canada Atlantique

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

From the U.S. Office of Naval Research

Welcome
To the 25th

Symposium on
Naval Hydrodynamics
Dr. Stephen Lubard
Technical Director

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

The world has changed

Wakes

and naval needs have changed.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Evolution of Naval
Requirements

New
Configurations

Expanded Range
Of Ship Size

USVs

Intratheater

Broad Impact on
Science and Technology Needs
New
Operating Ranges

Unprecedented
Performance

High
Speeds

Reduced
Detection

Copyright National Academy of Sciences. All rights reserved.

New
Technologies

Materials,
Efficiency,
Electric Ship

Twenty-Fifth Symposium on Naval Hydrodynamics

Smaller Craft
Unmanned Surface Vessels
High speed response craft

Challenge: Overcome seastate limits


Stability in waves
Ride control

High Speed Craft

Unmanned Surface Vessel

Deep-V Hull with Lifting Bodies

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Larger Craft / Small Ships


Multihulls for stability
Reduce Drag for efficiency and speed:
-Air cavities
-Evaluate friction drag reduction
SeaFlyer

Challenge: Achieve higher speeds

X-Craft Experimental Platform

SeaCoaster with Air Cavities

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Large Ships / Sealift


Minimize wave resistance
Reduce Friction Drag

Challenge: Achieve high speed, long range, large payload

Trimaran Sealift Concept

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Advanced Technology: Lifting Bodies for Craft and Smaller Ships


Combine buoyant and
dynamic lift
Reduce waterplane area
Provide ride control

Lifting Body Close-up

Hybrid Deep-V Hull with Lifting Bodies

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Advanced Technology: Electric Ship


Propulsors new possibilities
Superconducting Motors

Ducted Propulsor

Podded Propulsor (CFD)

Superconducting Motor

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Advanced Technology: Explosion Resistance


Shock-absorbing surface treatments

Protective Blister (Schematic)

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Enabling Technologies
Computational Fluid Dynamics
Fluid Structure Simulations
Microbubble and polymer drag reduction
Materials Science

40 micron
Sintered
Metal

Flow

MICROBUBBLES

AIR

Photo of Microbubbles
CFD of Propulsor/Hull Interaction
CFD of Lifting Body Pressures

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Education
Goals:
Near Term: Strengthen the
Ocean Engineering, Naval
Architecture and Marine
Engineering departments
Far Term: Recruit the
students to ensure a
strong naval engineering
community for the future

Government

National Naval
Responsibility for
Naval Engineering
Academia

Industry

Ensure the presence of a strong Naval


Engineering discipline in the U. S. to maintain
the U. S. Navys maritime capability

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

The seas have changed, and we are altering course!


We must build on the traditional strengths of Naval
Architecture such as:
Total system engineering
Reliability and safety
but adapt new technologies and methods.
Some examples have been shown here which, hopefully,
will inspire further advances within this community!
Lets plunge ahead!

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

BACKUPS

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

High Speed Craft Technical Issues


NAVY
3,000 tons
50 kts

Overarching Need Maintain useful payload


fraction and range while increasing speed

ARMY
10,000 tons
70 kts

Propulsion
Metrics:

Hull Forms
Metrics:

Hull Materials
Metrics:

Ride Control
Metrics:

Power Density
Efficiency

Minimize drag
(friction, form, and
wavemaking)

Stable, smooth
Controllable /
adjustable

Technologies:
Engine / Drivetrain
Mechnical drive vs.
electric drive
Propulsor choice

Technologies:
Optimize hull form
Control emersion
(dynamic lift)
Fluid drag reduction

Strength vs. weight


Cost
Corrosion resistance
Maintainability

Technologies:

High strength steel


Aluminum
Composites
Coatings

Copyright National Academy of Sciences. All rights reserved.

Technologies:
Environmental
sensing
Algorithms
Control surfaces and
actuators

Twenty-Fifth Symposium on Naval Hydrodynamics

Memorial University of
Newfoundland
Axel Meisen, Ph.D., P.Eng,
President and Vice-Chancellor

25th Symposium on Naval Hydrodynamics


St. Johns NL, Canada
9th August 2004

Ref: AM/HYDR0808.PPT (2004)

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Newfoundland and Labrador

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Area and Population

Newfoundland
and Labrador
Canada
USA
Great Britain

Area
(k km2)

People
(M)

People per
km2

404

0.5

10,000

31.0

9,600

293.0

30

244

60.3

247

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Main Industries

Traditional:
Fishing
Forestry
Iron ore mining

New:
Oil and gas
Nickel
Tourism
Knowledge

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

St. Johns Campus

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Marine Institute
St. Johns

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Sir Wilfred Grenfell College


Corner Brook

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Harlow Campus
England

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Enrollments

Students
Memorial University

17,000

Stanford University, USA

18,300

University of Cambridge, UK

17,300

University of Western Australia

17,000

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Academic Programs
Arts
Business
Education
Engineering
Fine Arts
Human Kinetics
Maritime Studies

Medicine
Music
Nursing
Pharmacy
Science
Social Work
Technology

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Newfoundland and Labradors


Petroleum Industry
Hibernia,
200,000 bbd

Terra Nova
150,000 bbd

Proven Reserves:
Oil: 2.1 B bbl Gas: 9.8 TCF
NGL: 0.4 B bbl
Copyright National Academy of Sciences. All rights reserved.

Whiterose
100,000 bbd

Twenty-Fifth Symposium on Naval Hydrodynamics

Very Large Flume Tank

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Institute for Ocean Technology (NRC)


Ice Tank

Wave Tank
Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Harsh Environment Bridge Simulator


(US$7 million)

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Marine Transportation of
Compressed Natural Gas

Trans Ocean Gas


Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

3D Visualization (US$15 million)

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Bonne Bay Marine Station

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Some Memorial Researchers in


Hydrodynamics
Don
Bass

Serpil
Kocabiyik

Claude
Daley
Neil Bose
Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Design Evaluation

Dan Walker
Oceanic

2003 Americas
Cup Winner
Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Inco Innovation Centre


(US$13 million)

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Petro-Canada Hall (US$1.4)

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Tuition Fees for International


Undergraduate Students (US$)
$30,000
$20,000
$10,000

Copyright National Academy of Sciences. All rights reserved.

W
.A
us
tr
al
ia

am

br
id

ge

or
d
St
an
f

M
em

or
ia
l

$0

Twenty-Fifth Symposium on Naval Hydrodynamics

Memorial University

Partner in Ocean Excellence

Outstanding facilities and people


Commitment to innovation
Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Opening Remarks
William B. Morgan, Dr. Eng. Retired
Former Head of Ship Hydromechanics Directorate
David Taylor Model Basin
Naval Surface Warfare Center, Carderock Division
Good morning, it is an honor for me to make a few opening remarks at this 25th
ONR Naval Hydrodynamics Symposium. In addition, it is always a pleasure to come to
St. Johns and visit friends at IMD (now called IOT) and Memorial University. I should,
particularly, mention Chairman David Murdey whom I first met at the 10th ITTC in
1963.
Since this is the Silver anniversary of the ONR Symposium, I thought it would
be appropriate to spend a little time looking back. About 400 scientists and engineers
attended the first symposium in 1956, which was dedicated in honor of Captain Harold E.
Saunders who was the U.S. Navy captain in charge of building the David Taylor Model
Basin. From the beginning, the US National Academy of Sciences through its National
Research Council, represented today by the Naval Studies Board, were partners with
ONR in these biennial symposia. The first two symposia were held in Washington, D.
C., but starting in 1960 the present rotation of venue was developed with the third
symposium being held in The Netherlands. From that time, the Symposium has
alternated location between the American continent and overseas.
The concept of such a symposium, I believe, was the brainchild of Marshall Tulin
who went from the David Taylor Model Basin to ONR with Phil Eisenberg in the mid1950's. Marshall formulated and developed the technical program of invited papers
which contained critical, timely surveys of various aspects of Naval Hydrodynamics.
The presenters of the surveys whom Marshall assembled were the Who's Who of that
day in hydrodynamics. Some of the presenters were: Louis Milne-Thomson, Sir James
Lighthill, Walter Munk, Georg Weinblum, John Wehausen, Herman Lerbs, Garrett
Birkhoff, John Parkinson, Murray Strasberg, Milton Plesset, Stanley Corrsin, George
Batchelor, among others. Fittingly, over the years the Symposium has evolved into
papers on research topics as proposed by the authors with the Keynote Lectures
becoming the survey papers. This symposium has carried on over the years as the
premier symposium on naval hydrodynamics in the world.
So what has changed over the years - not much as far as the general topics go.
There are participants from more countries today and papers on specific research
dominate. The big change has been the change in technology. Of particular note is the
use of computers for numerical calculations and development of laser-based, nonintrusive instrumentation, e.g. holography, LDV, PDV, etc. In 1956, there was only a
brief mention in one paper and in discussions of five papers of electronic computer
calculations. The only electronic computer data shown was by Stoker in a discussion
of Lighthill's paper entitled River Waves where the observed flood stages of a river
was compared to calculations performed using the UNIVAC. Also in regard to
electronic computer calculations, Saunders said in a discussion of Niedermair's paper
entitled Hydrodynamic Barriers in Ship Design, There has been a lot of talk about
electronic computers, but don't let us forget that, in general, in order to work an electronic

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

computer we have to have some mathematical formulas. In order to derive the


mathematical formulas we must know the physical processes. Have things really
changed? With regard to instrumentation, high-speed photography was available and
stereophotography but the non-intrusive instrumentation that we consider commonplace
today had not been invented. Of course, much of this progress depended on computer
development from the huge room-filled computers to the laptop.
For many years, there has been a steady increase in the number of computer-based
papers and detailed measurements of the flow with non-intrusive instrumentation. This
essentially leveled off a few years ago with few, if any, theoretical papers today. The
concentration is now more on numerical techniques rather than theoretical techniques and
detailed measurements of the flow. There has also been an increase movement from
linear to nonlinear solutions. Another change has been the increase in large-physical
model facilities with more sophisticated instrumentation. This has led to many more
studies of scale effects of the flow and to detailed uncertainty analysis of the
experimental data. This uncertainly analysis is being extended to numerical calculations
as noted by the work of the International Towing Tank Conference (ITTC).
In a cursory review of the ONR Naval Hydrodynamics Symposia over the years, I
found the first paper to give meaningful digital computer calculations for ship-type
flows was at the Second Symposium in 1958 in a paper entitled The Design and
Estimated Performance of a Series of Supercavitating Propellers by A. J. Tachmindji
and W. B. Morgan. The first paper on numerical calculations of viscous flows, which
was based on the Marker-and-Cell technique, was at the Seventh Symposium in 1968
entitled The Numerical Simulation of Viscous Incompressible Fluid Flows by C. W.
Hirt. How many are there today?
The first paper discussing detailed measurements of the flow with non-intrusive
instrumentation was in 1974 at the 10th ONR Symposium where S. J. Barker presented a
paper entitled Laser-Doppler Measurements of Trailing Vortices in a Water Tunnel. J.
H. J. van der Meulen followed in 1978 with a paper on holography at the 12th ONR
Symposium entitled A Holographic Study of the Influence of Boundary Layer and
Surface Characteristics on Incipient and Developed Cavitation on Axisymmetric Bodies.
I am a little hesitant of making future predictions as my crystal ball sometimes
is quite cloudy, but I feel that there will be continued emphasis on viscous flows and on
nonlinear codes. There will be continued coupling of numerical calculations and
experiments. I would think the RANS codes have about run out, as far as theoretical
development. What I mean is that there most likely will not be any major improvements
in the development of RANS codes because of the inherit assumptions in these codes.
RANS codes will continue to be used as a practical tool for design, for the foreseeable
future, and we will develop a better understanding of their limitation and application.
Where will LES go? I am uncertain about LES, per se, and have my doubts. I think for
the time being, until we can handle the full Navier-Stokes equations at high Reynolds
number, some combination of methods, such as, RANS and LES, or similar, will come to
the forefront in the future. Hopefully, sometime in the future, we will be faced with the
question of whether or not the full Navier-Stokes equations are useful to describe the
complicated flow around a ship.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

It is very obvious that the digital computer and non-intrusive flow-measurement


instrumentation have both had a major impact on the quality of the papers at this
symposium and on our understanding of the flow around ships. However, the problems
are far from solved. It seems that inventors will always be able to come up with
concepts with little merit. And there is always the problem with meaningless
calculations and flawed experiments. Therefore, we must always be aware of the many
pitfalls. From my long experience, I am one who has doubts about any calculation and
has distrust of any measurement. I remember years ago, in 1953, I was given the task of
designing the first supercavitating propeller based on the SC section work of Marshall
Tulin. The calculations were by hand with no experimental data for the SC section
performance. When the test data came, I was amazed, as well as everyone else, at how
well the propeller met the design conditions. In back analyzing and rechecking the
hand calculations, I found a grievous error. I had misread the design conditions as
given to me by Dr. Lerbs. I had mistaken a 4 for a 7. Dr. Lerbs writing in his
German style had confused me. In those days, I didn't have sense enough to check and
recheck everything. I don't remember the details but certainly remember the lesson and
the importance of dumb luck. There is certainly a big difference between 7000 and
4000. If I had designed using the 4000 instead of the 7000, the propeller would not have
worked and our work on supercavitating propellers would have come to a screeching halt.
An example of flawed experiments is a design I made many years ago of a set
of contrarotating propellers. The experimental data did not match well with my
calculations. Being young and brash, I said there had to be something wrong with the
experiments. This caused quite a turmoil with the experimenters but they agreed to retest
and double check all the measurements. I was lucky, an error in one of the simplest
measurements was found. The new rpm instrumentation was in error.
Another example, is the importance of intuition? In designing a supercavitating
propeller for a Canadian hydrofoil boat, I felt that the blades were coming out very thin
toward the blades leading edges. At that time, the blade strength was based on the
section modulus. I went to my boss and said I felt the blade leading edges were too thin
even though all the calculations indicate that the propeller would be strong enough. After
substantial debate, we decided to build a second propeller with twice the thickness at the
quarter chord point solely based on gut feeling. What happened? We sent the thinbladed propeller to the Canadians for testing for us and proceeded to build a thicker
propeller. Shortly after, we received a frantic call from the Canadians that they had bent
our propeller in their first run in Bedford Basin. My boss calmly said, Don't worry
about it. We were concerned the propeller might not be strong enough and have built a
second stronger one and we will bring it to you. These antidotes are given to indicate
the importance of dumb luck and your gut feeling in spite of all the tools you have.
In conclusion, I would like to leave you with the following thoughts:
1.

The early theoretical work has been the basis for the advancement in numerical
techniques and our understanding of fluid mechanics.

2.

Progress in computers has led to both improved numerical procedures and flow
measurements.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

3.

There will always be a critical need to look at both numerical and experimental
data. Do they make physical sense?

4.

All the problems are not solved. Hydrodynamics is still the key to the advances in
future ships. The future is still ahead.

Thank you all for making this 25th ONR Naval Hydrodynamic Symposium a crowning
success.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

National Research
Council Canada

Conseil national
de recherches Canada

25th Symposium on Naval


Hydrodynamics

St. Johns, Newfoundland and Labrador


Fairmont Newfoundland Hotel

Michael Raymont

9 August 2004

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics (9 August 2004) 2

Thank you very much, and good morning everyone.


It is a pleasure to be here.
Indeed, it is always a pleasure to visit St. Johns a unique and
beautiful seaport with both much history and an exciting future.
As Acting President of Canadas National Research Council I also
have the honour of reflecting the pride that all of us at NRC feel for
our Institute of Ocean Technology and our NRC-IRAP staff here in
Newfoundland.
I hope that you are already beginning to feel that you made a wise
and appropriate choice in selecting St. Johns as the site for this
symposium - the first to be held in Canada.
This venue gives you a chance to visit not only our extensive facilities
at the NRC Institute for Ocean Technology, but those at Memorial
University and at other organizations working with us to develop a
world-class cluster of ocean engineering and technology expertise
and facilities in this city.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics (9 August 2004) 3

And while these are certainly important attractions, there is another


reason that this city makes for an excellent venue for the
25th Symposium on Naval Hydrodynamics.
This city has for many years, indeed for centuries, been both a
product of, and a contributor to, our understandings of the science of
the seas.
Over five hundred years ago, the emergence of the nation-states, ,
the Renaissance, the accompanying scientific revolution, and the
overseas expansion of European states, all had significant influence
on the development of naval technologies.
As we all know, settlement of North America, principally by
Europeans, became possible after advances in naval technology
opened up the ocean routes to Western navigators.
Seaworthy, manoeuvrable ships which could deal effectively with
adverse sea and wind conditions, gave European adventurers access
to new horizons. It was in the early days of great advances in naval
design and construction that the Italian navigator Giovanni Caboto,
perhaps better known by his Anglicized name, John Cabot, reached
this island in 1497 while sailing under the English flag.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics (9 August 2004) 4

I have to confess to being fascinated also by theories that the North


American continent may have been visited, prior to the 15th century
visits of Europeans, by Chinese seafarers. Since, I arrived here this
morning at past 3 a.m. direct from Beijing, I am glad to know that
today its possible to cover such distances more quickly than those
intrepid seafarers whether from Europe or the Orient.
Cabot's enthusiastic reports attracted international attention, and
English, French, Basque, and Portuguese ships were drawn to fish in
the surrounding waters in the centuries the followed.
For generations, this island and region were regarded by Britain as its
nursery for navies, a training ground for the crews that were to
maintain England's maritime superiority into the distant future.
So, we are standing on land that has a special place in both the
history and the future of naval technology and hydrodynamics.
Today, of course, you are gathered to think about the future. And, of
course, in spite of centuries of advances in naval technologies, there
remains huge potential for future developments in naval design,
construction and ocean exploitation, your work this week is focused
more on the power of partnerships and collaboration sharing
knowledge and building bridges between disciplines.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics (9 August 2004) 5

One of the biggest stumbling blocks to developing new ideas and


technologies lies in the silos that the various disciplines and groups
build around themselves whether from industry, technology, or basic
research. This is the mindset we are challenging with our ocean
engineering cluster.
So, I am very pleased to be among a group of people who are not
shackled by this kind of silo thinking.
Naval hydrodynamics is a field that is I understand - a union of
physics, mathematics, and the practical study of the oceans.
At this symposium, you are clearly looking at new opportunities
which, for example, are coming from interactions with the field of fluid
mechanics and from new international partnerships. With NRC-IOT
and its partners we are also looking at industrial applications and new
business opportunities.
We are proud that you have chosen Canada and specifically
St. Johns for this event and very proud to join the University and our
friends from the United States: the National Academy of Sciences
and the US Department of the Navy Office of Naval Research, in
staging this Symposium.
Best wishes in your deliberations and again welcome to Canada and
St. Johns.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics


St. Johns, Newfoundland and Labrador, CANADA, 8-13 August 2004

Ships In Ice - A Review


Stephen J. Jones
(Institute for Ocean Technology, National Research Council, Canada)

ABSTRACT
A historical review of the literature on the
performance of ships in ice is given from 1888 to
2004.
INTRODUCTION
The object of this paper is to provide an up-to-date
(2004) review of the scientific literature on ship
performance in ice. This forms an updated version of
my previous review (Jones 1989), now 15 years old.
It considers only unclassified work in the open
literature and, unfortunately, probably gives too much
emphasis to those papers written in English. It was
not the intention to deal with the construction of, or
strength of, icebreaking ships, nor is the science of
modelling discussed in any detail.
HISTORICAL REVIEW
To 1900
Runeberg (1888/89) published the first scientific
paper on icebreakers with particular reference to the
Baltic. He discussed both continuous icebreaking
and charging and derived expressions for the
vertical pressure at the bow, the thickness of ice
broken, and the total elevation at the fore-end
calculated from ship geometry for the case of
continuous icebreaking. He claimed that the results
agreed, tolerably well with the actual performance
of six ships. He recognized the importance of hullice friction on resistance, taking, without any
apparent justification, a coefficient of friction of 0.05,
as well as the role of the stem angle of the bow:
the vertical component should be as large as
possible. This is effected by making the bow very
sloping at the waterline. This is still true today.
Nothing else was published in the 19th Century.
1900 1945
Kari (1921) gave, in a brief note, some empirical
equations for determining the required power,
displacement, length, and draught of an icebreaking

ship but no derivation or justification for them was


given. He, also, recognized the importance of low
stem angle to provide a downward force. Simonson
(1936), in what would appear to be the first
contribution from North America, was the first to
recognize the importance of the strength of the ice
and, referring to some experiments at the University
of Illinois (Beach et al. 1895) gave a tensile strength
of freshwater ice as 102-256 psi (0.7-1.8 MPa) for
temperatures of 19.4o to 23o F (-7o to 5o C). He also
showed that the stem angle was important and
derived a simple equation for stem angle as a
function of thrust, vertical force, and trim angle. He
concluded the maximum thickness of ice that can be
broken by a given ship without stalling depends upon
the limiting angle that can be built into the bow, and
he added the frame sections, they should show a
marked flare at the waterline to relieve the crushing
force of the ice.
The only other pre-World War II paper was
a detailed analysis by Shimanskii (1938) who
employed a semi-empirical method for investigating
continuous mode icebreaking resistance.
He
developed several parameters for icebreakers, which
he termed conditional ice quality standards, i.e. the
form of the equation was developed but certain
coefficients in the equation had to be determined
from full-scale data.
This paper must have
influenced the design of the seven large icebreakers
built by the Soviet Union at the end of the 1930s.
During the war an unconventional use of
ships in ice was explored, namely to build aircraft
carriers out of ice. The Habbakuk project (Gold,
1993) has now been well documented and while it led
to much Canadian research on ice properties, no such
carrier was ever built.
1945 - 1960
After World War II, Johnson (1946) described the
U.S. Coast Guards icebreaking vessels and
experience in considerable detail. His comprehensive
paper was more concerned with their strength and
design rather than their performance. He described
the Wind-class icebreakers in detail, which had

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

operated around Greenland and the Russian Arctic


during the war. Vinogradov (1946) described some
of the Russian experience as well as giving an
equation for the downward icebreaking force
developed.
A significant contribution to the literature
was made by Jansson (1956[a] and 1956[b]) with a
major review article. He discussed in detail the
history of icebreaking from what he considered the
earliest true icebreaker, Eisbrecher 1, which operated
between Hamburg and Cuxhafen and was built in
1871, their bow shapes and propellers, to 1956. He
described the history of the bow propellers, which
originated with ships operating on the Great Lakes
where pack ice was a major problem. There, vessels
that got into difficulties were able to force their way
through by backing into the ice. The natural
consequence was that ships were built with bow
propellers. Thus, in 1888, the ferryboat St. Ignace
was built, with a stern propeller driven by 2000 hp,
and fore propeller by 1000 hp. The primary action of
the fore propeller is to wash away water and broken
ice from the fore end of the ship and thus reduce
friction between the ice and the bow sides of the ship.
As mentioned previously, towards the end of the
1930s the Soviet ice breaking fleet had been
augmented by 7 large icebreakers, designed for work
in Arctic waters with three stern propellers, as it
would be useless to try and break the hard polar ice
with fore propellers. Seven Wind-class ships were
built in U.S.A. during and after the 2nd world war, as
well as the Mackinaw, all diesel electric with one fore
propeller and two stern. For operations in the Arctic,
the fore-propeller could be removed and all the power
(10,000 HP) could be split between the two stern
propellers. A major advance after the war was the
first icebreakers equipped with two bow propellers.
This idea originated with the Abegweit, a diesel
electric ferry built in Canada in 1947 for operations
in the Northumberland Straits. The Finnish Voima,
built in 1953, was the first real icebreaker to be
equipped with two bow propellers and two stern.
However, the interest in Arctic type icebreakers
without bow propellers also increased in the midfifties, particularly in Canada.
Jansson (1956[a] and 1956[b]) also
discussed the science of icebreaking. He quoted,
without reference, values for the physical properties
of freshwater ice, apparently at 3oC, as:Elastic Modulus = 70,000 kg/cm2 (6,900 MPa)
Tensile and bending strength = 15 kg/cm2 (1.5 MPa)
Compressive strength = 30 kg/cm2 (2.9 MPa)
Shear strength = 7 kg/cm2 (0.7 MPa)

and said he had failed to find any reliable values for


sea ice. He said that the strength increased with
lower temperatures and even followed a rule that
ultimate strength is approximately proportional
to the square root of the number of degrees below
freezing point. No details were given about these
experiments, which is unfortunate. He also quoted
values of the coefficient of friction between ice and
metal as 0.10 to 0.15 for fresh, or Baltic, ice and 0.20
for salt water and polar ice. He gave a simple
formula for the total ice resistance as:R ice = (C1h + C 2 hv 2 ) B

(1)

where C1 and C2 are experimental constants, h is ice


thickness, v is vessel speed and B is breadth of vessel
at waterline.
In December 1957 the Lenin was launched
in Leningrad (St. Petersburg). It was the first atomic
or nuclear powered icebreaker and represented a
major technological achievement (Alexandrov et al.
1959). It claimed to have a cruising speed of 2 knot
in ice 2.4 m (8 ft) thick, and could remain at sea for
one year.
At a Society of Naval Architects and Marine
Engineers (SNAME) Spring Meeting held to
celebrate the opening of the St. Lawrence Seaway,
German (1959) and Watson (1959[a][b]) both
reviewed the Canadian experience and described the
icebreakers then in service and those planned for the
Canadian Department of Transport. Thiele (1959)
described the technical aspects of icebreaking
operation stressing four problems including friction,
and Ferris (1959) discussed the proportions and
forms of icebreakers.
1960 1985
The vast majority of the literature has been published
since 1960. The Manhattan voyage in 1969, and the
dramatic rise in oil prices in 1973 and again in 1979,
which led to a promise of extensive Arctic
development, contributed to the importance of
icebreaker design and to a corresponding interest in
structures for use in ice-covered waters. The advent
of model tests, ice tanks, analytical and numerical
techniques has meant a more scientific approach to
the subject. One of the first model tests was
described by Corlett and Snaith (1964), who used a
wax-like substance for their ice, for the Perkun, a
small Baltic icebreaker.
Kashteljan et al. (1968) are usually credited
with the first detailed attempt to analyze level ice
resistance by breaking it down into components.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

They gave an equation for the total ice resistance,


RTOT,
R TOT = R1 + R 2 + R 3 + R 4

(2)

where:R1 = resistance due to breaking the ice


R2 = resistance due to forces connected with weight
(i.e. submersion of broken ice, turning of broken ice,
change of position of icebreaker, and dry friction
resistance)
R3 = resistance due to passage through broken ice
R4 = water friction and wavemaking resistance
Their equation is (without R4)
1 k4 k5
B v (3)
2
where is ice strength, B is ship beam, h is ice
thickness, v is ship speed, and i is the density of ice.
R TOT = k 1 o Bh + k 2 o B i h 2 + k 3

o and 2 are related to Shimanskys ice cutting


parameters and k1, k2, k3, k4, k5, are coefficients
experimentally determined (0.004, 3.6, 0.25, 1.65,
and 1.0 respectively). This equation was developed
from model and full-scale tests of the Ermak.
Lewis and Edwards (1970) gave a good
review of previous work and derived the equation
R im = C o h 2 + C1 i gBh 2 + C 2 iBhv 2

with C0=0.146, C1=8.840, and C2=5.905. They went


on to analyse the Wind-class data more thoroughly in
a similar non-dimensional way, and showed best-fit
curves between the full-scale Wind-class data and
their semi-empirical equation utilising (a) all data and
(b) just model data. Their model data, however,
predicted a v2 term, which was not found in the fullscale data. They included a snow cover term in their
regression analysis of the full-scale data, which gave
an added resistance of about 2 long tons/inch
(8kN/cm) for the Wind-class icebreaker.
White (1970) gave a purely analytical
method for calculating bow performance. His major
contribution was to identify those qualities of a bow
that would be desirable for (a) improved continuous
icebreaking, (b) improved ramming and (c) improved
extraction ability. He concluded that there were only
three qualities that would improve all three
capabilities simultaneously namely;
(a) decrease of spread angle complement (i.e. a
blunter bow)
(b) decrease of the coefficient of friction
(c) increase of thrust.
He proposed a bow form, shown in Fig. 1, which
incorporated these features. This form was used on
the Manhattan for its voyage in the Arctic.

(4)

Rim = mean resistance excluding water


g = acceleration due to gravity
Co, C1, C2 are non-dimensional coefficients to be
determined experimentally.
The first term represents ice breaking and
friction, the second accounts for all resistance forces
attributable to ice buoyancy, and the third accounts
for all resistance forces attributable to momentum
interchange between the ship and the broken ice.
They then non-dimensionalized the equation by
where

dividing by h to get
R ' = C o + C1B ' N + C 2 B ' NI

(5)

where R ' = R im h 2 , non-dimensional mean ice


resistance
B' = B h , non-dimensional beam
N = i gh , volumetric number
NI = i v 2 , inertial number

Fig. 1 Whites (1969) recommended bow form for


a polar icebreaker, as used in the design of the
Manhattan.
Crago et al. (1971) described a set of model
tests in wax-type ice on 11 icebreakers. By
considering a simple bow geometry and the vertical
force acting on the ice sheet, they derived a
theoretical equation for the ice thickness, h, broken:-

and then obtained a best fit with full-scale and modelscale tests of Wind-class, Raritan, M-9 and M-15

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

h
T

1.53

(6)

tan(i + )

where = ice tensile strength, T= thrust, i = stem


angle and = tan-1f, where f is the coefficient of
friction.
They then plotted
(1

(h

T)

Finncarrier, and Jelppari, and was able to compare


his results with limited full-scale data from all three.
From a combination of analytical work, dimensional
analysis, and a few assumptions, he derived a semiempirical equation for ice resistance based on three
terms:RTOT = C1 Bh + C2 BhT g + C3 Bhi v2 (8)

against

tan(i + ) ) , as shown in Fig. 2, and obtained

good correlation between equation (5) and their


model tests. They also had one full-scale data point
from the CCGS Wolfe. While the model data did fit
their equation quite well, but not with a slope of 1.53,
it seriously over-predicted the ice thickness broken
for a given thrust, compared to the one full-scale
point.

Fig. 2. Showing the relationship by Crago et al.


(1971) between their model data, one full-scale point,
and one new model ice result and their equation for
thrust.
However, they pointed out that the one model test in
new model ice was in good agreement with the
full-scale point, as shown in Fig. 2. It is, perhaps, not
surprising that the agreement with the full-scale data
was poor, since they considered only a simple
breaking term in their equation and neglected all
others, such as submersion of ice pieces, as had been
considered by Kashteljan et al. (1968). Crago et al.
(1971) also measured the friction of a dry, unpainted,
steel plate toboggan against a dry crusty snow cover.
They obtained mean values of static, fs, and kinetic
frictions, fk, of:(7)
fs = 0.30 0.35
fk = 0.07 0.23
Enkvist (1972) made a major addition to the literature
of ship performance in level ice. He conducted
model tests for three ships; Moskva-class,

where

T = draft of ship
w = density of water and = w - i .
By doing model tests at low speed (v=0) as well
as normal speeds he was able to isolate the velocity
dependent term, and by doing tests in pre-sawn ice
( = 0 ) he was able to isolate the submergence term.
He was able, therefore, to determine the relative
importance of the three terms in his equation.
Enkvist (1972) also conducted detailed tests on the
strength of his model ice, described strength tests on
natural ice, and carried out a considerable number of
friction tests on his model ice and on natural ice
surfaces using a towed sled the first person to
describe such tests in any detail. In a later study,
Enkvist (1983) applied his model-scale technique of
doing tests in pre-sawn ice and creeping speeds, to 16
full-scale tests. From these tests he obtained the
result that the breaking term at full-scale was greater
than he had previously estimated, between 40 and
80% of the total zero speed resistance, with the larger
figure applying to smaller ships. This is probably
still the most reliable published estimate of the
importance of the breaking term at full-scale. At
model scale, Poznak and Ionov (1981) showed that
for a medium size icebreaker the breaking term was
about 40% of the total ice resistance, and the friction
term about 30%.
Johansson and Mkinen (1973) applied
Enkvists method of analysis to model tests of a
parametric series of nine bulk carrier models. Their
results showed that
1. A reduction of bow angle from 82o to 20o
reduced the ice resistance by about 60%.
2. an increase in length of 38% increased the
ice resistance by about 30%. A decrease in
length of 38% decreased the ice resistance
by 10%.
3. An increase in beam of 33% increased the
ice resistance by about 40%. A decrease in
beam of 27% reduced the ice resistance by
about 36%.
They later (Virtanen et al., 1975) investigated the
effect of draft and found no effect on resistance,
within the errors of their experiment.
Edwards et al. (1972) conducted an extensive
set of full-scale and model-scale tests on a Great
Lakes icebreaker, the USCGC Mackinaw. Their full-

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

scale resistance was, however, determined


indirectly as the sum of the estimated thrust in each
of the three propeller shafts (two aft, one forward)
determined almost exclusively from electrical
readings of current, voltage, and r.p.m.
Milano (1973) made a significant advance in
the purely theoretical prediction of ship performance
in ice. He considered the energy needed for a ship to
move through level ice, which varied somewhat with
ice thickness. For example, for very thick ice the
ship moves through the ice-filled channel (E1),
impacts the various bow and cusp wedges causing
local crushing (E2), climbs onto the ice (E3) until
sufficient force is generated to cause fracture, at
which time the ship falls (E4), and moves forward,
forcing the ice downward (E5). The total energy loss
due to ship motion is then written as

E T = E1 + E 2 + E 3 + E 4 + E 5

dependence, at least in thick ice, is interesting


because of its complexity (Fig. 4) and shows what
has become known as a Milano hump. His
explanation for this hump is related to the different
mechanisms involved in the energy equation. Some
experimental evidence for such a hump has been
found by Tatinclaux (1984), Schwarz (1977) and
Narita and Yamaguchi (1981). Milano (1975) then
varied numerous ship and ice parameters and showed
how this affected his calculated resistance. His plots
showed the trend in resistance to be expected by
varying ship parameters such as beam, block
coefficient, waterplane coefficient, length, etc., and
also what would happen if ice properties such as
friction, tensile strength and compressive strength
were altered.

(9)

Then he derived explicit analytical expressions for


each of these terms and compared his predictions
with the data obtained on the Mackinaw, discussed
above, (Edwards et al., 1972) the Wind-class vessel
Staten Island, and the Raritan. He obtained good
correlation, as shown in Fig. 3, although this
correlation was dependent on the value of ice flexural
strength and friction coefficient used. He, also, used
a non-dimensional approach, following Lewis and
Edwards, and developed a design chart for

Fig. 4. Milanos (1975) plot of total energy lost


versus ship speed showing component energy terms
for Mackinaw in ice two feet thick, and showing the
development of a hump.
Fig 3. Plot of ship resistance versus speed for
USCGC Mackinaw as a function of ice thickness,
showing correlation with full-scale tests (Milano,
1973)
predicting total resistance for all icebreakers in
general and large polar-type icebreakers in
particular, a somewhat ambitious exercise! In later
papers Milano (1975, 1980, 1982) investigated in
detail the effect on his analytical model of varying
various ship or ice parameters. His proposed speed

Carter (1983) has also attempted an


analytical approach to ship resistance in ice. He
derived a relatively simple equation for the maximum
resistance to ship motion. He neglected inertial
forces and buoyancy forces entirely on the grounds
that the effect of upturning and submerging the ice
pieces was small and could be ignored. The net total
energy lost was set, in his theory, equal to that
absorbed in icebreaking by bending, buckling, or
crushing. However, model tests by Enkvist (1972)
and others, do not bear out this assumption. Despite

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

this, Carter (1983) obtained reasonable agreement


between his theory and data for six icebreakers.
Scarton (1975) investigated the role of
friction in icebreaking and specifically studied
theoretically the direction of the frictional force. He
derived a relationship between bow angles and the
coefficient of friction such that a ship would not get
stuck in the ice. Mkinen et al. (1975) showed
clearly the importance of friction in the most direct
way. By attaching stainless steel plates, which had a
friction coefficient of about half the remainder of the
hull, to the Jelppari, they showed that the resistance
dropped significantly, particularly at low speeds.
They also compared two full-scale ships with
different surface finishes as well as observing model
scale effects. All showed a significant drop in
resistance with reduced friction coefficient. They
described tests with the Murtaja using different
coatings at different places on the hull and found a
solvent free epoxy (INERTA 160) was the best in
terms of reducing friction and staying attached to the
hull.
Vance (1975) obtained an optimum
regression equation from five sets of model and fullscale data, of the Mackinaw (same data as used by
Edwards et al., 1972), Moskza, Finncarrier, Staten
Island, and Ermak. His equation was :R(ice) = CSgBh2 + CBBh+ CVi v2Lh0.65 B 0.35

(10)

They quote hull-ice friction coefficients varying from


0.08 to 0.48 but did not explain how these were
obtained.

Fig. 5. Vance (1975) analysis of Mackinaw data.


MSR is the model-scale regression curve (i.e.
obtained from model tests) and FSR is the full-scale
regression curve (i.e. obtained form full-scale tests).
FS is the actual full-scale data. All for three different
ice thicknesses, 0.3, 0.9, and 1.6 ft (9, 27 and 49 cm).

where R (ice) is the resistance due to ice, L is length of


vessel, and CS, CB, CV are empirically determined
values. The first term is a submergence term, the
second a breaking term, and the third term is a
velocity dependent resistance.
An example of a fit to his equation is shown
in Fig. 5, in which the Mackinaw full-scale data
(labeled FS) are shown fitted to his equation above
(labeled FSR) and a model-scale regression to his
equation (MSR) is also shown. Good agreement is
found between the model and full-scale results.
Edwards et al. (1976) presented full-scale
data for the Louis S. St. Laurent collected by
analyzing ramming type tests using a nondimensional equation:
R
v
= 4.24 + 0.05
+ 8. 9
(11)
2
w gh
w gBh
gh
which is linear in velocity. Their results are shown in
Fig. 6 in which the five lines are obtained from their
regression above using the values of and h
appropriate to the course. They compared these
results to two sets of saline ice model data, one
collected at a scale of 1:36 and one at a scale of 1:48.

Fig. 6. Edwards et al (1976) regression of full-scale


data from the Louis S. St. Laurent using values of
and h appropriate to the course.
They also conducted a parametric series of tests on
nine different models. They concluded that level ice
resistance was

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

(a) directly proportional to beam


(b) independent of length
(c) proportional to block coefficient, and
(d) proportional to draft.
These results are somewhat surprising, and in some
disagreement with the earlier work of Johannson and
Mkinen (1973).
Particularly surprising is the
independence of length, since Edwards et al. (1976)
had earlier shown the importance of a frictional term,
which one might expect to be a function of length.
They also conducted manoeuvring tests.
In a related series of experiments, Kitagawa
et al., (1982, 1983, 1986) investigated the effect of
parallel mid-body length, and beam, on an Arctic
tanker model

Fig. 7. Resistance per unit displacement for three


Arctic tanker models of different lengths as shown,
scaled-up to a ship of length 360 m (Kitagawa et al.,
1982)
They found a clear increase in resistance with
increase in length, using three models of lengths 3.75
m, 5.0 m, and 6.4 m. They plotted their results as
resistance per unit displacement against speed at both
model scale and scaled up to a 360 m long, 280,900
m3 displacement, vessel. This involved different
scaling factors for the three models, and therefore
ignored differences in model ice strength, which was
probably not insignificant (Kloppenburg, 1975).
Their results, both at model and full-scale, indicated
an optimum parallel mid-body as shown for example
in Fig. 7 in which the data are scaled up to a 360 m
long ship. Model B-003 has the least resistance per
unit displacement and is the 5.0 m long model i.e. the
middle length of the three tested, with a Lparallel/Lmodel
of 0.4. Correcting for model ice strength would have
the effect of increasing the resistance of B-004 even
more and reducing B-005 slightly. They concluded,
therefore, that for this particular hull form, the
maximum length of parallel body should be 0.4 LPP.
In a corresponding series of self-propulsion tests,

they concluded that a minimum parallel body of 0.25


LPP was needed top avoid excessive propeller-ice
interaction. When they varied the beam of the model
for a fixed draught, using values of L/B=8 and 6, they
found an increase in resistance with increase in beam.
However, the wider ship had a significantly lower
resistance per unit displacement.
They also
investigated the effect of an 8o side flare to the
parallel mid-body. While this increased the level ice
resistance somewhat, it had certain advantages; the
open-water channel width was slightly wider, and no
asymmetrical roll was observed, as had been seen
with the vertical sided model. They also observed
that a 5o rise of floor in the parallel body had a
significant beneficial effect in allowing broken ice
pieces to rise to the surface before reaching the
propellers.
In 1981, a STAR symposium held in Ottawa
published a number of model tests and some fullscale data. Narita and Yamaguchi (1981) published a
very detailed account of model tests, which had led to
the building of the Shirase. First, they tested three
model bows and showed that a cylindrical bow with a
low stem angle of 22.5o had less resistance than the
other two, because it avoided crushing at the bow.
They went on to test a triple-screw ship in resistance
and self-propulsion. They also showed, at modelscale, that the resistance almost doubled as the hullice friction coefficient doubled from 0.1 to 0.2.
Schwarz et al. (1981) published model tests of the
Polarstern, and Juurmaa and Segercrantz (1981)
stressed the importance of propulsion efficiency in
ice, rather than just resistance, pointing out that while
different models might have the same resistance, they
could have very different efficiencies due to
ice/propeller interaction.
They showed that a
propeller with a nozzle could have very low
efficiency if it became blocked with ice.
Full-scale data for the Canadian R-class
icebreakers were also presented at this conference
(Edwards et al., 1981; Michailidis and Murdey,
1981), as well as a set of parametric variation model
tests, which examined different bow forms based on
the R-Class as parent (Noble and Bulat, 1981).
Resistance tests only were conducted, and these,
again, showed the superiority of rounded bows with
low stem angle in breaking ice, but since no selfpropulsion tests were conducted it is impossible to
judge the overall performance of the different ships.
Vance (1980) and Vance et al. (1981) conducted fullscale tests of the 140 ft (43 m) Great Lakes
icebreaker, Katmai Bay. He analysed his results
somewhat differently from other workers, plotting
Propulsive Coefficient (PC) against velocity, where

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

models. Tatinclauxs (1984) tests in level and broken


ice allowed ice resistance to be divided into a
submergence-inertia component and an ice-breaking
component. The ice-breaking component was found
to be proportional to the Cauchy Number ( w h) ,
as expected, but was influenced by Froude Number
(v

gh ) . In particular a rapid change in the ice-

breaking resistance was found to occur at a Froude


Number of 0.4-0.5, as shown in Fig. 9.
Fig. 8. PC versus velocity for Katmai Bay in level
ice with no bubblers operating. Clearwater value, not
shown, was 0.565 (Vance, 1980).
PC = EHP/SHP
(12)
and EHP = Effective horsepower, SHP = Shaft
horsepower. EHP was calculated from
EHP = Resistance X Velocity

(13)

and resistance, R, was determined from


R = T(1-t)

(14)

where T was the thrust measured on the shaft, and t


was a thrust deduction factor, taken as 0.2. SHP was
calculated from
SHP = Measured torque X R.P.M.

(15)

He found that PC in level ice was always lower


(0.12-0.45) than in clearwater (0.565) as shown in
Fig. 8, and he suggested several reasons for this loss
in efficiency, namely:1.
2.
3.
4.

increase in t
decrease in w (ice-free wake fraction)
decrease in relative rotative efficiency by
disturbances to flow pattern by ice blocks
decrease in propeller efficiency for other
reasons, as discussed in the paper.

Tatinclaux (1984) performed resistance tests on two


models of the Katmai Bay, 1:10 scale and 1:24 scale,
in level and brash ice.
He found that the
dimensionless ice resistance in level ice was
essentially the same for both models i.e. no scale
effect. Newbury and Williams (1986) did find a scale
effect when testing 1:40 and 1:20 scale models of the
R-Class icebreakers, but they attributed it to
differences in the friction coefficient between the

Fig. 9. Plot of (Rbk Bhi ) as a function of Froude


number, Fn (dimensionless quantities) for Katmai Bay
model tests (Tatinclaux, 1984).
This behaviour was attributed to corresponding
observed changes in the amplitude of pitching and
heaving motions of the models, and may correspond
to a Milano Hump as discussed earlier.
Comparison with full-scale data (Vance, 1980)
indicated that the model resistance was significantly
larger, when scaled up, than the full-scale data, and
several possible reasons were suggested. A further
set of self-propelled model tests (Tatinclaux, 1985)
showed reasonable agreement with full-scale but
several possible sources of error were identified; the
Froude Number was not the same for model and fullscale tests, a stock propeller was used in the model
tests which might not have been as efficient the real
propeller, and the model friction coefficient may have
been higher than the full-scale value.
Bulat (1982) investigated the effect of snow
cover on level ice resistance. He used published data
from five full-scale trials (Radisson, Franklin, Staten
Island, Mackinaw, Wolfe) and plotted the percentage
increase in resistance against non-dimensional snow
cover, as shown in Fig. 10.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

R S ,R Sf = normal and frictional resistance


due to submerging broken ice floes.

Fig. 10. Plot of actual data points from full-scale


trials to show the influence of snow cover on ice
resistance (Bulat, 1982).
This was done by comparing full-scale data points
with and without snow cover, but in similar ice
conditions. He concluded that there was no evidence
that the actual hull shape influenced ship
performance in snow covered ice (at least within the
range of hull forms tested) and that the relative
resistance increased with decrease of speed, increase
of snow cover, and decrease of ice thickness.
Lewis et al (1982) re-analysed earlier data in
a different manner, the major improvement being the
inclusion of a thrust deduction factor applied to their
full-scale measurements. They assumed that t i = t p ,
the thrust deduction factor in open water tests, which
they used to convert the thrust measured during ice
breaking trials to resistance which could be compared
to model-scale resistance measurements. This they
did for the Mackinaw, Katmai Bay, and Radisson,
plotting model-test resistance against full-scale
resistance and found reasonable agreement.
Kotras et al. (1983), in a paper based on
Nagles thesis (Nagle, unpublished) describe yet
another semi-empirical approach in which the total
ice resistance is given by
R ice = R B + R Bf + R T + R Tf + R S + R Sf

where

R ice

(16)

= total ship-ice resistance

R B ,R Bf = normal and frictional resistance


due to breaking of level ice
R T ,R Tf = normal and frictional resistance
due to turning broken ice floes.

The resulting equation contained four empirical


coefficients these were determined from best fits to
some of the data from Katmai Bay, Mackinaw,
Radisson, Staten Island, and Manhattan.
The
remainder of the data, plus that used to optimize the
coefficients was then plotted as measured ice
resistance against the ice resistance predicted from
their equation. The measured ice resistance had
been obtained from the full-scale measurements by
applying a thrust deduction factor as discussed by
Lewis et al. (1982). 72% of the data fell between
25% of the perfect correlation line, which they
claimed was a significant improvement over the
equation given by Lewis et al. (1982).
1985-2004
This modern period has seen the development of new
icebreaking forms, and a more scientific approach to
the modeling of ships in ice with extensive model
testing and, most recently, numerical methods.
Canadian Arctic oil exploration and development led
to new designs such as the Kigoriak, and Terry Fox,
while other activities led to the Oden, double acting
tankers (DAT) with Azipods, FPSOs in ice, and
research ships such as the Nathaniel B. Palmer,
USCGC Healy, and the converted CCGS Franklin
now called CCGS Amundsen.
Baker and Nishizaki (1986) described a new
bow form for an Arctic tanker and compared model
tests done by several laboratories. The results were
somewhat disappointing scientifically because the
full-scale predictions by the three laboratories
differed widely as shown in Fig. 11. Reasons for this
were suggested by the authors as differences in ice
modeling and analysis procedures, as well as a lack
of understanding of friction and thrust deduction
effects. Similar comparison work by the ITTC, Fig.
12, on a model of the R-Class icebreaker, has also
shown a certain lack of agreement, but closer than the
Baker and Nishizaka (1986) comparison. This
disagreement is attributed to the different model ices
and analysis procedures used by the tanks. However,
Takekuma and Kayo (1988) apparently obtained
quite good agreement in two ice tanks with both
structure and ship models.
A study of the dynamics of continuous mode
icebreaking (Ettema et al., 1987) using 1:48 scale
model of the Polar-class icebreaker, showed that a
free hull (free in pitch, heave, and roll) experienced
larger values of mean resistance than did a fixed hull.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

The 1980s saw the design and construction


of icebreakers with unconventional bow forms all of
which have low stem angles of approximately 20o
and are different from the classical wedge-shaped
bow. These are the spoon-shaped bows of the
Canmar Kigoriak, Robert LeMeur, and other similar
designs, and the Thyssen-Waas bow form of the
modified Max Waldeck and the converted Mudyuq.
Fig 13 shows these types of bow at model scale, in

Fig. 11. Comparison of full-scale thrust predictions


from three laboratories, ACL, HSVA and WARC, for
the old bow of the MV Arctic. Ice strength = 500
kPa. Considerable differences in predicted thrust are
apparent (Baker and Nishizaki, 1986).

Fig. 13. Models of typical icebreaking bows showing


from L-R, the original bow of a CCG Navaids tender
Bernier, an R-Class bow, a Beaufort Sea type bow,
and a Thyssen-Waaas bow (Glen et al. 1998).

Fig. 12. Comparison of resistance tests conducted by


different ice tanks on a model of the R-Class
icebreaker (18th ITTC Proceedings).
The dominant frequency of ice resistance and hull
motions experienced by the free hull occurred either
at integer fractions of icebreaking frequency, b , or
at the hulls natural frequency of coupled pitch and
heave, n . The fixed hull experienced cycles of
resistance predominantly at frequency, b . Further
experiments such as these to study the dynamics of
icebreaking, should help us to understand more
clearly what is happening in the icebreaking process.

which the bow forms of four ships are compared: an


original bow of the Bernier a CCG Navaids vessel, an
R-Class bow, a Beaufort Sea bow typical of the
Kigoriak, and a Thyssen-Waas type bow.
The general design and operation of these
ships has been published (Churcher et al., 1984;
Ghonheim et al., 1984; Freitas and Nishizaki, 1986;
Schwarz, 1986[a]; German, 1983; Tronin et al., 1984;
Johansson and Revill, 1986) but little in the way of
full-scale trials or even detailed model tests.
Hellmann (1982) described model and full-scale tests
with the Max Waldeck before and after conversion to
a Thyssen-Waas bow form.
He showed an
approximate 25% drop in resistance model tests, and
100% increase in speed, for the same power, in
propulsion tests. Full-scale data gave reasonable
agreement. Enkvist and Mustamki (1986) have
published results of model and full-scale tests of a
bow, which was derived from tests of a circular and
square bow form. They showed, first of all, that ice
crushing at the stem of two small ships accounted for
20-40% of the total low-speed resistance. By cutting

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

slots in the ice ahead of the stem and removing the


ice, the resistance was reduced by this amount.
Clearly this is the major advantage of low stem angle,
non wedge-shaped bows.
Their model tests
compared a circular bow, a square bow, and the
original Mudyuq bow, and showed that the circular
bow had the lowest resistance. They then selected an
experimental bow for further testing and analysis and
after model testing, made a full-scale bow to attach to
the Protector. Their full-scale results showed a
considerable improvement in the Protectors
performance in level ice although they admitted that
the original Protector was not particularly efficient.
They measured full-scale friction using two panels
installed on the bow of the Protector and obtained
somewhat scattered results as shown below:Low pressure panel, f = 0.16-0.26
High pressure panel, f = 0.05-0.13
Similar panels were installed on the
Polarstern (Schwarz et al., 1986), and results
(Schwarz, 1986[b]; Hoffmann, 1985) also show a
decrease in friction coefficient with increasing
normal force. Good correlation with model data, of
the performance of the new Protector bow, was
obtained with a model friction coefficient of 0.05 as
against the measured full-scale values shown above.
A major disadvantage of the bow was higher
slamming pressures. A similar disadvantage was
noted by Freitas and Nishizaki (1986) who tested an
ice class bulk carrier model with a Thyssen/Waas
bow, which otherwise showed considerable
improvement in icebreaking ability. This bow form
was fitted to the Mudyuq and results showed that in
snow-free ice, hull speed increased 50 to 100%
without the aid of the Jastram hull lubrication plant
(Varges, 1987, 1988). Improvements in turning
circle and in clearing of ice from a broken channel
were also reported, as well as agreement with model
tests. A series of comparison tests by Glen et al.
(1998) on four bows, one of which was a ThyssenWaas form, showed that while it was superior in
breaking level ice, this had little real significance on
the overall performance of a Navaids vessel in
service with the Canadian Coast Guard, which spent
a lot of its time in open water. For such a vessel a
conventional R-Class type bow was superior overall.
An interesting development in the mid-80s
was a full-scale towed resistance trial of the Mobile
Bay in uniform level ice (Zhan et al., 1987). In
principle, this parallels the open water trials of the
Greyhound (Froude, 1874) and Lucy Ashton (Denny,
1951). While such tests are clearly difficult to
perform, in theory they provide a direct measurement
of full-scale resistance. They also conducted full-

scale propulsion tests. They found the best fit to their


towed resistance results was with the equation (one of
15 equations that they analyzed):Ri
w gBh 2

where

= C o + C1

v2
gB

L
h

(17)

Co = 4.25
C1 = 3.96 X 10-5

Which implies a v2 dependence of resistance on


speed, as well as an h2 dependence. From their
propulsion data they determined a thrust deduction
fraction as a function of ice thickness, but as I have
commented in a discussion to their paper, their range
of thickness (and strengths) was so small, and the
normal errors associated with thrust and torque
measurements so large, that such a relationship is
difficult to justify. However, it is a valuable addition
to the literature and, hopefully, could be repeated in
the future with significantly different ice conditions,
for comparison.
Since 1990 the major development has
undoubtedly been that of using podded propellers in
ice with double acting tankers (DAT), which has
taken place principally in Finland (Juurmaa et al.,
2001) and appropriate for the Baltic Sea. Starting in
1990 with a 1.3 MW buoy tender, MV Seili, podded
propellers have been used in conjunction with
designs which allow the ship to go astern in heavy ice
and forward in open water and light ice. Full power
can be applied in either direction by rotating the
Azipod. Fig 14 shows the stern of the Seili with an
Azipod fitted.

Fig. 14. MV Seili, the first ship to be fitted with an


Azipod
The development has now progressed to a 16 MW
tanker with one Azipod unit, two of which were
recently (2003) delivered to Fortum Shipping by
Sumitomo Heavy Industries, for use in the Baltic.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

The idea was to design an efficient icebreaking stern


for the vessel, while keeping an efficient open water
bow. Fig. 15 shows one of the ships going astern in
ice during its ice trials.

Fig. 17. Stern model of the DAT in Fig. 15.


The most recent new icebreakers in North
America are the Nathaniel B. Palmer (1991)and the
USCGC Healy (2000) designed principally to be
Antarctic and Arctic research/supply ships. The
Healy is shown in Fig. 18.
Fig. 18. USCGC Healy entering St. Johns harbour.
Fig. 15. A 106,000-dwt Masa-Yards-developed DAT
crude carriers built by Sumitomo Heavy Industries in
2003.

Fig. 16. Bow form of the DAT in Fig. 15.


When entering a ridge field at slow or moderate
speed, a DAT vessel lets its pulling propeller chew
up the ridge and slowly pull the vessel through,
without any need for ramming. Whether this would
work on a massive arctic ridge without damage to the
propulsion unit seems unlikely, but the vessels are
well suited to Baltic ice conditions.

It has a conventional bow form with two


conventional propellers. A complete set of trials in
ice was conducted with this ship in 2000 with the
results published in the literature (POAC 2001). The
design icebreaking capability of the Healy was for
continuous icebreaking at 3 knots through 4.5 ft (1.37
m) of ice of 100 psi (690 kPa) strength. The fullscale trials were conducted in ice half this strength,
but by extrapolation from model-scale tests (Jones
and Moores, 2002) the ship was shown to meet this
requirement.
An icebreaker for the Great Lakes, GLIB, to
be named USCGC Mackinaw and scheduled for
delivery in October 2005 will have two Azipod units
of 3.3 MW each. The ship, shown in Fig. 19, is

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

icebergs are towed away by support ships, as shown


in Fig. 21.

Fig. 19. Profile of the GLIB to be delivered in 2005.


designed to break 32(0.82 m) of ice at 3 knots ahead
and 2 knots astern. In addition it should be very
maneuverable with the Azipod units, which can turn
360o. Full-scale trials of that ship will also be
conducted.

Fig. 21. Towing a medium sized iceberg off


Newfoundland
The oil is transported to market in ice strengthened
tankers. At IOT we have been conducting a major
research program into the impact of a ship with a
small iceberg, or bergy bit. The results remain
confidential for a little longer but Cumming et al.
(2001) has described the extensive model and fullscale experiments including full-scale impact tests
with the Terry Fox, shown in Fig. 22.

Fig. 22. The Terry Fox impacting a bergy bit.

Fig. 20. The Terra-Nova FPSO off Newfoundland


With the advent of the offshore oil industry
in Newfoundland two FPSOs have been built for the
ice infested waters, the Terra-Nova FPSO, shown in
Fig. 20, and the Sea Rose. These ships are not
icebreakers but are ice strengthened and can
withstand pack ice forces. They are designed to
disconnect if threatened by a large iceberg. Smaller

The last twenty years has seen advances in


ship-ice modeling techniques both experimental and
numerical. Jones et al (1989) has described the
different model ices in use throughout the world. In
short, large ice tanks have allowed model scales of
around 1:20 and at that scale the model ice properties
of strength, stiffness and brittleness are reproduced
remarkably well. Different tanks have used different
chemical dopants to give the best ice properties at
model scale, and at IOT we have always used a
combination of Ethylene Glycol (EG, 0.39%)
aliphatic detergent (AD, 0.036%) and sugar (S,
0.04%), thus giving us EGADS model ice. The
glycol acts like the salt in real sea ice forming brine,

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

or glycol, pockets on freezing, the detergent reduces


the surface tension at the growing interface allowing
more dopant to be included in the model ice, and the
sugar acts as a long chain molecule to keep the grain
size small as the ice grows. The resultant model ice
and its properties have been described by Timco
(1986) who concluded that it was a significant
improvement in model ice technology.
Another other major advance in the last
twenty years has been in numerical methods to
predict resistance in ice.
Valanto (2001) has
developed a 3-D numerical model of the icebreaking
process on the ship waterline, which predicts the
forces on the waterline. These were compared with
load panel measurements on the MS Uisko with good
agreement. He then calculated the resistance in ice
for several ships using his numerical model,
combined with a semi-empirical model of Lindqvist
(1989) for the underwater components of resistance,
and obtained good agreement with measured values,
as shown in Fig 22.

REFERENCES
Alexandrov, A.P. et al., 1958. The atomic icebreaker
Lenin. Proc. 2nd Int. Conf. on Peaceful Uses of
Atomic Energy, Geneva, U.N., New York, Vol. 8,
Nuclear Power Plants, Part I, p. 204-219.
Baker, D., and Nishizaki, R., 1986. The MV Arctic
new bow form and model testing. Trans. SNAME,
Vol. 94, p. 57-74.
Beach, C.W., Munn, A.M., and Reeves, H.E., 1895.
Strength of ice. The Technograph, Vol. 9, p.38-48.
Bulat V., 1982. The effect of snow cover on ice
resistance. The Naval Architect, Nov. 1982, p.
E.253.
Carter, D., 1983. Ship resistance to continuous
motion in level ice. Transportation Development
Centre, Transport Canada, Monteal, Canada, Report
No. TP3679E.
Churcher, A., Kolomojcev, A., and Hubbard, G.,
1984. Design of the icebreaking supply ship Robert
LeMeur. Marine Technology, Vol. 21, No. 2, p. 134146.
Corlett, E.C.B., and Snaith, G.R., 1964. Some
aspects of icebreaker design. Trans. RINA, Vol. 106,
No. 4, p. 389-413.

Fig. 22. Measured and computed resistance values in


level ice for the Otso-class icebreakers (Valanto,
2001).
In future, further developments in numerical methods
will continue to take place
SUMMARY
Enormous technological progress has been made in
the last 100 years from Eisbrecher I to Double Acting
Tankers. Ice will continue to be important factor for
oil exploration and production in certain offshore
areas as well as for marine transportation. Increased
tourist, as well as commercial, traffic in the Arctic
and Antarctic will bring demands for safer and more
efficient travel in such areas.
Modelling will
continue to improve with emphasis on numerical
simulations as well as physical modeling.

Crago, W.A., Dix, P.J., and German, J.G., 1971.


Model icebreaking experiments and their correlation
with full-scale data. Trans. RINA, Vol. 113, p. 83108.
Cumming, D., Gagnon, R.E., Derradji, A., and Jones,
S.J., 2001. Overview of research underway at the
institute for marine dynamics related to the
investigation of interactions between floating
structures and bergy bits. Proc. 6th Can. Marine
Hydrodynamics
and
Structures
Conference,
Vancouver.
Denny, Sir Maurice E., 1951. BSRA resistance
measurements on the Lucy Ashton, Part 1.: Full-scale
measurements. Trans. RINA, Vol. 93, International
Conference of Naval Architects and Marine
Engineers, p. 40-57.
Edwards, R.Y. Jr., et al., 1972. Full-scale and model
tests of a Great Lakes icebreaker. Trans. SNAME,
Vol. 80, p. 170-207.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Edwards, R.Y. Jr., et al., 1976. Influence of major


characteristics of icebreaker hulls on their powering
requirements and maneuverability in ice. Trans.
SNAME, Vol. 84, p. 364-407.

four bow designs for an icebreaking Navaids vessel.


Marine Technology, Vol. 35, No. 4, p. 200-218.
Gold, L.W., 1993. The Canadian Habbakuk project.
International Glaciological Society, Cambridge,
322pp.

Edwards, R.Y. Jr., Dunne, M.A., and Johnson, B.,


1981. Results of full-scale trials in ice of CCGS
Pierre Radisson. Proc. 6th STAR Symposium,
Ottawa, 17-19 June 1981, SNAME, New York, p.
291-310.

Hellmann, J-H., 1982. Model and full-scale tests in


ice with the icebreaker Max Waldeck.
Proc.
Intermaritec 82, Hamburg, p. 373-384.

Enkvist, E., 1972. On the ice resistance encountered


by ships operating in the continuous mode of
icebreaking. The Swedish Academy of Engineering
Sciences in Finland, Helsinki, Report No. 24, 181 pp.

Hoffmann, L., 1985. Impact forces and friction


coefficient on the forebody of the German polar
research vessel Polarstern. Proc. POAC 85, Vol. 3,
p. 1189-1202.

Enkvist. E., 1983.


A survey of experimental
indications of the relation between the submersion
and breaking components of level ice resistance to
ships. Proc. POAC 83, p. 484-493.

Jansson, J-E., 1956[a]. Ice-breakers and their design,


Pt.I. European Shipbuilding, No. 5, p. 112-128.

Enkvist, E., and Mustamki, E., 1986. Model and


full-scale tests with an innovative icebreaker bow.
Trans. SNAME, Vol. 94, p. 325-351.
Ettema, R., Stern, F. and Lazaro, J., 1987. Dynamics
of continuous-mode icebreaking by a polar-class
icebreaker hull. Iowa Institute of Hydraulic Research
Report No. 314, June 1987, 177pp.
Ferris, L.W., 1959. The proportions and forms of
icebreakers. Trans. SNAME, Vol. 67, p. 6-25.
Freitas, A. and Nishizaki, R.S., 1986. Model test of
an ice class bulk carrier with Thyssen/Waas bow
form. Trans. ASME, Vol. 108, p. 26-69.
Froude, W., 1874. On experiments with HMS
Greyhound. Trans. RINA, Vol. 15, p. 36-73.

Jansson, J-E., 1956[b]. Ice-breakers and their design,


Pt.II. European Shipbuilding, No. 6, p. 143-151.
Johansson, B.M., and Makinen, E., 1973.
Icebreaking model tests: systematic variation of bow
lines and main dimensions of hull forms suitable for
the Great Lakes. Marine Technology, Vol. 10, No. 3,
p. 236-243.
Johansson, B.M., and Revill, C.R., 1986. Future
icebreaker design.
Proc. International Polar
Transportation Conference (IPTC), Vancouver, 4-8
May 1986, D.F. Dickins Associates Ltd., Vol. 1, p.
169-200.
Johnson, H.F., 1946. Development of ice-breaking
vessels for the U.S. Coast Guard. Trans. SNAME,
Vol. 54, p. 112-151.

German, J.G., 1959. Design and construction of


icebreakers. Trans. SNAME, Vol. 67, p. 26-69.

Jones, S.J., Timco, G.W., Frederking, R., 1989. A


current view on sea ice modelling. Proceedings 22nd
ATTC, St. John's, Newfoundland, August 8-11, 1989,
National Research Council Canada, p. 114-120.

German, J.G., 1983. Hullform of icebreaker ships


background and progress. Proc. Annual Technical
Conference of CSSRA, Montreal, p. 49-69.

Jones, S.J. 1989. A review of ship performance in


level ice. Proc. 8th Int. OMAE Conf., ASME, New
York, Vol. IV, p. 325-342.

Ghonheim, G.A.M., et al., 1984. Global ship ice


impact forces determined from full-scale tests and
analytical modelling of the icebreakers Canmar
Kigoriak and Robert LeMeur. Trans. SNAME, Vol.
92, p. 253-282.

Jones, S.J. and Moores, C., 2002. Resistance tests in


ice with the USCGC Healy. IAHR 16th International
Symposium on Ice, Dunedin, New Zealand, Vol. 1, p.
410-415.

Glen, I.F., Jones, S.J., Paterson, R.B., Hardiman,


K.C. and Newbury, S., 1998. Comparative testing of

Juurmaa, k., and Segercrantz, H., 1981.


On
propulsion and its efficiency in ice. Proc. 6th STAR
Symposium, Ottawa, 17-19 June 1981, SNAME,
New York, p. 229-237.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Juurmaa, K., Mattson, T., and Wilkman, G., 2001.


The development of the new double acting ships for
ice operation. Proc. POAC 2001, NRC, Ottawa,
Vol.2, p. 719-726.

Michailidis, M., and Murdey, D.C., 1981.


Performance of CGS Franklin in Lake Melville 1980.
Proc. 6th STAR Symposium, Ottawa, 17-19 June
1981, SNAME, New York, p. 311-322.

Kari, A., 1921.


The design of icebreakers.
Shipbuilding and Shipping Record, No. 18, 22nd
December 1921, p. 802-804.

Milano, V.R., 1973. Ship resistance to continuous


motion in ice. Trans. SNAME, Vol. 81, p. 274-306.

Kashteljan, V.I., Poznyak, I.I., and Ryvlin, A.Ya.,


1968.
Ice resistance to motion of a ship.
Sudostroyeniye, Leningrad.

Milano, V.R., 1975. Variation of ship/ice parameters


on ship resistance to continuous motion in ice. Proc.
SNAME Ice Tech. Symposium, Montreal, Paper B,
30pp.

Kitagawa, H., et al., 1982. Vessel performance in ice


(Report No. 1) Abstract note of 40th General Meeting
of S.R.I., December 1982, p. 14-24.

Milano, V.R., 1980. A re-analysis of ship resistance


when in continuous motion through solid ice. Proc.
Intermaritec Symposium, Hamburg, p. 456-475.

Kitagawa, H., et al., 1983. Vessel performance in ice


(Report No. 2) Abstract note of 42nd General Meeting
of S.R.I., December 1983, p. 25-32.

Milano, V.R., 1982.


Correlation of analytical
prediction of ship resistance in ice with model and
full-scale test results. Proc. Intermaritec Symposium,
Hamburg, p. 350-372.

Kitagawa, H. et al., 1986.


A study of ship
performance in ice-covered waters (1st Report)
Effect of parallel mid-body. NRC, Institute for
Marine Dynamics, Report No. LM-AVR-14.
(Translated and revised version of Kitagawa et al.
1982, above ).

Narita, S., and Yamaguchi, M., 1981.


Some
experimental study on hull forms for the new
Japanese Antarctic research ship. Proc. 6th STAR
Symposium, SNAME, New York, p. 253-271.

Kloppenburg, M., 1975. Ice resistance of a cargo


vessel in the continuous mode of icebreaking. Proc.
Ice Tech 75, SNAME, New York, Paper L, 12 pp.

Newbury, G.S., and Williams, F.M., 1986. R-class


icebreaker model experiment results.
Ice
Technology, Computational Mechanics, SpringerVerlag, Berlin, p. 333-348.

Kotras, T.V., Baird, A.V., and Naegle, J.W., 1983.


Predicting ship performance in level ice. Trans.
SNAME, Vol. 91, p. 329-349.

Noble, P., and Bulat, V., 1981. Icebreaker bow


forms a parametric variation. Proc. 6th STAR
Symposium, SNAME, New York, p. 273-287.

Lewis, J.W., and Edwards, R.Y. Jr., 1970. Methods


for predicting icebreaking and ice resistance
characteristics of icebreakers. Trans. SNAME, Vol.
78, p. 213-249.

POAC 2001.
Proceedings of International
Conference on Port and Ocean Engineering under
Arctic Conditions, Ottawa, National Research
Council. Also see USCGC Healy Ice Trials 2000,
consolidated report, USCG-ELC-023-01-09.

Lewis, J.W., DeBord, F.W., and Bulat, V.A., 1982.


Resistance and propulsion of ice-worthy ships.
Trans. SNAME, Vol. 90, p. 249-276.
Lindqvist, G., 1989. A straightforward method for
calculation of ice resistance of ships. Proc. 10th Int.
Conference on Port and Ocean Engineering under
Arctic Conditions (POAC), Lulea, Sweden, p. 722735.
Mkinen, E., Lahti, A., and Rimppi, M., 1975.
Influence of friction on ice resistance; search for low
friction surfaces. Paper G, SNAME IceTech 75, 9-11
April, 1975, Montreal.

Poznyak, I.I., and Ionov, B.P., 1981. The division of


icebreaking resistance into components. Proc. 6th
STAR Symposium, SNAME, New York, p. 249-252.
Runeberg, R., 1888/89. On steamers for winter
navigation and ice-breaking. Proc. of Institution of
Civil Engineers, Vol. 97, Pt. III, p. 277-301.
Scarton, H.A., 1975. On the role of bow friction in
icebreaking. J. Ship Research, Vol. 19(1), p. 34-39.
Schwarz, J., 1977. New developments in modelling
ice problems.
Proc. POAC 77, St. Johns,
Newfoundland, p. 45-61.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Schwarz, J., Jochmann, P. and Hoffman, L., 1981.


Prediction of the icebreaking performance of the
German polar research vessel. Proc. 6th STAR
Symposium, Ottawa, 17-19 June 1981, SNAME,
New York, p. 239-248.

conditions. Proc. 7th IAHR Symposium on Ice, Vol.


2, p. 255-266.

Schwarz, J., 1986[a]. Some latest developments in


icebreaker technology. Journal of Energy Resources
Technology, Vol. 108, p. 161-167.

Vance, G.P., 1975. A scaling system for ships


modelled in ice.
Proc. SNAME Ice Tech.
Symposium, Montreal, Paper H1, 28pp.

Schwarz, J., 1986[b].


Advances in icebreaker
technology in West Germany. Proc. International
Polar Transportation Conference (IPTC), Vancouver,
4-8 May 1986, D.F. Dickins Associates, Vol. 1, p.
201-219.

Vance, G.P., 1980. Analysis of the performance of a


140-ft Great Lakes icebreaker : USCGC Katmai Bay.
U.S. Army CRREL Report 80-8, 28 pp.

Schwarz, J., et al., 1986. Erste Eisbrechtechnische


Expedition mit F.S. Polarstern, Report of
Hamburgische Schiffbau-Versuchsanstalt GmbH
(HSVA).

Valanto, P., 2001. The Resistance of Ships in Level


Ice, Trans. SNAME, Vol. 109, p. 53-83.

Vance, G.P., Goodwin, M.J. and Gracewski, A.S.,


1981. Full scale icebreaking test of the USCGC
Katmai Bay. Proc. 6th STAR Symposium, Ottawa,
17-19 June 1981, SNAME, New York, p. 323-344.
Varges, G., 1987. Advances in icebreaker design
the conversion of the Soviet polar icebreaker Mudyuq
into a Thyssen/Waas ship. Schiff und Hafen,
Kommandobrucke, Ht. 12, p. 20-30.

Shimanskii, Yu.A., 1938. Conditional standards of


ice qualities of a ship. Trans. Arctic Research
Institute, Northern Sea Route Administration
Publishing House, Vol. 130, Leningrad. Translation
T-381-01 by Engineering Consulting and Translation
Center (ECTC), P.O. Box 1377, Jackson Heights,
New York, NY 11372.

Varges, G., 1988. Revolution in polar shipping


Thyssen/Waas in Arctic ice.
Proc. 9th IAHR
Symposium on Ice, Sapporo, Japan, 23-27 August
1988, Vol. 1, p. 719-734.

Simonson, D.R., 1936. Bow characteristics of


icebreaking. Journal of American Society of Naval
Engineers, Vol. 48, No. 2, p. 249-254.

Vinogradov, I.V., 1946.


Vessels for Arctic
navigation (Icebreakers). Library of Congress, No.
(VM) 451.VS. Paraphrased in Ferris (1959).

Takekuma, K., and Kayo, Y., 1988. Comparative


resistance tests in ice on offshore structure models
and a ship model between two ice model basins.
Proc. 9th IAHR Symposium on Ice, Sapporo, Japan,
23-27 August 1988, Vol. 1, p. 570-579.

Virtanen et al., 1975. Great Lakes ore carrier series


ice resistance model tests draft variation. Wrtsil
Icebreaking Model Basin, Test Report No. A-34 to
U.S. Department of Commerce, Maritime
Administration.

Tatinclaux, J. C., 1984. Model tests on two models


of WTGB 140-ft icebreaker. U.S. Army CRREL
Report, 84-3, 17 pp.

Watson, A., 1959[a]. Operation of Department of


Transport icebreakers in Canada. Trans. SNAME,
Vol. 67, p. 140-161.

Tatinclaux, J.C., 1985. Propulsion tests in level ice


on a model of a 140-ft WTGB icebreaker. US Army
CRREL Report 85-4.

Watson, A., 1959[b]. The design and building of


icebreakers. London Transactions of the Institute of
Marine Engineers, Vol. 71, No. 2, p. 37-65.

Thiele, E.H., 1959. Technical aspects of icebreaker


operation. Trans. SNAME, Vol. 67, p. 162-184.

White, R.M., 1970.


Prediction of icebreaker
capability. Trans. RINA, Vol. 112, No. 2, p. 225251.

Timco, G.W., 1986. EG/AD/S: a new type of model


ice for refrigerated towing tanks. Cold Regions
Science and Technology, Vol. 12, p. 175-195.
Tronin, V.A., Malinovsky, V.A., and Sandakov, Yu.
A., 1984. Problems of river shipping in ice-bound

Zhan, P.B., Humphreys, D., and Phillips, L., 1987.


Full-scale towed resistance trials of the USCGC
Mobile Bay in uniform level ice. Trans. SNAME,
Vol. 95, p. 45-77.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Larry J. Doctors, The University of New South
Wales, Australia
Can you please clarify what property or
properties of the ice affect the size of the pieces that
break off? How does one control the creation of the
model ice to achieve the desired result?
AUTHORS REPLY
The piece size broken by a ship in model ice
has long been debated, but there are very few data.
Research done at our institute has suggested
(Newbury, 1989) that the piece size is indeed 3-5
times larger than it should be when compared with
limited full-scale data that is available. However,
some of this discrepancy might be due to the fact that
a 5 cm piece at full-scale would be measured, while
the same piece at a model scale of 1:20 might be
ignored, thus skewing the statistics of the piece size.
Undoubtedly the brittleness of the model ice
has a big effect on the piece size, and most model
ices are not as brittle as they should be at model
scale.
The only variables we can adjust are
temperature, and the chemical content of the water.
The EG/AD/S model ice, which we use at IOT, and
was mentioned in my talk, is more brittle than most
other ices, but even so it is less brittle than it should
be. Ice is a remarkably brittle substance considering
how close to its melting point it exists on earth and so
finding a perfect model ice will probably remain
impossible. We have reviewed the properties of the
different model ices in Jones et al (1989).

REFERENCES
Newbury, S.N., 1989. A preliminary investigation of
model ice failure pattern and piece size generated by
an icebreaker bow form. NRC/IMD Report LM1989-11, 4pp.
Jones, S.J., Timco, G.W., Frederking, R., 1989. A
current view on sea ice modelling. Proceedings 22nd
ATTC, St. John's, Newfoundland, August 8-11, 1989,
National Research Council Canada, p. 114-120.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics


St. Johns, Newfoundland and Labrador, CANADA, 8-13 August 2004

Nuclei Effects on Cavitation Inception and Noise


Georges L. Chahine
(DYNAFLOW, INC., U.S.A.)
community, but much more importantly to the
perceived impracticality of using the methods
developed with the existing design resources. This
has made the use of bubble dynamics seem
inconceivable but by experts.
Recently,
however,
there
have
been
tremendous advances in available computing
resources placed at the reach of non-experts.
Personal computers with phenomenal speed,
memory, and storage size, when compared to what
existed a decade ago, are now in the hands of most
engineers at a small fraction of the cost of an entry
computer a decade ago. This computer revolution
has definitely affected the operating procedures of
the designers. For instance, while a few years ago,
use of CFD viscous solvers by designers was out of
reach and only very simplified codes were used to
design and model rotating machinery, it is now
common to use repeatedly in-house or commercial
Navier Stokes solver CFD codes to seek better
solutions [11-13]. The challenge is thus presently
for the cavitation community to bring its techniques
to par with the single phase CFD progress. It is this
challenge that is been undertaken here and to
which we wish to significantly contribute.
In this paper, we discuss first the various
definitions of cavitation because of their significant
implications on modelling and then describe the
analytical and numerical tools that have become
available. We will try to convey the need to include
the presence of nuclei and nuclei dynamics in the
predictive tools for advanced designs. Some of
these tools are at the reach of all users and should
be adopted by the design community in
conjunction with the CFD tools presently used for
advanced design.

ABSTRACT
Cavitation is a problem of interaction between nuclei
and local pressure field variations including turbulent
oscillations and large scale pressure variations.
Various types of behaviors fundamentally depend on
the relative sizes of the nuclei and the length scales of
the pressure variations as well as the relative
importance of the bubble natural period of oscillation
and the characteristic time of the field pressure
variations. Ignoring this observation and basing
cavitation inception predictions on pressure
coefficients of the flow of the pure liquid, without
account for bubble dynamics could result in significant
errors in predictions. We present here a practical
method using a multi-bubble Surface Averaged
Pressure (DF-Multi-SAP) to simulate cavitation
inception and scaling, and connect this with more
precise 3D simulations.
INTRODUCTION
Cavitation and bubble dynamics have been the
subject of extensive research since the early works
of Besant in 1859 [1] and Lord Rayleigh in 1917 [2].
Thousands of papers and articles and several books
[e.g., 3-10] have been devoted to the subject.
Various aspects of the bubble dynamics have been
considered at length under various assumptions
and each contribution included one or several
physical phenomena such as inertia, interface
dynamics, gas diffusion, heat transfer, bubble
deformation, bubble-bubble interaction, electrical
charge effects, magnetic field effects, etc.
Unfortunately, very little of the resulting
knowledge has succeeded in crossing from the
fundamental research world to the applications
world, and it is uncommon to see bubble dynamics
analysis made or bubble dynamics computations
conducted for cavitation avoidance by the
hydrodynamics marine designer community, such
as propeller designers. This is due in part to the
failure of the scientific community to frame the
advances made in a format usable by the design

Definition(s) of Cavitation

Liquid phase only: Engineering definition


In the phase diagram of a substance the
curve which separates the liquid phase from the

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

through separation of the liquid molecules [3]. In


fact researchers agree that cavitation initiates at
weak spots of the liquid or nuclei. These are very
small microscopic bubbles or particles with gas
trapped in crevices in suspension in the liquid.
Several techniques have been used to measure
these nuclei distributions both in the ocean and in
laboratory cavitation channels. These include
Coulter counter, holography, light scattering
methods, cavitation susceptibility meters, and
acoustic methods [15-21]. Figure 1 shows typical
nuclei size distribution curves in cavitation water
tunnels and in the ocean [14, 22, 23]. The figure
shows the number density distribution, n, in m-4 ,
as a function of the bubble size, R. n(Rn) is the
number of nuclei bubbles in the range Rn to
Rn+dR. Distributions of the shape n( R) Rn4 are
usually reported.
Therefore, any fundamental analysis of
cavitation inception has to start from the
observation that, any real liquid contains nuclei
which when subjected to variations in the local
ambient pressure will respond dynamically by
oscillating and eventually growing explosively (i.e.
cavitating).

vapour phase defines the liquid vapour pressure


values at different temperatures. Any process
that raises the temperature or reduces the
pressure will result in a phase change from liquid
to vapour. Conventionally, boiling is defined as
the phase change resulting from raising the
temperature at ambient pressure, while cavitation
is the process inducing phase change at ambient
temperature through a pressure drop. This has
provided the following traditional cavitation
engineering definition: a liquid flow experiences
cavitation if the local pressure drops below the liquid
vapour pressure, Pv.
One root of the technology transfer problem
discussed above stems from this accepted
engineering definition of cavitation. Even though
this definition has allowed significant progress in
practical cavitation studies and design work, it is
responsible for a lack of further advance of the
technology, since it has been used at many
decision points to ignore bubble dynamics effects.
Indeed this definition assumes that the process
occurs in the regime where heat transfer is
negligible and where a large free surface is
present.
This over-simplification serves the
purpose in most engineering cases but could lead
to erroneous conclusions if used to explain or
model new complex problem areas. The
dangerous implication of this definition is that
understanding of the liquid one-phase flow only
is sufficient to predict and therefore avoid
cavitation.
We discuss in the following more advanced
definitions, which can help us to better
understand the scaling of the cavitating results
between laboratory and full scale. They should
also help cavitation test results comparison
between different testing facilities, and enable
making more accurate cavitation predictions.

Presence of cavitation nuclei

Figure 1. Nuclei Size distribution as measured in the


ocean and in the laboratory (from [22].

The above definition of cavitation inception


is only true in static conditions when the liquid is
in contact with its vapour through the presence of
a large free surface. For the more common
condition of a liquid in a flow, or in a rotating
machine, liquid vaporization can only occur
through the presence of micro free surfaces or
microbubbles, also called cavitation nuclei.
Indeed, a pure liquid free of nuclei can sustain
very large tensions, measured in the hundreds of
atmospheres, before a cavity can be generated

Cavitation inception in fact appears under


several forms, such as:
a. Explosive growth of individual travelling
bubbles,
b. Sudden appearance of transient cavities
or flashes on boundaries,
c. Sudden appearance of attached partial
cavities, or sheet cavities,

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

d. Appearance, growth, and collapse of


bubble clouds, behind attached cavities or
a vibrating surface.
e. Sudden appearance of cavitating rotating
filaments, or vortex cavitation.
Upon further analysis, all these forms can be
related initially to the explosive growth of preexisting cavities or nuclei in the liquid when
subjected to pressure drops generated by various
forms of local pressure disturbances1. These are
either imposed pressure variations, uniform
pressure drops due to local liquid accelerations, or
strongly non-uniform pressure fields due to
streamwise or transverse large vortical structures.
The presence of nuclei or weak spots in the liquid is
therefore essential for cavitation inception to occur
when the local pressure in the liquid drops below
some critical value, Pc , which we address next.

a gas partial pressure which varies with the bubble


volume. For quasi-steady equilibrium, Pg, as
considered in this section, the gas follows an
isothermal compression law, and is related to the
initial or reference values, Pgo , R0, and to the new
bubble radius R through:
3

R
Pg = PgO o .
(2)
R
The balance of pressures at the bubble wall
becomes:
3

R 2
L ( R ) =v +go o - ,
(3)
R R
where the notation PL(R) is meant to associate the
liquid pressure, PL to the bubble radius, R. An
understanding of the bubble static equilibrium can
be obtained by considering the curve; PL(R). As
illustrated in Figure 2, this curve has a minimum
below which there is no equilibrium bubble radius.
Only the left side branch of the curve corresponds
to a stable equilibrium.

Bubble Static Equilibrium


A first correction to the common engineering
definition of cavitation inception is based on
consideration of the static equilibrium of a bubble
nucleus. The nucleus is assumed to be spherical
and to contain non condensable gas of partial
pressure, Pg, and vapor of the liquid of partial
pressure, Pv. Therefore, at the bubble surface, the
balance between the internal pressure, the liquid
pressure, and surface tension can be written:
2
,
PL = Pv + Pg
(1)
R
where PL is the pressure in the liquid, is the
surface tension parameter, and R is the radius of
the bubble.
If the liquid ambient pressure changes very
slowly, the bubble radius will change accordingly
to adapt to the new balance. This is accompanied
with a modification of the pressure inside the
bubble. The vaporization of the liquid at the
bubble-liquid interface occurs very fast relative to
the time scale of the bubble dynamics, so that the
liquid and the vapor can be considered in
equilibrium at every instant, and the partial
pressure of the vapor in the bubble remains
constant. On the other hand, gas diffusion occurs at
a much longer time scale, so that the amount of gas
inside the bubble remains constant2. This results in

1.E+05
P= Pv
Ro=1 mic
Ro=2 mic
Ro=5 mic
Ro=10 mic
Critical Values

8.E+04
6.E+04

A m b ie n t P r e s s u r e , P a .

Stable
4.E+04
2.E+04
0.E+00
-2.E+04
-4.E+04

Unstable
-6.E+04
-8.E+04
-1.E+05
1

10

100

1000

Bubble Radius, m

Figure 2. Static equilibrium curves of spherical


bubbles and definition of critical curves. Example
given for bubbles of 1, 2, 5, and 10m initially at
equilibrium at a pressure of 1 atmosphere.
Solving for the minimum of PL(R) using
Equation (3) provides the values of the critical
pressure, Pc, and corresponding critical radius, rc,:
4
Pc = Pv
,
3rc
(4)
1
2.
3R03
2
rc =
PLo Pv +
.
Ro
2

This could be followed by extreme bubble deformation and


merger to result in the various cavitation forms.
2
More generally, both gas diffusion and vaporization can be
modeled and taken into account
3

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

addition to the temporal) strongly couple with the


actual bubble motion (i.e. position vs time) to
result in a driving force that depends on the
resonator reaction. This makes such a case much
more complex than what occurs for a travelling
bubble about a foil where, relatively speaking, the
position of the bubble is less coupled to its
dynamics.
The flow field pressure fluctuations have
various time scales: e.g. relatively long for
travelling cavitation bubbles over a blade or
captured in a vortical region flow, or very short
for cavitation in turbulent strongly sheared flow
regions. The amplitudes of these fluctuations and
the relationship between the various characteristic
times determine the potential for cavitation
inception.

If the pressure in the flow field drops below


the critical pressure an explosive bubble growth,
i.e. cavitation, is provoked. This provides an
improved definition for cavitation inception: a
liquid flow experiences cavitation if the local pressure
drops below the critical pressure, Pc.. The reason such
a definition has not been adopted is that the critical
pressure is not a liquid only definition and a
different value is obtained for each nucleus size. To
use it, one requires knowledge of the nuclei size
distribution in the liquid (which is ultimately
needed for any serious scale up study of
cavitation.) Expression (4) illustrates the fact that
the critical pressures are always lower than the
vapor pressure. Pc is close to Pv only for very
large initial nuclei sizes. This probably explains
why such a criterion has been ignored by the
practicians, the reasoning being that using Pv is on
the safe side. This reasoning, however, results in
large margins of safety. In addition, this cannot be
used to scale up experimental small scale tests,
since cavitation would occur when bubbles actually
grow explosively in the laboratory experiments and
not when p  Pv , but the scaling would assume
p = Pv .
Cavitation inception cannot be defined
accurately independent of the liquid bubble
population (sometimes characterized by liquid
strength [24]) and independent of the means of
cavitation detection. The cavitation inception is in
fact a complex dynamic interaction between the
nuclei and their surrounding pressure and velocity
fields; interaction that can be different between
small and large scales.
In addition, the
experimental means to detect and decide cavitation
inception (practical threshold used by the
experimentalists) will affect the results and could
be different between a laboratory experiment and
full scale.

Spherical Bubble Dynamics


The first improvement to the static
equilibrium analysis of a bubble nucleus is to
consider the nucleus dynamics when it is assumed
to conserve a spherical shape during its motion.
This has been extensively studied following the
original works of Rayleigh [1] and Plesset [25]. For
instance, if we limit the phenomena to be modelled
to inertia, small compressibility of the liquid,
compressibility of the bubble content, we obtain the
Gilmore [26] differential equation for the bubble
radius R(t). We modified this equation to account
for a slip velocity between the bubble and the host
liquid, and for the non-uniform pressure field
along the bubble surface [27]. The resulting
Surface-Averaged Pressure (SAP) equation applied
to Gilmores equation[ 27-28] becomes:
(1

R  3
R
1
R R d
) RR + (1 ) R = (1 + +
)

c
2
3c
c c dt

2
R ( u ub )

+
p
p
p
4
,
g
encounter
v

R
R
4

Dynamical Effects
When the pressure variations to which the
bubble is subjected are not slow compared to the
bubble response time, the nuclei cannot
instantaneously adapt to the new pressure, inertia
effects become important, and thus one needs to
consider the bubble dynamics equation. This is
the case for nuclei travelling through a rotating
machinery. The nuclei /bubbles then act as
resonators excited by the flow field temporal and
spatial variations. In the case of a vortical flow
field the strong spatial pressure gradients (in

(5)

where c is the sound speed, is the liquid


viscosity, u is the liquid convection velocity and u b
is the bubble travel velocity.
Equation (5) degenerates to the classical
Rayleigh-Plesset [10] equation for negligible
compressibility effects. If in addition, gas diffusion
effects are neglected and a polytropic law of gas
compression is assumed, the resulting modified
SAP equation becomes:

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

deformation, elongation, splitting, coalescence,


and non-spherical sound generation.
One such refinement, important for
propulsor studies, consists of considering the case
of bubbles captured on a vortex axis. The bubble
then elongates along the axis and may split into
two or more sub-bubbles, and/or form jets on the
axis. In order to investigate this behavior the
commercial
boundary
element
method
axisymmetric bubble dynamics code 2DYNAFS
[33-38] was exercised and was able to simulate
bubble dynamics through reentrant jet formation,
jet break through, and bubble splitting. The code
can handle as input vortex flow fields obtained
from CFD viscous computations or from
experimental measurements.
p

3k

3 2 1
R0

RR + R = pv + pg 0 Pencounter +


2
R

(6)
2

1 2 4 R ( u ub )
,

+
+
R
4
R
where k is the polytropic compression law
constant.
In the Surface-Averaged Pressure (SAP)
bubble dynamics equation, we have accounted for
a slip velocity between the bubble and the host
liquid, and for a non-uniform pressure field along
the bubble surface. In this SAP method the
definition of Pencounter as the average of the liquid
pressures over the bubble surface results in a
major improvement over the classical spherical
bubble model which uses the pressure at the
bubble center in its absence [27-29]. For instance,
a bubble does not always continuously grow once
it is captured by a vortex. Instead, it is subjected
to an increase in the average pressure once it
grows and this leads to a more realistic bubble
dynamics. In general, the gas pressure, pg, is
obtained from the solution of the gas diffusion
problem and the assumption that the gas is an
ideal gas [30].
The bubble trajectory is obtained using the
following motion equation [32]
du b 3
3
= P + C D ( u u b ) u u b

4
dt
(7)
3
+CL ( u u b ) u ub + ( u ub ) R ,
R
where the drag coefficient CD is given by an
empirical equation such as that of Haberman and
Morton [31]:
24
1.38
(1 + 0.197 Reb0.63 + 2.6 104 Reb
);
CD =
Reb
(8)
2 R u ub
Reb =
.

100

10-1

P acoustic / P amb

10-2

10

-3

10-4

10-5

10-6
0.75

0.8

0.85

0.9

0.95

1.05

/ -Cpmin

Figure 3: Illustration of the acoustic pressure emitted


by a bubble in a vortex field as a function of the
cavitation number. Note that the bubble behaviour
near and above the cavitation inception is quasispherical [35].
By simulating the dynamics behavior of a
bubble captured on a vortex axis under a
significant number of conditions using the
2DYNAFS, the followings conclusions illustrated
in Figure 3 were found [35,41]:

If the bubble is captured by the vortex far


upstream from the minimum pressure, it
remains spherical while oscillating at its natural
frequency.

When the bubble reaches the axis just


upstream of the minimum pressure, it develops
an axial jet on its downstream side which shoots
through the bubble moving in the upstream
direction. Even at this stage, the spherical
model provides a very good approximation
because the bubble is more or less spherical
until a thin jet develops on the axis.

Non-spherical Bubble Dynamics: Axisymmetry


Spherical bubble models, as briefly described
above, can be efficient tools for studying
cavitation inception, scaling, bubble entrainment,
and cavitation noise. They can become more
powerful if they are provided with further
intelligence based on more precise nonspherical models which account for bubble
behavior near boundaries, in pressure gradients,
and in high shear regions, resulting in bubble

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

jets formed after the splitting.


This is an
important conclusion that has been preliminarily
confirmed experimentally [35,38].

The bubble behavior becomes highly nonspherical once it passes the minimum pressure
location. It elongates significantly and can
reach a length to radius ratio that can exceed 10.
The bubble then splits into two or more
daughter bubbles emitting a strong pressure
spike followed later by other strong pressure
signals when daughter bubbles collapse. Two
axial jets originating from the split and a strong
pressure signal during the formation of the jets
are observed.

Three types of tests were conducted and are


still on-going: spark generated bubbles, laser
generated bubbles, and electrolysis bubbles
injected in vortex lines. Figure 4 shows high speed
photography and acoustic signals of bubble
splitting between two rigid walls. A small but
distinct pressure spikes is formed at splitting
followed but a more significant spike during the
collapse of the sub-bubbles. The second set of
experiments was conducted in a vortex tube,
where bubbles generated by electrolysis were
injected and observed once captured by the vortex
line.
Figure 5 shows the elongated bubble
dynamics and the corresponding signals
measured by a hydrophone [38]. The third set of
experiments was conducted at the University of
Michigan [39] using laser induced bubbles (in the
vortex and far upstream) and the flow field of a
tip vortex behind a foil. Comparisons between the
observations and the 2DYNAFS simulations
showed very good correspondence as shown in
Figure 6.

Signal from
the collapse

10

Pressure Signal (mV)

-5

-10
0.02

0.025

0.03

0.035

Time (sec)

Figure 4. High speed photos of spark-generated


bubble collapsing between two solid walls, and
resulting acoustic signal indicating peak
signal at splitting and subsequent sub-bubbles
collapse [35]

14000
12000
10000
8000
6000
4000
2000
0
0.75

1.25

1.5

1.75

2.25

2.5

Figure 6. Bubble behaviour in a vortex flow field:


Rc = 4.51mm , = 0.2123 m 2 /s , = 1.72;

2.75

-2000
-4000
-6000

U = 10 m/s; R0 = 750 m . Three-dimensional


view of the bubble just before splitting predicted
by 2DYNAFS and observed at University of
Michigan using laser-induced bubbles at
the center of the vortex [39].

t (m s)

Figure 5. High speed photos of an electrolysis


bubble captured in a line vortex, and resulting
acoustic
signal
indicating
peak
signals
at splitting and collapse [40].

Bubble splitting criteria

Experimental Verification

A large series of computer simulations of


axisymmetric bubbles captured in a vortex
indicated some definite trends, which can be used
in a predictive model [34-38]:

This behavior supports the hypothesis that


the noise at the inception of the vortex cavitation
may originate from bubble splitting and/or the

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

An explosively growing bubble splits into


two sub-bubbles after it reaches its maximum
volume, (equivalent radius, Rmax) and then drops
to 0.95 Rmax.
The two resulting sub-bubbles have the
following equivalent radii: 0.90 Rmax and 0.52 Rmax.
The locations of the two sub-bubbles after
the splitting are at -0.95 Rmax and 4.18 Rmax.
The pressure generated by the subsequent
formation of reentrant jets in the sub-bubbles can
be approximated by a function of [41]:
Since the noise associated with the jet
formation appears to be much higher than the
pressure signal from the collapse of a spherical
bubble, it is desirable to include the splitting and
the associated jet noise in simulations with
multiple bubble nuclei.

various stages of the interaction between a bubble


and a tip vortex flow.

Validation of the SAP model


In order to evaluate the various models, we
combined the SAP spherical model and the twoway interaction non-spherical bubble dynamics
model to predict tip vortex cavitation inception
for a tip vortex flow generated by a finite-span
elliptic hydrofoil [28].

Re=2.88x106
R0=50m
=2.56
Re=2.88x106
R0=50m
=2.50

Fully Non-spherical Bubble Dynamics


In order to study the full 3D interaction
between a bubble and a complex flow field, two
methods were developed. The first, using the
commercial boundary element code, 3DYNAFS
[42], enables study of full bubble deformations
during capture but neglects the effects that the
bubble may have on the underlying flow field.
The second method accounts for the full two-way
bubble/flow field interaction, and considers
viscous interaction. This model is embedded in an
Unsteady Reynolds-Averaged Navier-Stokes
code, DF-UNCLE3, with appropriate free surface
boundary conditions and a moving Chimera grid
scheme [28,42]. This full two-way interaction nonspherical bubble dynamics model has been
successfully validated in simple cases by
comparing the results with reference results
obtained from the Rayleigh-Plesset equation and
3DYNAFS for bubble dynamics in an infinite
medium both with and without gravity [43].
As an illustration Figure 7 shows results of a
bubble interacting with the tip vortex of an
elliptical foil. The bubble elongates once it is
captured, and depending on the cavitation
number, either forms a reentrant jet directed
upstream or splits into two sub-bubbles. When
two-way interaction is taken into account further
smoothing of the bubble surface is exerted by
viscosity resulting in a more distorted but overall
more rounded bubble. Figure 8 illustrates the

One-Way
Interaction

Re=2.88x106
R0=50m
=2.50

Two-Way
Interaction

Figure 7. Bubble behavior in a vortex flow field [28].

s.
re s
r P
.
n te re s s
P
cou
E n u s ti c
o
Ac

Figure 8. Sketch of the successive phases of a


bubble behavior in a vortex flow field.
The flow field was obtained by a RANS
computation and provided the velocity and
pressure fields for all compared models: the
classical spherical model; the SAP model, the oneway interaction model where the bubble
deformed and evolved in the vortex field but did
not modify it, and finally the fully coupled 3D

DF_UNCLE is a DYNAFLOW modified version of UNCLE


developed by Mississippi State University
7

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

model due to the interaction between the


bubble and the vortex flow field.

model in which unsteady viscous computations


included modification of the flow field by the
presence of the bubbles.
R e = 2 .8 8 x 1 0 6 R 0 = 5 0 m = 2 .5

S p h e r ic a l
S p h e r ic a l
O ne -W a y
T w o -W a y

0 .0 0 7

Conventional
Spherical

0 .0 0 6

Differences between the one-way and two


interaction models exist but are not major.
Using the Surface Averaged Pressure (SAP)
scheme significantly improves the prediction
of bubble volume variations and cavitation
inception. SAP appears to offer a very good
approximation of the full two-way
interaction model.

M odel N o S AP
M o d e l W it h S A P
N o n - S p h e ric a l M o d e l
N o n - S p h e r ic a l M o d e l

R (m)

0 .0 0 5
0 .0 0 4
0 .0 0 3

1-Way 3D

0 .0 0 2

2-Way 3D
0 .0 0 1

SAP Spherical
0 .0 4

0 .0 5

0 .0 6

0 .0 7

0 .0 8

Nuclei Distribution

0 .0 9

0 .1

T im e ( s e c )

Release
Area

Figure 9. Comparison of the bubble radius versus time


for the spherical models (the conventional and the SAP
model) and the 3D 1-way and 2-way UnRANS
computations [28].

Re=2.88x10

l = U t
t Investigation time

R 0=50m C 0 =1m

0.008
0.007

Conventional
Spherical

Spherical Model No SAP


Spherical Model With SAP
One-Way Non-Spherical Model
Two-way Non-Spherical Model

R0 = 5-50 micron Total nuclei=568


Void Fraction = 3.4x10

0.005

-8

250

1-Way 3D
200

0.004
0.003

Number of Nuclei (N/m^3)

Rmax(m)

0.006

2-Way 3D
SAP Spherical

0.002
0.001
0
2.4

2.5

2.6

Bubble Released

150

100

50

2.7

Cavitation no.
0
5

10

15

20

30

40

50

Nucleus Size (micron)

Figure 10. Comparison of the maximum bubble radius


vs. cavitation number between the spherical models
(the conventional and the SAP model) and the 3D 1way and 2-way UnRANS computations [28].

Figure 11. Illustration of the fictitious volume feeding


the inlet to the nuclei tracking computational domain
(or release area) and example of resulting nuclei size
distribution satisfying a given nuclei density
distribution function.

Comparisons between the various models


of the resulting bubble dynamics history, and of
the cavitation inception values obtained from
many tested conditions, reveal the following
conclusions, illustrated in Figures 9 and 10:

Modelling of a Real Nuclei field

In order to simulate the water conditions with a


known distribution of nuclei of various sizes, as
illustrated in Figure 11. the nuclei are considered
to be distributed randomly in a fictitious supply
volume feeding the inlet surface or release area of
the computational domain. The fictitious volume
size is determined by the sought physical

The bubble volume variations obtained from


the full two-way interaction model deviate
significantly from the classical spherical

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

affect bubble dynamics in a way as to explain the


above observations.
To address this issue we exercised the
methods described above to study the effect of
vortex/vortex interaction on bubble dynamics
and cavitation noise [45,46]. The liquid phase
flow was solved by direct numerical simulation of
the Navier-Stokes equations and was coupled
with the SAP spherical bubble dynamics model to
track the evolution of the bubbles at each time
step.

duration of the simulation and the characteristic


velocity in the release area as illustrated in Figure
11. The liquid considered has a known nuclei size
density distribution function, n(R), which can be
obtained from experimental measurements [1621] and can be expressed as a discrete distribution
of M selected nuclei sizes. Thus, the total void
fraction, , in the liquid can be obtained by
M
4 Ri3
= Ni
,
(10)
3
i =1
where Ni is the discrete number of nuclei of
radius Ri used in the computations. The position
and thus timing of nuclei released in the flow
field are obtained using random distribution
functions, always ensuring that the local and
overall void fraction satisfy the nuclei size
distribution function.

Vortex/Vortex Interaction and Inception

Tip Leakage
Vortex

Recent experiments on ducted propellers


[44] have shown a cavitation inception value and
a cavitation inception location which were very
different than those predicted by several state of
the art CFD RANS codes. The most disturbing
conclusion made from this comparison was that
cavitation inception does not occur in the
minimum pressure region, which would
contradict our understanding of cavitation
inception as the explosive growth of nuclei in low
pressure regions.
One hypothesis for explaining this
doubtful conclusion was that in both the
experiments and the simulations unsteady effects
were not accounted for, with the RANS solutions
smearing out the computed fluctuations and the
experimental measurement techniques filtering
them through time averaging. The reason this
effect was enhanced in the concerned experiment
is that the flow field was inherently unsteady and
thus significantly affected bubble dynamics in a
complex fashion. Indeed, in the considered
ducted propeller there is strong interaction
between a tip-leakage vortex and a trailing-edge
vortex as illustrated in Figure 12 at a low
cavitation
number,
showing
cavitation
development and interaction between the two
structures. This evolving vorticity may cause
early cavitation wherever the two vortices
strongly interact. We therefore set out to analyze
whether unsteady vortex/vortex interactions

Trailing Edge
Vortex

Figure 12. Advanced cavitation on Propeller 5206


visualizing both the tip leakage vortex and the trailing
edge vortex (taken from [44])
Canonical problem
A canonical problem was considered first.
Bubble dynamics in the flow field of two-unequal
co-rotating vortices with different configurations
was considered and resulted in the following
conclusions [45].
A stronger interaction between the two
vortices was observed when the strengths of the
two vortices were closer.
The minimum pressure value and
location is strongly affected by the two-vortices
interaction. It could occur at, before, or after the
two vortices have completely merged depending
on the relative strength of the two vortices (see
Figure 13).
The pressure reaches its minimum when
the vorticity of the weaker vortex is spread and
sucked into the stronger vortex. This also results
in an acceleration of the flow and leads to a
maximum streamwise velocity at the vortex center.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

minimum of the pressure is also seen at 0.5 chord


length and has value of -10.8.

The shape, size and location of study of


the window of opportunity, i.e. area in the inlet
to the computational domain from which emitted
nuclei are captured in the vortices, are highly
dependent on the relative strength of the two
vortices and on the nuclei sizes. A large size of
window of opportunity was found for the
stronger interaction case and for larger nuclei.
The unsteady flow resulting from the
interaction of the two vortices may results in
some nuclei initially starting to be entrapped by
one vortex to be ejected by the other during the
merging process (see Figure 13).

2=1/3

~ 0.6 C0

2=1/2.5

Propulsor study
The same approach as discussed above was
applied to the David Taylor Propeller 5206 [44]
shown in Figure 12. Three RANS codes have been
used by three groups to simulate numerically
cavitation inception on this propulsor [46-49]. All
three codes followed the simple engineering
criterion for cavitation inception, i = c p , and
min

~ 0.3 C0

2=1/2

Figure 13. Effect of the strength of two vortices on the


location and intensity of the minimum pressure [45].

gave close inception values: 6.5 < < 8, at a


location 0.1 chord length downstream of the trailing
edge of the blade. The experiments however,
showed a much higher inception value,  11,
and, more disturbing, at a location much further
downstream, 0.5 chord length!, much far away for
the c p location.
min

To improve the numerical solution from


these RANS computations, we considered a
reduced computational domain behind the
trailing edge of the propulsor blade that
encompassed only the region of interaction of the
two vortices. The RANS solution of Yang [49]
provided the initial conditions for the grid points
of the reduced domain, and the boundary
conditions everywhere but at the downstream
end of the domain, where an extrapolation
scheme was used [46]. As in the previous section,
a direct numerical simulation (DNS) of the Navier
Stokes equations was performed, for a set of
increasingly finer grids.
Figure 15 shows a comparison of the
resulting pressure coefficient, Cp, along the vortex
center line between the RANS computation [49]
and the DNS computations for three different
grids. As the gridding is refined, Cpmin converge
to about -11 at a location 0.34 chord length
downstream from the tip trailing edge. Another

Figure 14. Interaction between two vortices resulting


in ejection of initially trapped nuclei out of the main
vortex [45].
The two co-rotating vortices periodically
approach each other during the vortex merger. As
they move closer, the flow in the axial direction is
accelerated and results in a decreased pressure in
the vortex center. Figure 16 indicates that this
pressure drop is directly connected to the
enhancement of the axial velocity to a maximum
value by the merger.
A bubble population was allowed to
propagate through the propeller flow field and
the resulting dynamic cavitation inception was
studied using both 3D bubble dynamics and SAP
[46]. Figure 17 illustrate where the cavitation
event occurs in the flow field, the bubble
trajectory and size variations are plotted with the

10

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

propulsor blade and iso-pressure surface. It is


seen that the cavitation event occurs at a location
very close to the experimental observation, since
the bubble grow to their maximum size near a
location 0.5 chord length downstream of the tip
trailing edge.

R0=20m, =10.75

R0=20m, =10.85

s/C=0

s/C=0

Cp=-5.6

Pressure coefficient along vortex center

Cp=-10.9

Cp=-5.6

-4

s/C=0.5
DNS 61x61 grid
DNS121x121 grid
DNS 181x181 grid
RANS

Cp

-6

blow up

-8

s/C=0.5

blow up

Figure 17. Bubble trajectories and size variations


during bubble capture by the two-vortex system [46].

-10

0.1

0.2

0.3

s/C

0.4

0.5

0.6

0.7

Figure 15. Pressure coefficient at various distances


from the propeller blade as computed by RANS and by
the direct Navier Stokes solution with an increasing
number of grids [46].

3.5

-8

Cp
Axial velocity

-9

3.3

Vs

Cp

3.4

3.2

-10

3.1

0.1

0.2

0.3

0.4

s/C

0.5

0.6

0.7

Figure 16. Pressure coefficient and axial velocity at


various downstream distances from the
propeller blade [46].

Figure 18. Results of SAP simulation of cavitation


with gas diffusion on the Prop5168 propeller. The
bubbles in Section A are to scale. The bubbles
upstream and downstream were magnified by 5 to
become visible.

Conclusions
Difficulties in considering real fluid effects
have led the user community to select a liquid
only simple engineering definition of cavitation
inception as the basis for cavitation predictions
and scaling. While this has served the community
very well for decades, advances in silencing and
detection has made such a definition unsuitable
for advanced designs.

One has then to resort to the more basic


definition of cavitation as that of the explosive
growth of initially microscopic nuclei in a liquid,
resulting in visible bubbles (optical criterion) or
in detectable emitted sound signals (sub-visual
cavitation and acoustical criterion). The required
simulations of the nuclei behavior in complex
flow fields and turbulent structures were
previously out of reach of the community and

11

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

3.

thus were out the question. However, with the


advent of desktop high speed computers and with
the development of advanced computational
techniques, this is now within the reach of
designers who are increasingly using CFD (such
as RANS) to select their designs. In this
communication, we have proposed a practical
method to actually conduct bubble dynamics
numerical experiments as in the real flow field.
This allows actual nuclei fields to interact with the
computed flow field. This method, which we
have successfully used with RANS solvers, a DNS
solver, could be used with experimentally
measured flow fields and become a design tool
for cavitation avoidance.
We have shown some simulations in the
body of the communication for cavitation
inception. In fact, the method has also been use
more recently to simulate advanced cavitation
such as shown in Figure 18 and could prove with
further development to be a powerful design and
scaling tool. One of its major strengths is that it
allows the engineer to reproduce and mimic the
actual experimental procedures. For instance,
both acoustical and optical criteria of cavitation
inception could be measured. Concerning the
acoustical criteria, the technique provides in
addition to amplitude of measured signals, the
number of events per second, and the spectra of
the sound generated, which both could be used to
simulate detection. The engineer could therefore
utilize the same criteria and tools as used in the
real life experiments to conduct the predictions.

4.

5.
6.

7.

8.
9.

10.

11.

12.

Acknowledgment

13.

This work could not have been done


without the sustained support of ONR, Dr. KiHan Kim monitor, and the significant
contribution of many DF colleagues, most
particularly Dr. Chao-Tsung Hsiao and Dr. JinKeun Choi who developed and conducted most
of the numerical simulations presented here.

14.

15.

References
1.

2.

Besant,
W.H.
Hydrostatics
and
hydrodynamics, Cambridge University
Press, London, Art. 158, 1859.
Lord Rayleigh, On the pressure developed in
a liquid during collapse of a spherical
cavity, Phil. Mag; 34:94-98, 1917.

16.

Knapp R.T., Daily, J.W., and Hammitt F.G.,


Cavitation, McGraw Hill Book Co., NY,
1970.
Hammitt F.G., Cavitation and Multiphase
Flow
Phenomena,
McGraw-Hill
International Book Co., NY,1980.
Young, F.R. Cavitation. McGraw Hill Book
Co., NY, 1989.
Franc, J. P., Avellan, F., Belhadji, B., Billard,
J.Y., Brianon-Marjollet, Frchou, D.,
Fruman, D.H., Karimi, A., Kueny, J.L.,
Michel, J.M., La Cavitation. Mcanismes
Physiques
et
Aspects
Industrielles,
Collection Grenoble Sciences, Presses
Universitaires de Grenoble, 1995.
Brennen, C.E., Cavitation and Bubble
Dynamics, Oxford Engineering Sciences
Series 44, Oxford University Press, 1995.
Isay, W. H. Kavitation, Schiffahrts-Verlag
Hansa C. Shroedter & co. Hamburg, 1981
Leighton, T.G.
The Acoustic Bubble,
Academic Press, 1994.
Plesset, M. S., Bubble Dynamics in Cavitation
in Real Liquids, Editor Robert Davies, Elsevier
Publishing Company, pp. 1-17, 1964
Hsiao, C.-T. and Pauley, L.L., Numerical
Computation of the Tip Vortex Flow Generated
by a Marine Propeller, ASME Journal of Fluids
Engineering, Vol. 121, No. 3, pp. 638-645,
1999.
Chen, B., and Stern, F., Computational Fluid
Dynamics of Four-Quadrant Marine-Propulsor
Flow, ASME Symposium on Advances in
Numerical Modeling of Aerodynamics and
Hydrodynamics
in
Turbomachinery,
Washington, D.C., 1998.
Gorski, J,J., Present State of Numerical Ship
Hydrodynamics and Validation Experiments,
Journal of Offshore Mechanics and Arctic
Engineering -- May 2002 -- Volume 124, Issue
2, pp. 74-80.
Wu J., Whitecaps, Bubbles and Spray, in
Oceanic Whitecaps, Editors Monahan, E.C.
and Niocaill, G. M., D. Reidel Publishing
Company, pp. 113-123, 1986.
F. MacIntyre, On reconciling optical and
acoustical bubble spectra in the mixed
layer, in Oceanic Whitecaps, edited by E.C.
Monahan and G. Macniocaill, Reidell, New
York, 75-94, 1986.
D.M. Oldenziel, A new instrument in
cavitation
research:
the
cavitation
susceptibility meter, Journal of Fluids
Engineering., 104, 136-142, 1982.

12

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

28. Hsiao, C.-T., Chahine, G. L., Prediction of


Vortex Cavitation Inception Using Coupled
Spherical and Non-Spherical Models and
Navier-Stokes Computations, Journal of
Marine Science and Technology, Vol. 8, No.
3, pp. 99-108, 2004.
29. Hsiao, C.-T., Chahine, G. L., Scaling of Tip
Vortex Cavitation Inception Noise with a
Statistic
Bubble
Dynamics
Model
Accounting for Nuclei Size Distribution, to
appear ASME Journal of Fluid Engineering
2005.
30. G.L. Chahine and K.M. Kalumuck, The
Influence of Gas Diffusion on the Growth of
a Bubble Cloud, ASME Cavitation and
Multiphase Flow Forum, Cincinnati, Ohio,
Vol. 50, pp. 17-21, June 1987.
31. Haberman, W.L., Morton, R.K., An
Experimental Investigation of the Drag and
Shape of Air Bubbles Rising in Various
Liquids, Report 802, DTMB, 1953.
32. Johnson V.E., Hsieh, T., The Influence of the
Trajectories of Gas Nuclei on Cavitation
Inception, Sixth Symposium on Naval
Hydrodynamics, pp. 163-179, 1966.
33. Chahine, G.L., Choi. J.-K., 2DYNAFSTMAxisymmetric Boundary Element Method
for
Bubble
and
Free
Surface
Dynamics, DYNAFLOW, INC. User Manual
7-026, February 2004.
34. Choi, J.-K and Chahine, G. L., Non-Spherical
Bubble Behavior in Vortex Flow Fields,
Computational Mechanics, Vol. 32, No. 4-6,
pp.281-290, December 2003 (also in IABEM
2002 Conference, Univ. of Texas at Austin,
TX, May 2002).
35. Choi, J.-K and Chahine, G. L., Noise due to
Extreme Bubble Deformation near Inception
of Tip Vortex Cavitation, FEDSM03,
International Symposium on Cavitation
Inception, 4th ASME/JSME Joint Fluids
Eng. Conference, Honolulu, Hawaii, July
2003.
36. Choi, J.-K and Chahine, G. L., A Numerical
Study on the Bubble Noise and the Tip
Vortex Cavitation Inception, Eighth
International Conference on Numerical
Ship Hydrodynamics, Busan, Korea,
September 2003.
37. Choi, J.-K, Chahine, G. L., and Hsiao C.-T.,
Characteristics of Bubble Splitting in a Tip
Vortex Flow Fifth International Symposium

17. M. L. Billet, Cavitation nuclei measurements


a review, ASME Cavitation and
Multiphase Flow Forum, FED-vol 23, June
1985.
18. N. Breitz and H. Medwin, Instrumentation
for in situ acoustical measurements of
bubble spectra under breaking waves, J.
Acoust. Soc. Am., 86, 739-743, 1989.
19. R. Duraiswami, S. Prabhukumar and G.L.
Chahine, Bubble Counting Using an
Inverse Acoustic Scattering Method,
Journal of the Acoustical Society of
America, Vol. 105, No. 5, November 1998.
20. Chahine, G.L., Kalumuck, K.M., Cheng, L.-Y.,
and G. Frederick, 2001, Validation of
Bubble Distribution Measurements of the
ABS Acoustic Bubble Spectrometer with
High Speed Video Photography, 4th
International Symposium on Cavitation,
California
Institute
of
Technology,
Pasadena, CA.
21. Chahine, G.L. and Kalumuck, K.M.,
Development of a Near Real-Time
Instrument for Nuclear Measurement: The
ABS Acoustic Bubble Spectrometer, 4th
ASME-JSME Joint Fluids Engineering
Conference, Honolulu, HI, July 6-10/2003
22. Franklin, R.E., 1992, A note on the Radius
Distribution Function for Microbubbles of
Gas in Water, ASME Cavitation and
Mutliphase Flow Forum, FED-Vol. 135,
pp.77-85.
23. Katz, J. and Acosta, A. Observations of
nuclei in cavitating flows, in Mechanics
and Physics of Bubbles in Liquids, L. van
Wijngaarden, editor, Martinus Nijhoff
Publishers, pp. 123-133, 1982.
24. Arndt, R.E., Keller, A.P., Water Quality
Effects on Cavitation Inception in a Trailing
Vortex,
ASME
Journal
of
Fluid
Engineering, Vol. 114, pp. 430-438, 1992.
25. Plesset, M. S., Dynamics of Cavitation
Bubbles, Journal of Applied Mechanics,
Vol. 16, 1948, pp. 228-231, 1948.
26. Gilmore, F. R., The growth and collapse of a
spherical bubble in a viscous compressible
liquid, California Institute of Technology,
Hydro. Lab. Rep, 26-4, 1952.
27. Hsiao, C.-T., Chahine, G. L., Liu, H.-L.,
Scaling Effects on Prediction of Cavitation
Inception in a Line Vortex Flow, Journal of
Fluid Engineering, Vol. 125, pp.53-60, 2003.

13

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

38.

39.

40.

41.

42.

43.

44.

45.

46.

47. Brewer, W.H., Marcum, D.L., Jessup, S.D.,


Chesnakas, C., Hyams, D.G., Sreenivas, K.,
An Unstructured RANS Study of TipLeakage Vortex Cavitation Inception,
Proceedings of the ASME Symposium on
Cavitation Inception, FEDSM2003-45311,
Honolulu, Hawaii, July 6-10, 2003.
48. Kim, J., Sub-Visual Cavitation and Acoustic
Modeling for Ducted Marine Propulsor,
Ph.D. Thesis, 2002, Department of
Mechanical Engineering, The University of
Iowa, Adviser F. Stern.
49. Yang, C.I., Jiang, M., Chesnakas, C.J., and
Jessup, S.D., 2003, "Numerical Simulation of
Tip Vortices of Ducted-Rotor", NSWCCD50-TR-2003/46

on Cavitation, CAV2003, Osaka, Japan,


November 2003.
Choi, J.-K., Chahine, G.L., Noise due to
Extreme Bubble Deformation near Inception
of Tip Vortex Cavitation, Physics of Fluids,
Vol.16, No.7, July 2004.
Rebow, M., Choi, J., Choi, J.-K., Chahine, G.L.,
and Ceccio, S.L., Experimental Validation
of BEM Code Analysis of Bubble Splitting
in a Tip Vortex Flow, 11th International
Symposium on Flow Visualization, Notre
Dame, Indiana, Aug. 9-12, 2004.
Chahine, G.L., Hsiao, C.-T., and Delfino, C.
Cavitation Inception and Noise Scaling
Using Measured Bubble Size Distribution,
Dynaflow Technical Report. 2M2014-1DTRC, July, 2004.
Choi, J.-K., Hsiao, C.-T., and Chahine, G.L.,
Tip Vortex Cavitation Inception Study
Using the Surface Averaged Pressure (SAP)
Model Combined with a Bubble Splitting
Model, 25th Symposium on Naval
Hydrodynamics, St. Johns, Newfoundland
and Labrador, Canada, Aug. 8-13, 2004.
Chahine, G.L., Hsiao, C.-T., 3DYNAFSTM - A
Three dimensional Free Surface and Bubble
Dynamic Code, DYNAFLOW, INC. User
Manual 7-038, October 2004
Hsiao, C.-T., Chahine, G.L., Numerical
Simulation of Bubble Dynamics in a Vortex
Flow Using Navier-Stokes Computations
and Moving Chimera Grid Scheme, 4th
International Symposium on Cavitation
CAV2001, Pasadena, CA, June 20-23, 2001.
Chesnakas, C. J., Jessup, S. D., Tip-Vortex
Induced Cavitation on a Ducted Propulsor,
Fifth
International
Symposium
on
Cavitation
CAV2003,
Osaka,
Japan,
November 1-4, 2003.
Hsiao, C.-T. and Chahine, G. L., Effect of
Vortex/Vortex Interaction on Bubble
Dynamics and Cavitation Noise, Fifth
International Symposium on Cavitation
CAV2003, Osaka, Japan, November 1-4,
2003.
Hsiao, C.-T., Chahine, G. L., Numerical
Study of Cavitation Inception due to
Vortex/Vortex Interaction in a Ducted
Propulsor, 25th Symposium on Naval
Hydrodynamics, St. John, Newfoundland
and Labrador, Canada, August 8-13, 2004.

14

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Martin Renilson
QinetiQ, United Kingdom
Thank you for a very interesting paper. Can
you please explain what influences the frequency of
oscillation of the spherical bubble, and how this
scales?
AUTHORS REPLY
Thank you for your comment.
The
frequency of the spherical bubble is proportional to
the inverse of its radius and the square root of the
local pressure. Simplified scaling would follow from
the previous sentence. However, difficulties arise
when one has to select, in a dynamic environment,
the appropriate bubble radius, and the appropriate
pressure. The maximum bubble radius is appropriate
for one part of the spectrum, while the minimum
bubble radius is more appropriate for the collapse
phase.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Yin Lu Young
Princeton University, USA
Excellent work and presentation!
questions:

I have two

1) Have you investigated which effect is more


important in the modification of the R-P
equation? The averages of the pressure along the
surface of the bubble or the slip velocity?
2) I assume the bubbles are not interacting with
each other. Do you prescribe some criteria to
ensure the bubbles do not collide with each
other?
AUTHORS REPLY
Thank you for your compliments and
questions. Yes, we have investigated the importance
of the two corrections. In the flow field of a vortex
and for bubbles captured by the vortex averaging is
the predominant factor and affects both quantitative
and qualitative results. Inclusion of the pressure
correction due to the slip velocity brings in a
quantitative correction to the results.
Since we were originally concerned with
cavitation inception, bubble interactions were not
included and are not expected to play any role. The
void fractions due to initial nuclei distribution are in
the 10-7 to 10-6 range. Now that we feel the method
applicable to a larger range of applications, we will
implement a model for interaction, where necessary,
following our previous approach in [a, b]. Probably,
as you suggested we will use some distance criterion
to determine coalescence.
[a] A Singular Perturbation Theory of the
Growth of a Bubble Cluster in a Superheated
Liquid, G.L. Chahine and H.L. Liu, Journal of Fluid
Mechanics, Vol. 156, pp. 257-279, July 1985.
[b] Cloud Cavitation: Theory, G.L.
Chahine, Proceedings, 14th Symposium on Naval
Hydrodynamics, Ann Arbor, Michigan, National
Academy Press, pp. 165-195, Washington, D.C.,
1983.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

St. Johns, Newfoundland and Labrador, CANADA, 8-13 August 2004

Ship Maneuverability in Shallow Water


Katsuro Kijima (Kyushu University, Japan)
ABSTRACT
 


On a ship maneuvering performance, there are some


published papers dealing with the hydrodynamic forces
acting on a ship in shallow water. Ship maneuverability
will basically depend on the hydrodynamic forces acting
on a ship. Especially, its force in shallow water will be
several times of the forces in deep water in certain case.
The ship maneuvering characteristics in shallow water
will remarkably be different with it in deep water.
However, as there are few papers dealing with the
maneuvering characteristics in shallow water, this paper
investigates the maneuvering characteristics in its water.
This paper shows some examples on a ship maneuvering motion in shallow water by numerical simulation
and by measured results due to the model test. As first
stage on this investigation, the relationship between ship
maneuvering characteristics and its forces has shown as
one example.


 !" #

XCYZ[\ZL]6^7_4`

FCGHIJHLKMON P

_a

NEQ

R4FTSVU
N PWS1F

bcXCd a
_deZ`fgX
uCv:u1wx
yzwx{wLu6|7v }
$
%'&'()+* %-,
3

8:9;<=

$
%/.102

>?<@ACBED

*4%5(".6(7*4%
hCijkmljLn6o7p q

pEr

osq-rcklh
q tTh

Figure 1 Total evaluation of ship maneuverability

INTRODUCTION

papers have dealt with its hydrodynamic force or estimation method for the force by using strip theory, slender
body theory or CFD etc.
But there is little published paper on relating with
the detail investigation for ship maneuvering characteristics in shallow water. This paper, therefore, deals with
the detail maneuvering characteristics in shallow water.
Consequently, the results based on this investigation will
be useful for ship operation and ship handling in shallow water for preventing marine disaster and for safety
of navigation.

There are still so many marine disasters in restricted


water such as harbor, bay or canal. In these restricted
water, the maneuvering motion of a ship are affected by
the water depth or bank wall.
Generally speaking on the analyzing for marine disaster, we can pick up the following items shown in Figure 1 as most important factor.
On the prevention of marine disaster, we have to investigate the inherent ship maneuverability, human factor and environment condition such as wind, wave and
water depth for analyzing the disaster. The relationship
between the inherent ship maneuverability and human
factor correspond to evaluate of response, environmental
condition and human factor correspond to evaluate its influence, and inherent ship maneuverability and environmental condition correspond to evaluate of performance.
Finally, we have to consider the total evaluation included
the above each evaluations. Therefore, of course it is indispensable to take into consideration both environmental condition and human factor in the inherent ship maneuverability.
When a ship is running in shallow water, the water depth will give some influence to the maneuvering
characteristics. The main factor on this will be hydrodynamic forces acting on a ship. Then so many published

HYDRODYNAMIC FORCES IN SHALLOW WATER


Hydrodynamic forces acting on a ship will change
remarkably depending on the water depth. For instance,
lateral force, YH0 , and yawing moment, NH0 , which acting on a ship hull as function of drift angle, , and nondimensional yaw rate, r0 , when water depth / draft ratio
H/d (H : water depth, d : draft) is equal to 1.2 and 1.5 are
shown in Figure 2. Each symbol in the figure indicates
measured hydrodynamic force. The gradient of lateral
force and yawing moment over and r0 is different for
each water depth condition. It means that maneuvering
characteristic of a ship will be changed because ship mo-

1
Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

tion is dominated by these hydrodynamic forces. Then it


is very important to understand the characteristics of the
hydrodynamic forces precisely.

There are some studies about prediction of hydrodynamic forces acting on a ship hull in shallow water theoretically or experimentally. The prediction method of
hydrodynamic forces in shallow water based on slender
body theory was proposed by Nonaka et al. (Nonaka et
al., 1997). As examples of experimental study, there are
papers by Hirano et al. (Hirano et al., 1985) or Kijima
et al. (Kijima et al., 1985). Recently application of CFD
for prediction of hydrodynamic forces in shallow water
was done by Ohmori (Ohmori, 1998).
From the practical point of view, simple prediction
method is required at the initial design stage of a ship.
Therefore the author had proposed approximate formulae to estimate hydrodynamic derivatives in deep water
(Kijima and Nakiri, 1999). The approximate formulae
consist of the principal particulars and some parameters
which represents hull shape as follows,



1

Y0 = k + 1.9257 Cb B/L a ,

0
0
0
0

Yr = k + 0.052ea 0.457 + m + m x ,

Y = 1.199Cb a + 1.05,




0
0
(1)

Yrr = 0.225 dCb /B ea 0.12,

n 
 o

Yrr
= 7.1256 d 1 Cb /B ,

hn 
 o i2

0
0

Yr = 10.443 d 1 Cb /B ea

n 
 o

9.374 d 1 C /B e + 1.227,

; r = 0.0
; r = 0.3125
; r = 0.50

YH
1.0
0.5
0.0
0.5
1.0
5

10
(deg)

10
(deg)

NH
0.2

0.0

0.2
5

(a) H/d = 1.2

; r = 0.0
; r = 0.3125


n 
o2

N0 = k 150.668 d 1 Cb /B e0a K

n 

o
23.819 d 1 Cb /B e0a K + 1.802 ,

; r = 0.50
; r = 0.70

YH
0.5

Nr0 = 0.54k + k2 0.0477e0a K + 0.0368,


n 
o2

0
N
= 43.857 d 1 Cb /B e0a K
n 
o

3.671 d 1 Cb /B e0a K + 0.086,

0.0

0.5
5

10
(deg)

0
Nrr
= 0.15K 0.068,
0
Nrr
= 0.4086Cb + 0.27,
n 
 o
0
Nr = 0.826 d 1 Cb /B e0a 0.026,

NH
0.2

(2)

0.0

where,
0.2
5

L
B
d
Cb
k
m
mx

10
(deg)

(b) H/d = 1.5


Figure 2 Comparison between measured and estimated hydrodynamic forces

:
:
:
:
:
:
:

ship length,
ship breadth,
draft,
block coefficient,
2d/L,
ship mass,
longitudinal component of added
mass.

2
Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

PREDICTION OF SHIP MANEUVERING MOTION

ea , e0a , a and K are coefficients defined as follows. They


express characteristics of aft hull shape.

ea = (1 C pa ),

B e

a
0

,
ea = q

1
1

4
(B/d)2
(3)

1 Cwa

a , =

1 C pa

1.5
1

0.33 (0.95a + 0.40).


K= 0 +

ea L/B

Generally maneuvering characteristics changes considerably depending on the depth of water. Figure 3
shows turning circles for different water depth. These
lines in this figure were obtained by the numerical simulation by using the proposed approximate formulae, and
also they agreed completely well with measured results
in model test as shown in reference (Kijima and Nakiri,
2003). It is observed that turning circle becomes larger if
the depth of water shallows. Therefore the effect of water
depth upon precise hydrodynamic coefficients should be
taken into account when we estimate maneuvering performance by numerical simulations.

As described above, hydrodynamics forces and moment acting on a ship hull will change remarkably depending on water depth. The author had proposed the
extended formulae to estimate hydrodynamic derivatives
in shallow water based on that in deep water (Kijima and
Nakiri, 2003). When hydrodynamic derivatives in deep
water and shallow water are noted with D0 () and D0 (h)
(h = d/H : draft / water depth ratio, d : draft, H : water
depth) respectively, there are following correlation between them.
0
1) for Y0 , N0 and Yrr
)
(
1
h f (h) D()
D(h) =
(1 h)a

(4)

2) for Yr0
D(h) = (1 + a1 h + a2 h2 + a3 h3 )


 

D() m0 + m0x () + m0 + m0x (h)

Figure 3 Turning trajectories depending on water


depth in Esso Osaka

(5)
0
0
0
0
3) for Nr0 , Y
, Yrr0 , Yr
, Nrr0 , Nr
and Nrr

D(h) = (1 + a1 h + a2 h2 + a3 h3 ) D()
4) for

(6)

x0

0
N

x, X

D(h) = (1 + a1 h + a2 h2 + a3 h3 ) f (h) D()


(7)
G

a1 , a2 and a3 are constants and f (h) is a function of h.


They take different value and form which change with
the kind of derivatives. Detailed expressions for these
constants and functions are shown in reference (Kijima
and Nakiri, 2003).
Comparison between measured and estimated lateral
force and yawing moment for H/d = 1.2 and 1.5 is
shown in Figure 2. Each symbol in the figures indicates
measured value and each line represents estimated value
using equations (1) to (7). There are a few discrepancies
in lateral force when yaw rate is big, estimated results
almost agree with measured results.

r, N
y, Y

y0

Figure 4 Coordinate systems

3
Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

XH0 , YH0 and NH0 , are expressed as follows,

The equations for surging, swaying and yawing motion of ship can be written in the following form using
coordinate systems in Figure 4 (Kijima et al, 1990),
!
  L  U
cos sin
U U


+ m0 + m0y r0 sin = X 0 ,
!

  L  U
0
0

sin + cos
m + my
U U

+ m0 + m0x r0 cos = Y 0 ,
!

  L 2 U
U 0
0
0
0
r + r
= N0.
Izz + izz
U
L
L
m0 + m0x

(8)

0
0
Xuu
is resistance coefficient and Xr
can be estimated
by Hasegawas chart (Hasegawa, 1980). Hydrodynamic
derivatives for YH0 and NH0 can be estimated using equations (1) to (6).
Mathematical model for thrust produced by propeller
is expressed using thrust coefficient KT defined as function advance coefficient JP :

The superscript 0 in the equations refers to the nondimensional quantities defined by:
m

, m0x , m0y

m, m x , my
1
2
2 L d

X, Y
X ,Y = 1
,
2
2 LdU
rL
,
r0 =
U
0

Izz0 , i0zz

Izz , izz
1
2
2
2 L dU

N
N = 1
,
2
2
2 L dU
0

 1
XP0 = (1 tP0 )n2 D4P KT JP / LdU 2 ,
2

KT JP = C1 + C2 JP + C3 JP2 ,


JP = U cos 1 wP / nDP ,
wP = wP0 exp(4.00P 2 ),
0P = x0P r0 ,
x0P ' 0.5,

(9)

where,
L, d
m
m x , my

:
:
:

Izz , izz

U,
r
X, Y

:
:
:

:
:

ship length and draft,


ship mass,
x and y-axis components of added
mass of a ship,
moment and added moment of inertia
of a ship,
ship speed and drift angle,
angular velocity,
x and y-axis components of external
force acting on a ship,
yaw moment acting on a ship,
the density of fluid.

(12)

(13)

where,

The non-dimensional external forces X 0 & Y 0 and


moments N 0 can be expressed assuming that they consist of hull, propeller and rudder components noted with
subscripts H , P and R ,
X 0 = XH0 + XP0 + XR0 ,
Y 0 = YH0 + YR0 ,
N 0 = NH0 + NR0 .

(11)

0 0
0
XH0 = Xr
r sin + Xuu
cos2 ,

0
YH0 = Y0 + Yr0 r0 + Y
|| + Yrr0 r0 r0


0
0
+ Yr
+ Yrr
r0 r0 ,

0
NH0 = N0 + Nr0 r0 + N
|| + Nrr0 r0 r0


0
0
+ Nr
+ Nrr
r0 r0 ,

tP
tP0

:
:

n
DP
C1 , C2 , C3
wP

:
:
:
:

wP0

thrust reduction coefficient,


thrust reduction coefficient in straight
forward motion,
propeller revolution,
propeller diameter,
constants,
effective wake fraction coefficient at
propeller location,
effective wake fraction coefficient at
propeller location in straight forward
motion.

Mathematical models for terms on rudder forces are


assumed as:

XR0 = (1 tR )F N0 sin ,

0
0
(14)
YR = (1 + aH )F N cos ,

0
0
0
0
NR = (xR + aH xH )F N cos ,

(10)

where,

There are many kinds of mathematical models for


each component in equation (10). Hereafter the mathematical model which proposed by the author is shown
as an example.
Hydrodynamic forces and moment acting on a hull,

tR
aH
x0H

:
:
:

coefficient for additional drag,


ratio of additional lateral force,
non-dimensional distance between
the center of gravity of ship and
the center of additional lateral force
(x0H = xH /L),

4
Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

xR0

non-dimensional longitudinal distance between the center of gravity


of ship and the center of lateral force
(xR0 = xR /L),
rudder angle.

2 y/L
6

wR0

P
0

1 , 2

:
:
:
:

; r
;

; U/U0
;

(deg)
60

r, U/U0

=35

Time history

(deg)

1.0
40

400

20

200

0.5

0.0

0
100 time(sec)

50

Figure 5 Comparison of trajectories and time histories (Cargo ship, H/d = 1.2, = 35 )

where,

wR

Symbol; measured
Line; calculated

:
:
:
:
:
:

x/L

Following expressions are assumed for the normal


force acting on rudder F N0 with normal force coefficient
CN :

F N0 = AR /Ld C N UR02 sin R ,
(15)

C N = 6.13KR /(KR + 2.25),

2

02

UR = 1 wR {1 + Cg(s)} ,

g(s) = K{2 (2 K) s}/ (1 s) ,

= DP /hR ,




K = (1 0.6 |sin |) 1 wP / 1 wR ,

(16)


s = 1.0 1 wP U cos /nP,

n
o

wR = wR0 exp 4.0 + r0 /2 2 2

R = + 0 1 R ,

0
0
0

R = 2xR r ,

x ' 0.5,

AR
hR
KR
UR
R
C

Trajectory

=35

rudder area,
rudder height,
aspect ratio of rudder,
effective rudder inflow speed,
effective rudder inflow angle,
coefficient for starboard and port rudder,
effective wake fraction coefficient at
rudder location,
effective wake fraction coefficient at
rudder location in straight forward
motion,
propeller pitch,
toe angle of offset rudder,
flow straightening coefficient,
functions which express effect of
rudder angle on and wR0 .

x/L
Trajectory

=35

Symbol; measured
Line; calculated

2
0
2
0

; r
;

r, U/U0
Time history

6 y/L
; U/U0
;

(deg)
60

=35

(deg)

1.0
40

400

20

200

0.5

Solving equation (8) with consideration of the influence of water depth on the hydrodynamic coefficients,
ship maneuvering motion can be estimated.
Figure 5 and 6 show turning trajectory and time histories of non-dimensional yaw rate, r0 , speed reduction
ratio, U/U0 (U0 : initial speed), drift angle, , and heading angle, for a cargo ship and a coal carrier at H/d =
1.2 respectively. Each line indicate simulation result the
symbols indicate measured values. It can be said from
these comparison between numerical simulation and measured results that this approximate formulae have enough

0.0

50

0
100 time(sec)

Figure 6 Comparison of trajectories and time histories (Coal carrier, H/d = 1.2, = 35 )

5
Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

; H/d=6.0

x/L

; H/d=6.0

; H/d=1.5

x/L

; H/d=1.2

; H/d=1.5
; H/d=1.2

2
5

y/L

Figure 7 Change of turning trajectory depending on


water depth (Coal carrier)

y/L

Figure 8 Change of turning trajectory depending on


water depth (Chemical tanker)

1.0

1.0

0.0

0.0

; H/d=6.0
; H/d=1.5

; H/d=6.0

; H/d=1.2

; H/d=1.5
; H/d=1.2

1.0

40

1.0

20

20

40

40

(deg)

Figure 9 Change of r0 curve depending on water


depth (Coal carrier)

20

20

40

(deg)

Figure 10 Change of r0 curve depending on water


depth (Chemical tanker)

r0 for both ships.


There is a difference on the transition of advance and
tactical diameter for both ships. On Figure 7, both advance and tactical diameter becomes bigger evenly as the
depth of water shallow. On the other hand, advances for
three different water depth have almost same value and
only tactical diameter becomes bigger on the Figure 8.
Figure 11 indicates transition of advance, AD , transfer, T r , and tactical diameter, DT for three kinds of ship
as function of draft / water depth ratio, d/h. It is observed that each ship has inherent tendency on change of

accuracy to estimate maneuvering motion in shallow water.


Figure 7 and Figure 8 show the change of turning
trajectory for a coal carrier and a chemical tanker depending on the depth of water. As water depth becomes
shallower, namely H/d becomes smaller, the radius of
turning circle becomes bigger on both figures. It is well
known that course stability is stabilized generally as the
water depth becomes shallower. It can be understood
from Figure 9 and Figure 10 which represent correlation
between rudder angle, and non-dimensional yaw rate,
6

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

tively. Therefore it means that course stability is stable


if `r0 is bigger than ` .
Figure 12 shows the transition of `r0 and ` for the
change of water depth. Shallow water effect appears
conspicuously on a container ship shown in Figure 12
(a). The course stability of the ship is stable when d/H
is equal to 1/1.2. As for the rest ships, the sign of `r0 `
is always negative, namely course stability is unstable
regardless of water depth. However the shape of `r0 `
curve is each different for these four ships.
CONCLUDING REMARKS
Ship maneuvering characteristics in shallow water will
be remarkably different with it in deep water. The most
important factor on its deference will be hydrodynamic
forces acting on ship, especially it seems the deference
depend on the position of the acting point of yaw damping force and sway damping force. These forces mostly
depend on ships form such as body plan or flame line of
the hull body.
REFERENCE
Hasegawa, K., On a Performance Criterion of Autopilot Navigation, Journal of the Kansai Society of Naval
Architects, Japan, Vol.178, 1980, pp.93104.

Figure 11 Change of advance, transfer and tactical diameter depending on water depth

Hirano, M., Takashina, J., Moriya, S. and Nakamura,


Y., An Experimental Study on Maneuvering Hydrodynamic Forces in Shallow Water, Transaction of the
West-Japan Society of Naval Architects, No.69, 1985,
pp.101110.

these parameters.
As shown in the above, a turning advance and tactical
diameter will change as function of water depth and in
ship type in spite of same rudder execution.

Kijima, K., Katsuno, T., Nakiri, Y. and Furukawa, Y.,


On the Manoeuvring Performance of a Ship with
the Parameter of Loading Condition, Journal of the
Society of Naval Architects of Japan, Vol.168, 1990,
pp.141148.

CORRELATION BETWEEN HYDRODYNAMICS


FORCES AND MANOEUVRING CHARACTERISTICS

Kijima, K., Murakami, M., Katsuno, T. and Nakiri, Y.,


A Study on the Ship Manoeuvring Characteristics in
Shallow Water, Transaction of the West-Japan Society
of Naval Architects, No.69, 1985, pp.111122.

It is well known that the linear terms of hydrodynamic derivatives are important parameters to evaluate
the course stability of a ship. If the linear terms of hydrodynamic derivatives satisfy the following equation, the
course stability is stable.


Y0 Nr0 + N Yr0 m0 + m0x > 0
(17)

Kijima, K. and Nakiri, Y., Approximate Expression


for Hydrodynamic Derivatives of Ship Manoeuvring
Motion taking into account of the Effect of Stern
Shape, Transaction of the West-Japan Society of
Naval Architects, No.98, 1999, pp.6777.

Equation (17) can be rewritten as follows,


N0
Nr0
 0 = `r0 `0 > 0,
Yr0 m0 + m0x
Y

Kijima, K. and Nakiri, Y., On the Practical Prediction


Method for Ship Manoeuvrability in Restricted Water, Transaction of the West-Japan Society of Naval
Architects, No.107, 2003, pp.3754.

(18)

where `r0 and `0 indicate the position of the acting point


of yaw damping force and sway damping force respec7

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

lr

0.5

1
h(=d/H)

0.5

0.5

1
h(=d/H)

lr l

0.5

1
h(=d/H)

0.5
lr l

0.5

0.5

1
h(=d/H)

1
0

0.5
lr l

1
h(=d/H)

0.5

1
h(=d/H)

0.5

1
h(=d/H)

0.5

1
h(=d/H)

1
0
1

lr l

0.5

(d) Coal carrier

1
h(=d/H)

1
h(=d/H)

0.5

1
h(=d/H)

(c) Chemical tanker

1
0

0.5
lr l

lr

1
0
1

1
0

1
0

1
0

1
h(=d/H)

(b) Cargo ship

1
h(=d/H)

1
0

1
h(=d/H)

(a) Container ship


lr

0.5
l

5
0

1
0
1

1
0
1

1
h(=d/H)

5
0

lr

1
0
1

5
0
5

lr

1
0

(e) VLCC
Figure 12 Change of `r0 and `0 depending on water depth

Nonaka, N., Haraguchi, T., Nimura, T., Ueno, M., Fujiwara, T., Makino, M., Kodama, Y. and Yoshino, Y.,
Research on Flow Field around a Ship in Manoeuvring
Motion, Papers of Ship Research Institute, Vol.34,
No.5, 1997, pp.168.
Ohmori, T., A Study on Hydrodynamic Characteristics of a Maneuvering Ship in Shallow Water by a
Finite-Volume Method, Proceeding of International
Symposium and Workshop on Force Acting on a
Maneuvering Vessel (MAN98), 1998, pp.15-38.

8
Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Martin Renilson
QinetiQ, United Kingdom
I wonder if the author could comment on
scale effects in shallow water. Perhaps these would
be more severe than in deep water, and may be the
cause of some of the unusual results.
AUTHORS REPLY
Thank you for your question. On scale
effects of ship maneuvering motion in shallow water,
it seems it will be very difficult to investigate it,
especially in the full scales ship. Because, we have no
data for comparing the estimated results with
measured results in full scales ship in shallow water.
We know only a few published paper relating with
the data of full scales ship maneuvering
characteristics in shallow water.
We need to collect the data of its in shallow
water.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Jinzhu Xia
Australian Maritime College, Australia
It would be interesting if Prof. Kijima could
provide some discussion on wave effects and the
assessment of wave effects on maneuvering and
maneuvering in shallow water.
AUTHORS REPLY
Thank you very much for your discussion.
In this paper, the wave effects was not discussed, but
if you use our mathematical model, you can estimate
and consider the wave effect in ship maneuvering
motion by adding the wave force coefficients in the
mathematical model. Of course, you have to estimate
the wave drift force and wave exciting force acting
on ship hull.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics


St. Johns Newfoundland and Labrador, CANADA, 8-13 August 2004

Numerical Simulations of Breaking Wave


Around a Wedge
R. Broglia, A. Di Mascio and R. Muscari
(INSEAN, Italian Ship Model Basin, Italy)
breaking are responsible for both air entrainment
and downstream convection of air bubbles.
The numerical algorithm adopted is based on
the general pseudocompressible formulation of the
Reynolds Avereged Navier Stokes equations. In order to overcome the difficulties connected with gravity waves in breaking regimes, the solver has been
coupled with a non standard single phase levelset
approach (Di Mascio et al., 2004). In this approach
only the liquid phase is simulated, and the level set
function is used as a tracking device to locate the
intersection of the free surface with the grid lines.
A single fluid being computed, there is no transition zone across which the fluid properties vary, and
consequently no uncertainty is related to interface
smoothing. In addition, the algorithm is such that
the level set function satisfies the free surface kinematic boundary condition only on the free surface
itself, whereas it turns out to be a distance function
at any other point, which guarantees mass conservation properties.
The paper is organized as follows: in the next
section, the physical problem is defined, and the
conservation equations governing the flow of an incompressible viscous fluid are briefly recalled. Then,
the numerical algorithm used for the simulation of
the bulk flow and the free surface will be summarized, and the singlephase level set approach recalled. The second last section is dedicated to the
description of the numerical results, and to the comparisons with available experimental data in terms
of both wave profiles and cross sections of the free
surface; moreover, analysis of the vorticity field in
the water jet formation and in the impact regions
will be shown. Conclusions will be drawn in the last
section.

ABSTRACT
This paper describes a numerical study of the
flow around a sharp wedge in both breaking and
nonbreaking regimes. The wedge is fixed on a free
slip bottom. In the numerical simulations two bow
angles and three water depths are considered; the
simulations correspond to some experimental tests
carried out by Waniewski et al. (2002) with which
comparisons are made. Froude number, based on
water depth and free stream velocity, ranges from
2.57 to 3.29, whereas Reynolds number is around
1.86 106 . The numerical method used is a solver for
steady incompressible free surface flows based on the
pseudocompressible formulation of the Reynolds Averaged Navier Stokes equations (RANSE); free surface is handled by a singlephase level set approach
(Di Mascio et al., 2004).
INTRODUCTION
This paper deals with numerical simulations of
the flow around a wedge, fixed to a free slip bottom. This kind of flow is taken as a prototype of the
flow around the bow of a ship hull moving through
the water: the main characteristic of this flow is the
presence of a surface wave which, under some circumstances, can break. Bow wave breaking is one
of the main source of air entrainment and of the so
called whitewater wake phenomena; moreover, the
dispersion of bubbles in water can have strong negative effects in terms of, for example, signature of
the ship and propeller efficiency.
Objectives of this work are the study of the bow
wave in breaking condition and the analysis of the
vorticity production due to the formation of the
water jet and to the breaking phenomena. Attention must be devoted to this aspect, as coherent
structures like the streamwise vortices produced by

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

GOVERNING EQUATIONS

is a material surface, and allows to determine its unknown configuration:

The turbulent motion of an incompressible (constant density) viscous fluid can be described by the
Reynolds Averaged Navier Stokes equations:
uj
=0
xj
ui
uj ui
p
ij
+
+
=
t
xj
xi
xj

D F (x, y, z, t)
=0
Dt

Initial conditions have to be specified for the velocity


field and for the free surface configuration.

(1)
i = 1, 2, 3.

NUMERICAL METHOD
In this section the numerical method used is recalled; for a complete description of the algorithm
the reader is referred to Di Mascio et al. (2001,
2004). When only the average steady state has to
be computed, the system of equations (1) can be
conveniently replaced by the pseudocompressible
formulation (Chorin, 1967), that reads, in integral
form:
Z
I

p dV +
ui ni dS = 0
t V
S(V)
Z

(4)
ui dV +
t V
I
[ui uj nj + pni ij nj ] dS = 0

A reference length l and velocity U have been chosen to make the equations nondimensional. In the
previous equations, ui is the ith Cartesian component of the velocity vector (in the following, the
Cartesian components of the velocity will be also
denoted with u, v, and w); p is a variable related
to the pressure P and the acceleration of gravity g
(parallel to the vertical axis z,downward oriented)
by p = P + z/F n2 , F n = U / gl being the Froude
number. Finally, ij = t (ui,j + uj,i ) is the stress
tensor, t = 1/Rn+T is the global kinematic viscosity, with Rn = U l/ the Reynolds number, the
kinematic viscosity and T the turbulent viscosity.
In the present work, the turbulent viscosity was calculated by means of the Spalart and Allmaras one
equation model (Spalart and Allmaras, 1994). In
what follows, the water depth d and the free stream
velocity U are used as reference quantities.
The problem is closed by enforcing appropriate
conditions at physical and computational boundaries. On solid walls, velocity is set to zero (whereas
no condition on the pressure is required); at the (fictitious) inflow boundary, velocity is set to the undisturbed flow value, and the pressure is extrapolated
from inside; on the contrary, the pressure is set to
zero at the outflow, whereas velocity is extrapolated
from inner points.
At the free surface, whose location is one of the
unknowns of the problem, the dynamic boundary
condition requires continuity of stresses across the
surface; if the presence of the air is neglected, the
dynamic boundary condition reads:
p = ij ni nj +
ij ni t1j = 0

+
We F n2

(3)

S(V)

where is the pseudocompressibility factor, V a


control volume, S(V) its boundary, and ~n the outward unit normal. The previous system of equations
is approximated by a finite volume technique, with
pressure and velocity colocated at the cell center;
the residual on each control volume is computed as
flux balance at the cell interfaces:
Z
6 Z
X

~q dV +
(F~sc F~sd ) dS = 0
(5)
t Vijk
S
s
s=1
where ~q = (p, u, v, w)T is the state variables vector
for pseudocompressible flows, and F~sc and F~sd are
the convective (inviscid and pressure) and diffusive
normal fluxes at the sth interface Ss of the finite
volume Vijk . In order to obtain second order accuracy in space, convective and viscous fluxes are
computed by means of the trapezoidal rule.
The computation of the viscous fluxes requires
the value of the stress tensor at cell interfaces; for
instance, at the interface i + 12 , j, k :


um
ul
lm |i+ 1 ,j,k = i+ 21 ,j,k
+
(6)
2
xl
xm i+ 1 ,j,k

(2)

ij ni t2j = 0

2
where We = U
l/ is the Weber number ( being
the density of the fluid and the surface tension
coefficient) and the surface curvature; ~n, ~t1 and ~t2
are the surface normal and two tangential unit vectors, respectively. In this work surface tension effects have been neglected. The kinematic boundary
condition states that the free surface F (x, y, z, t) = 0

Velocity gradients are computed by means of a standard second order centered finite volume approximation.
2

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

set function remains unchanged at material points.


In classical twophase level set approaches, the density and the molecular viscosity of the fluid are then
assumed to depend on the sign of the level set function; to avoid numerical difficulties related to sharp
discontinuities, density and viscosity have a smooth
transition around the zero level (x, y, z, t) (Sussman et al., 1994).
In order to maintain the thickness of the interface constant in time and avoid mass loss (Sussman et al., 1994), level set function has to remain
a distance function for t > 0. In the socalled re
initialization step, the level set function (x, y, z, t)
is replaced, at each time step, by a new function
y, z, t) with the same zero level, but again rep(x,
resenting the distance from the interface; the func y, z, t) is computed at each physical time
tion (x,
step as the asymptotic (steady state) solution, with
respect to the pseudotime , of the equation:

For the inviscid part, a second order Essentially


Non Oscillatory (ENO) scheme (Harten et al., 1987)
has been adopted; convective fluxes are computed
as the solution of a Riemann problem at the cell
interface:
c
~ c qi+ 1 ,j,k ) = F~ c (~ql , ~qr )
F~i+
1
,j,k = F (~
2

(7)

whose left and right states are given by:


~ql = ~qi,j,k +

minmod( ~q|i1/2 , ~q|i+1/2 )

~qr = ~qi+1,j,k

2
minmod( ~q|i+1/2 , ~q|i+3/2 )
2

(8)

where:
~q|i+1/2 = ~qi+1,j,k ~qi,j,k

(9)

and minmod(x, y) is a function (to be applied to


each vector component) defined as:

0
xy 0
minmod(x, y) =
sign(x)(min |x|, |y|) xy > 0
(10)
Time integration of the discrete model is achieved
by means of an implicit Euler scheme; the resulting
discrete system of equations is solved in delta form
as in the Beam and Warmings scheme (1978)
Convergence to steady state is accelerated by local time stepping and a multigrid algorithm (for
more details and performances of this technique applied to time marching ENO schemes, see Favini
et al. (1996)).


1) = 0
+ sign()(||

(12)

where sign is the sign function (Rouy and Tourin,


1992).

SINGLEPHASE LEVEL SET


In level set approaches (see for example (Osher
and Sethian, 1988) and (Sussman et al., 1994)), a
smooth function (x, y, z, t), whose zero level coincides for t = 0 with the free surface, is defined in
the whole physical domain (i.e. in both liquid and
air phases); the kinematic boundary condition (3) is
extended to all the points in the domain, yielding a
transport equation for the level set function:
(x, y, z, t)
+ ~u (x, y, z, t) = 0
t
(x, y, z, 0) = d(x, y, z)

Figure 1: Computational domain. Squares: nodes in the


liquid phase; circles: nodes in the gas phase. Full symbols:
nodes where the kinematic condition (15) is enforced.

(11)
In the singlephase algorithm adopted only the
liquid phase of the fluid is computed; the computational domain is formally decomposed in (see figure (1)):

~u being the velocity of the underlying flow, and


d(x, y, z) the signed distance from the free surface
at t = 0. The zero level of the function (x, y, z, t)
represents the free surface location for t > 0; moreover, initializing (x, y, z, t) as the signed distance
from the surface of discontinuity, the sign of the level

grid points close to the surface of discontinuity


(full circles and full squares in figure (1)): the
level set function is computed by means of the
3

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

The interface values i+ 12 ,j,k is computed as:

evolution equation (11); velocity and pressure


in the liquid region (full squares) are computed
by means of the Navier Stokes and continuity
equations (5); pressure is evaluated by using
the dynamic boundary condition and the velocity is extrapolated at the points in the air
phase (full circles);

i+ 12 ,j,k = i,j,k +

i+ 12 ,j,k = i+1,j,k

minmod( |i+ 3 , |i+ 1 )


2

1
if Ui,j,k
0, and similarly for i 12 ,j,k . It has been
proved by Harten et al. (1987) that this procedure
yields a second order approximation to (11). Time
advancement of equation (15) is achieved by means
of a standard twolevel multistage second order
RungeKutta scheme (Jameson et al., 1981).
For the nodes in the water region which are not
close to the surface of discontinuity (empty squares
in figure (1)), the level set function is enforced to
represent the distance from the interface when the
steady solution is attained. To this aim, the constrain || = 1 is enforced by means of an iterative
marching ENO scheme with second order accuracy;
first, the above condition is rewritten as an evolution
equation for the level set function (x, y, z, t):



+ sign()
1 = 0
(17)
t
||

grid points in the air region (empty circles in


figure (1)): the level set function is computed
from equation (12) in order to get a distance
function, and an extension velocity is computed as done by Adalsteinsson and Sethian
(Adalsteinsson and Sethian, 1999).
It has to be noted that, in the singlephase formulation, the solution outside the water region is not
required; however, the extension of the velocity field
outside the water region ensures second order accuracy also close to the interface.
It was found convenient, when computing free
surface flows around complex geometries with curvilinear grids, to split the level set function as:

then, by using the definition (13), the previous equation is rewritten in term of (x, y, z, t) as:

(13)

where the function (x, y, z, t) is the solution (from


equation (11)) of:

+ ~u + w = 0
(14)
t
By doing so, it is easier to assign the boundary condition for the level set function at inflow, that reduces to (x, y, z, t) = 0.
An ENO technique (similar to the one used for
the bulk flow) is used to discretize equation (14); to
this end, the equation is first rewritten in terms of
curvilinear coordinates:

+ Um
+w =0
(15)
t
m

where:

+w
~ + b = 0
t

(18)

||


z
b = sign()
1
||

(19)

w
~ = sign()

This equation is solved with respect to (x, y, z, t)


with a scheme analogous to the one used to solve
the kinematic equation (14) at the points adjacent
to the free surface, with characteristic speed w.
~ The
same equation is used to update the level set function values in the air phase. It can be seen, that,
if the solution converges toward a steady state, it
provides a function (x, y, z, t) that satisfies equation (11) on the surface (x, y, z, t) = 0 and it is
a distance function at any other points. Note that
(x, y, z, t) is a distance function also at points close
to the free surface (full symbols in figure (1)) because of equation (25) (see later).
Once the function (x, y, z, t) is known throughout the whole domain, the free surface is located by

m
being the contravariant components
xi
of the velocity vector. The derivatives of the function (x, y, z, t) at cell center are approximated by
a second order finite difference formula; considering,
for instance, the (i, j, k)cell center, for the coordinate line 1 it reads:
U m = ui

= i+ 12 ,j,k i 12 ,j,k
1

1
if Ui,j,k
0, or:

grid points in the liquid phase region (empty


squares in figure (1)): the solution is computed by the numerical solution of the governing equations (5), and the level set function is
enforced to be a distance function by means of
equation (12);

(x, y, z, t) = (x, y, z, t) + z

minmod( |i+ 1 , |i 1 )

(16)
4

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

with

pi,j,k + pi1,j,k
(23)
2
where pF S is computed from the dynamic boundary
condition (2).
Once the pressure is known, the normal velocity
at (i + 12 , j, k) is computed by solving the Riemann
problem:




pi+ 12 ,j,k pw + un |i+ 12 ,j,k un |w = 0 (24)
pi 12 ,j,k =

i+1,j,k

FS

i,j,k

i+1/2

j+1/2

j1/2

p
being = u
n + u
2n + and where pw and un |w represent the known state on the water side, computed
as in (8). The tangential velocity is simply extrapolated along the normal to the free surface, given by
/||, as in the following equation (25). The remaining dynamic boundary conditions for the tangential stresses in (2) are explicitly enforced when
computing the viscous fluxes at the cell interface
i + 12 , j, k .
Outside the water region, extension velocities are
computed as:

i1/2

Figure 2: Free surface detection and extrapolation of pressure.

the surface (x, y, z, t) = (x, y, z, t) + z = 0. The


intersection of this surface with the underlying grid
line is computed as follows (see figure 2). Consider,
for instance, the coordinate line 1 ; when in two adjacent points (i, j, k) and (i + 1, j, k) the condition

ui = 0

i = 1, 2, 3

(25)

holds, it means that the free surface cuts the segment Pi,j,k Pi+1,j,k at some point PF S (Pi,j,k is the
position vector that locates the i, j, k point). Then,
the portion of the segment below the free surface is:

which guarantees that (x, y, z, t) evolves as a distance function also at the points adjacent to the
free surface (full circles in figure (1); see, for the
proof, Adalsteinsson and Sethian (1999)). Since the
steady state solution is the goal of the computation,
the previous relation is substituted by an evolution
equation for the velocity components ui :

|PF S Pi,j,k |
|i,j,k |
=
|Pi+1,j,k Pi,j,k |
|i+1,j,k i,j,k |

ui
+ ui = 0
t

i,j,k i+1,j,k 0

(20)

(21)

which is solved by a second order ENO scheme analogous to the one used to solve the kinematic equation (14) and equation (18), characteristic speed being . Note that the values of the velocity, pressure and turbulent viscosity in the air phase are useless when steady state is attained, and therefore they
do not affect the formal accuracy of the numerical
scheme. Nevertheless, their estimation is of great
importance during the iterative procedure at those
points that change their physical state from air to
water, for which an initial estimate is needed.

if i,j,k < 0, the level set function having been defined as the distance from the interface. A similar
relation holds for i+1,j,k < 0, with i and i + 1 interchanged.
The computation of the residuals for the RANS
equations at those points whose neighboring cells
are not all into the water region need some attention. In fact, in these cases the numerical convective
and viscous fluxes at interfaces that separate two
cells, of which one is in the air region (as the interface i + 12 , j, k in figure (2)), must be evaluated; in
these points the proper information to compute the
correct flux are needed to retain second order accuracy. To circumvent this difficulty, the following

procedure is applied. The pressure at i + 12 , j, k is
extrapolated as:
pi+ 12 ,j,k = pi 12 ,j,k +

1
2

1
(pF S pi 12 ,j,k )
+

(26)

RESULTS
Some numerical simulations of the flow around
an infinitely long wedge have been carried out; the
problem consists in a vertical piercing wedge fixed
on the bottom (in the simulations, the bottom plate
is treated as a free slip wall). Test conditions are
reported in table (1), where d represents the draft

(22)

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Run
L1
L2
L3
H1
H2
H3

13.4o
13.4o
13.4o
26.6o
26.6o
26.6o

d (cm)
6.66
7.55
9.21
6.45
7.62
9.32

U (m/s)
2.44
2.43
2.46
2.61
2.40
2.46

Fn
3.01
2.81
2.59
3.29
2.77
2.57

Table 1: Test conditions: is the wedge half angle, d is the


water depth, U is the incoming velocity and F n is the Froude
number

(which is equal to the water depth of the unperturbed flow), U is the free stream velocity, and is
the bow half angle; two different wedge angles and
three different water depths were analyzed. In the
same table, the values of the Froude number, based
on the free stream velocity and the water depth,
are reported for each conditions. The geometry and
flow field conditions under investigation correspond
to the large flume experiments made by Waniewski
et al. (2002) (tests 1l to 6l).
The physical domain is discretized by means of a
single block grid with CH topology (see figure (3)),
with a total of 144 128 128 volumes, for the
simulations with = 26.6o and 72 64 64 volumes, for = 13.4o , along streamwise, normal to
the wedge side wall, and vertical directions, respectively; points are clustered toward both the leading edge and the side wall of the wedge, whereas
an uniform distribution has been adopted along the
vertical direction. At least four grid levels were used
for multigrid acceleration. A uniform velocity field
(equal to the upstream value), zero pressure field
and a flat free surface have been considered as initial conditions. As boundary conditions, free stream

Figure 4: Wedge flow, = 26.6o d = 9.32 cm (run H3):


perspective view of the threedimensional flow field.

velocity is imposed at the inflow, free slip condition


are enforced at the bottom, and zero pressure and
velocity gradients are prescribed at the outflow.
Figure (4) shows a threedimensional view of the
flow field (test case H3) around the wedge (its side
wall is colored in red and the symmetry plane is
colored in blue): the presence of a strong three
dimensional breaking wave is evident. In figures (5)
and (6) perspective view of free surface sections (for
all test cases considered) perpendicular to the side
wall of the wedge are presented; sections are taken
from r = 0 cm to r = 150 cm every 10 cm, r being the distance from the leading edge.
From
these figures some features of the flow, already observed both theoretically and experimentally (see for
example (Noblesse et al., 1991; Dong et al., 1997;
Waniewski et al., 2002)) are highlighted. As the
flow impinges on the wedge a vertical component of
the velocity appears (the discontinuity in the free
surface elevation at the leading edge is clearly observable from the figures); this induces a flow which
rides up on the plate creating a liquid sheet along
the wall. Further downstream, the formation of a
water jet can be observed; due to gravity, this jet
plunges toward the free surface. After the impact
on the free surface, a splashup and a second jet are
clearly seen. In figures (5) and (6), it has also to be
noticed that in the splash region the resolution of
the grid is too coarse to resolve the flow adeguately
(see figure (3)).
Top views of the wave created by the wedge are
shown in figures (7) and (8) for all test cases considered; contour lines of free surface elevation are plotted. As it can be seen, for wedge angle = 13.4o
the wave does not break, whereas for wedge angle
= 26.6o wave jet profiles, impact line and splash
up are present.

250

y (cm)

200

150

100

50

0
-100

-50

50

100

x (cm)

150

200

250

Figure 3: Wedge flow, medium mesh (72 64 64) used for


largest wedge angle simulations.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Z
X

Z
X

Z
X

Figure 5: Wedge flow, = 13.4o : sections perpendicular to


the side wall of the wedge for L1 (top), L2 (middle) and L3
(bottom) runs.

Figure 6: Wedge flow, = 26.6o : sections perpendicular to


the side wall of the wedge for H1 (top), H2 (middle) and H3
(bottom) runs.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

175

175

10.0
9.0
8.0
7.0
6.0
5.0
4.0
3.0
2.0
1.0
0.0

125

75
50
25
0

50
25
0

175

175

10.0
9.0
8.0
7.0
6.0
5.0
4.0
3.0
2.0
1.0
0.0

75
50
25
0

150

100
75
50
25
0

-25

-25

175

175

150

100
75
50
25
0
-25

150

10.0
9.0
8.0
7.0
6.0
5.0
4.0
3.0
2.0
1.0
0.0

125

25

50

21.0
18.0
15.0
12.0
9.0
6.0
3.0
0.0
-3.0

125

x (cm)

100
x (cm)

75

-25

125

x (cm)

100

-25

150

21.0
18.0
15.0
12.0
9.0
6.0
3.0
0.0
-3.0

125

21.0
18.0
15.0
12.0
9.0
6.0
3.0
0.0
-3.0

125
100
x (cm)

x (cm)

100

150

x (cm)

150

75
50
25
0
-25

75 100 125 150


y (cm)

Figure 7: Wedge flow: top view of the waves for L1 (top), L2


(middle) and L3 (bottom) runs.

25

50

75 100 125 150


y (cm)

Figure 8: Wedge flow: top view of the waves for H1 (top),


H2 (middle) and H3 (bottom) runs.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25

20
144x128x128
72x64x64
36x32x32
18x16x16

15

15

z (cm)

z (cm)

20

10

144x128x128
72x64x64
36x32x32
18x16x16

10

0
20

0
25
z (cm)

15

z (cm)

20
15

10

10

0
20

5
15
z (cm)

z (cm)

25

10

20

15

10

15

20
y (cm)

25

30

35

40

10

Figure 10: Wedge flow: contact lines computed on four different grid levels, and for H1 (top), H2 (middle) and H3 (bottom) test cases.

5
0
-10

10

20

30
r (cm)

40

50

60

70

maximum wave elevation on the plate increases with


the wedge entrance angle; for = 13.4o the maximum increases with the draft (i.e. decreases with
the Froude number based on the water depth); for
the higher wave angle = 26.6o this dependence is
not clear. By comparison with experimental data
made by Waniewski et al. (2002), for the higher
wedge entrance angle the overall agreement is quite
good, even if maximum values are predicted a bit
downstream with respect to the measured data; for
the lower angle the agreement is only qualitative.
Such discrepancies could be due to viscous effects
on the bottom plate that were not taken into account; such viscous effects probably have a stronger
influence for the lower entrance angle, the wave elevation being much lower.
In figure (12) the comparison for the maximum
elevation of the wave profiles between present numerical results and experimental data is reported;
non dimensional free surface elevation is defined as

Figure 9: Wedge flow = 26.6o : plunging jet profiles at


impact on four different grid levels, and for H1 (top), H2
(middle) and H3 (bottom) test cases.

A grid convergence study is carried out for the


larger entrance angle and the three water depths;
in figures (9) and (10) contact lines (defined as the
wave profile on the wedge side wall) and plunging
jet profiles at impact, computed on four grid levels,
are compared. In these figures, the leading edge is
placed at r = 0, y is the distance from the side wall
and z = 0 is the location of the undisturbed free
surface; plunging jet sections are taken perpendicularly to the side wall at about r = 78 cm, r = 66 cm
and r = 67 cm, for the H1, H2 and H3 test case respectively. Convergence when refining the grid can
be clearly inferred.
Contact lines for all the simulations performed
are presented in figure (11); as it can be seen, the

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

30

L1
L2
L3

25

20

H1
H2
H3

10
x

+
+ ++ +

+
++

xxxxxx

z (cm)

z (cm)

20
15

++
x xx

xx

++++
++ +
x x
xx

0
20

2.2
2.0
1.8
1.6
1.4
1.2
1.0
0.8
0.6
0.4
0.2
0.0

10

20

30
r (cm)

40

50

60

70
z (cm)

15

Figure 11: Wedge flow: contact lines, vertical tiny line indicates the leading edge. Experimental data (Waniewski et al.,
2002): ( ) L1 run, ( ) L2 run, ( ) L3 run, ( + ) H1 run,
( ) H2 run and ( 4 ) H3 run.

10

0
20

15

z (cm)

2.5
2.0
Z*max

10

0
-10

1.5

+ ++

10

1.0
0.5
0.0

2.2
2.0
1.8
1.6
1.4
1.2
1.0
0.8
0.6
0.4
0.2
0.0

15

10

15

20
y (cm)

25

30

35

40

2.2
2.0
1.8
1.6
1.4
1.2
1.0
0.8
0.6
0.4
0.2
0.0

Figure 13: Wedge flow = 26.6o : plunging jet profiles at


impact for H1 (top), H2 (middle) and H3 (bottom) test case.

Fr

case respectively. As it can be seen, the impact of


the plunging jet on the unperturbed free surface is
well captured and defined. The shape of the plunging jet for the = 26.6o looks very similar to the
experimental results obtained by Waniewski et al.
(2002) (see, for example, figure (11) in the cited
paper); moreover, the plunging jet thicknesses are
about 2 cm for H1 and H2 simulations and around
3 cm for the H3 test case, values in agreement with
the towed model experiment by Waniewski et al.
(2002). Impact velocities (which are around 2 m/s
for all cases, with a small increase with the water
depth) and angle of impingements (around 58o , 60o
and 52o for H1, H2 and H3 test cases) are in good
agreement with the value of 2.3 m/s for the impact
velocity and 58o for the angle of impingement provide by Waniewski et al. (2002).
In figures (14) and (15) cross sections of streamwise (xdirection) vorticity fields for all test cases
considered are presented. Sections are taken perpendicularly to the wedge side wall at different distances from the leading edge. The first series of
sections (figure (14)) are taken when the water jet
starts to separate from the side wall of the wedge,
i.e. where the wave profiles reach the maximum el-

Figure 12: Wedge flow: nondimensional maximum wave


height as a function of Froude number. Waniewski et al.
(2002): ( N ) small flume experiments = 13.4o , (  ) large
flume experiments = 26.8o and ( ) towing tank experiments = 26.0o ; Olgivie (1972) towing tank experiments
with = 15.0o : ( ) d = 10.2 cm, ( 4 ) d = 20.4 cm, (
)
d = 30.5 cm and ( + ) d = 40.6 cm. Present: (  ) = 13.4o
and (  ) = 26.6o and.

(see Waniewski (1999) and Olgivie (1972)):


90 Zmax
?
Zmax
=
F n1.5 d

(27)

The figure confirms that the large entrance angle


results are in a good agreement with experimental
data, at least for the two smaller Froude number,
whereas the maximum wave elevations for the small
entrance angle are somewhat underpredicted.
In figures (13) sections at the impact of the free
surface for the largest wedge angle and the three
water depth test cases considered are shown. Sections are taken around the first impact, perpendicular to the side wall of the wedge, which correspond
to a distance from the wedge leading edge of about
78 cm, 66 cm and 67 cm, for the H1, H2 and H3 test
10

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

15

Run L1
r=35 cm

5
0

Run L2
r=35 cm

50
40
30
20
10
0
-10
-20
-30
-40
-50

Run L3
r=100 cm

50
40
30
20
10
0
-10
-20
-30
-40
-50

5
0
-5
15

Run L3
r=35 cm

15
12
9
6
3
0
-3
-6
-9
-12
-15

5
0

10

15

20

25

y (cm)

30

10

z (cm)

10

5
0
-5

35

10

15

20

y (cm)

25

25

20

Run H1
r=35 cm

15
10
5
0
-5
20
15

5
0
-5
20

10

0
-5

Run H2
r=75 cm

20

15

10
5
0
-5

5
0

10

15

20

y (cm)

25

30

35

Run H3
r=80 cm

20

50
40
30
20
10
0
-10
-20
-30
-40
-50

10

50
40
30
20
10
0
-10
-20
-30
-40
-50

15

25

Run H3
r=35 cm

50
40
30
20
10
0
-10
-20
-30
-40
-50

50
40
30
20
10
0
-10
-20
-30
-40
-50

10

35

15

25

Run H2
r=35 cm

30

Run H1
r=90 cm

20

50
40
30
20
10
0
-10
-20
-30
-40
-50

z (cm)

z (cm)

10

z (cm)

15

z (cm)

Run L2
r=100 cm

-5

15
12
9
6
3
0
-3
-6
-9
-12
-15

z (cm)

50
40
30
20
10
0
-10
-20
-30
-40
-50

z (cm)

z (cm)

10

z (cm)

Run L1
r=100 cm

15

15

-5

10

50
40
30
20
10
0
-10
-20
-30
-40
-50

15

z (cm)

z (cm)

10

15
12
9
6
3
0
-3
-6
-9
-12
-15

z (cm)

15

10
5
0
-5

40

10

15

20

25

y (cm)

30

35

40

45

50

Figure 15: Wedge flow: vorticity fields at 100 cm from the


leading edge for the = 13.4o entrance angle case, and just
after water jet impact for the = 26.6o entrance angle case.

Figure 14: Wedge flow: vorticity fields at sections around


water jet formation.

11

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

evation (r = 35 cm for all cases). The second series


is taken at r = 100 cm for = 13.4o ; for = 26.6o
the sections are extracted just downstream the impact of the water jet. As shown in these figures, the
formation of the bow wave and the impact of the
plunging jet involve production of streamwise vorticity; the main features of these vorticity fields share
some similarities with the experimental observations
by Dong et al. (1997), who have studied the structure of mild and steep (depending on the Froude
number) bow waves around a ship model. In agreement with experimental observation by Dong et al.
(1997), at the formation of the water jet (for the
larger wedge angle, and further downstream for the
smaller entrance angle) negative vorticity is present
in the forward face of the wave, whereas the entire
wave crest region of the wave has positive vorticity;
this phenomena is more pronounced when increasing the Froude number. For the smaller angle of entrance, production of vorticity at the section where
the water jet starts (figure (14)) seems caused by
free surface curvature. At the impact section, for
the larger entrance angle, strong negative vorticity
due to the wave overturning can be clearly observed.

Dong et al. (1997), such as the presence of negative


vorticity on the forward face and a positive peak on
the crest have been clearly shown.
ACKNOWLEDGMENTS
The authors wish to thank Dr. Tricia Waniewski
Sur for the experimental data kindly provided.
References
Adalsteinsson, D. and Sethian, J. A. (1999). The
Fast Construction of Extension Velocities in Level
Set Methods. J. Comput. Phys., 148:222.
Beam, R. and Warming, R. (1978). An Implicit Factored Scheme for the Compressible Navier-Stokes
Equations. AIAA Journal, 16:393402.
Chorin, A. (1967).
A Numerical Method for
Solving Incompressible Viscous Flow Problems.
J. Comput. Phys., 2:1226.
Di Mascio, A., Broglia, R., and Favini, B. (2001).
A Second Order GodunovType Scheme for Naval Hydrodynamics, pages 253261. Kluwer Academic/Plenum Publishers.

CONCLUSIONS

Di Mascio, A., Broglia, R., and Muscari, R. (2004).


A SinglePhase Level Set Method for Solving Incompressible Viscous Free Surface Flows. Submitted for publication.

A single phase level set approach for free surface viscous flows at high Reynolds number have
been used to study the flow around a sharp wedge.
Numerical simulations at two different entrance angles and three different water depths are carried out;
in agreement with available experimental data (see
(Waniewski et al., 2002) and (Waniewski, 1999)) the
waves generated by the presence of the wedge is in
breaking conditions at the higher half bow wedge angle. Threedimensional views of the breaking flow
clearly show the formation of a water jet, plunging
toward the free surface where it impacts; further
downstream a splashup region develops. Contact
lines (the profile of the wave on the wedge side wall)
compare well with experimental date for = 26.6o
and the three water depths, whereas, for = 13.4o ,
agreement is only qualitative. At present the reasons for such discrepancies are not clear; however,
the waves height being smaller for the second case,
viscous effects on the bottom could be important.
More investigation is needed in the future. Plunging jet shapes, angle and velocity of impact compare
well with experimental data provided by Waniewski
(1999). Cross sections of vorticity field at regions
where the water jet starts to form, and just after
the impact of water jet happened were presented;
similarities with experimental observation made by

Dong, R. R., Katz, J., and Huang, T. T. (1997).


On the Structure of Bow Waves on a Ship Model.
J. Fluid Mech., 347:77115.
Favini, B., Broglia, R., and Di Mascio, A. (1996).
Multigrid Acceleration of Second Order ENO
Schemes from Low Subsonic to High Supersonic
Flows. Int. J. Num. Meth. Fluids, 23:589606.
Harten, A., Engquist, B., Osher, S., and
Chakravarthy, S. R. (1987). Uniformly High
Order Accurate Essentially NonOscillatory
Schemes. J. Comput. Phys., 71:231303.
Jameson, A., Schmidt, W., and Turkel, E. (1981).
Numerical Solutions of the Euler Equations by Finite Volume Methods Using RungeKutta Time
Stepping Schemes. AIAA paper, 811259.
Noblesse, D., Hendrix, D., and Kahn, L. (1991).
Nonlinear Local Analisys of Steady Flow About
a Ship. J. Ship Research, 35:288294.
Olgivie,
F. (1972).
The Wave
erated by a Fine Ship Bow.
9th Symposium of Naval Hydrodynamic.
12

Copyright National Academy of Sciences. All rights reserved.

GenIn

Twenty-Fifth Symposium on Naval Hydrodynamics

Osher, S. and Sethian, J. A. (1988). Fronts Propagating with CurvatureDependant Speed: Algorithms Based on HamiltonJacobi Formulations.
J. Comput. Phys., 79:1240.
Rouy, E. and Tourin, A. (1992).
A Viscosity Solutions Approach to ShapefromShading.
SIAM. J. Numer. Analy., 29:867884.
Spalart, P. R. and Allmaras, S. R. (1994). A One
Equation Turbulence Model for Aerodynamic
Flows. La Recherche Aerospatiale, 1:521.
Sussman, M., Smekerda, P., and Osher, S. J.
(1994). A Level Set Approach for Computing Solutions to Incompressible TwoPhase Flow.
J. Comput. Phys., 114:146159.
Waniewski, T. A. (1999). Air Entrainment by Bow
Waves. Doctoral Dissertation, California Institute
of Technology, Pasadena, CA.
Waniewski, T. A., Brennen, C. E., and Raichlen, F.
(2002). Bow Wave Dynamics. J. Ship Research,
46(1):115.

13

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Larry J. Doctors
The University of New South Wales, Australia
I would like to congratulate the authors for
this very interesting paper. My question relates to
Table 1, listing the test conditions. Why werent
exactly the same values of the water depth and speed
chosen for the two different wedge half angles? It
would have been interesting to examine the linearity
(or otherwise) of the wedge-flow contact lines in
Figure 11 with respect to the wedge half angle. Since
the data is almost the same for the two angles of
13.4 and 26.6, one can see, at least approximately,
such a behavior in Figure 11.
AUTHORS REPLY
Thank you for the comment. With regard to
the choice of flow parameters, the only reason why
the water depths were not exactly the same is that we
wanted to validated the simulations against the
experimental data collected by Waniewski et al.
(2002), and therefore we used the same values as in
their channel and towing tank tests. The investigation
on the behaviour of the maximum wave height on the
side wall with the dihedral angle will be the topic of
our future research activity on wave breaking. We
agree with Dr. Doctors as to the expected linearity for
the aforementioned function.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics


St. Johns, Canada, 8-13 August 2004

A BEM-Level set Domain Decomposition for


Violent Two-phase Flows in Ship Hydrodynamics
G. Colicchio , M. Greco , O.M. Faltinsen

( INSEAN, The Italian Ship Model Basin, Roma - Italy, Centre for Ships and
Ocean Structures, NTNU, Trondheim - Norway)
ABSTRACT
In many practical circumstances relevant in ship hydrodynamics, complex water flow evolutions occur, involving large deformations of the free-surface, breaking and fragmentation. The resulting water-ship interactions may lead to dangerous structural loads and they
can concern the safety and the stability of the vessel.
The analysis of such problems is challenging both in
terms of the capability to handle the physics involved,
and of the CPU time and memory space needed for sufficiently accurate investigations. The latter become crucial when three-dimensional effects are accounted for.
Here we present a numerical study to deal with
these types of phenomena, their occurrence, evolution
and possible structural effects. The used method is based
on the Domain Decomposition (DD) philosophy where
the problem is split into sub-problems, each one analyzed by the most efficient and suitable solver. In particular, a Boundary Element Method (BEM) is used in
the fluid regions where the air-water interface can be
modeled as a smooth surface and vorticity and viscosity effects are negligible. A Navier-Stokes (NS) solver,
coupled with a Level-set (LS) technique for capturing
the air-water interface, is applied in the fluid areas interested by breaking phenomena and vorticity generation.
Since the air evolution becomes relevant in case of air
entrainment, both liquid and gas phases are simulated
by the BEM and NS-LS solvers. The most important
feature of the DD method proposed here is the presence
of an air-water interface in the exchange region between
the different solvers.
In this paper, the features of the present domain-

decomposition strategy are described and the challenges


connected with the coupling are deeply discussed. The
numerical investigation highlighted the importance of a
proper rational study when CFD methods are considered. In the present case a crucial aspect is represented
by the domain composition (DC) step, where the information from one solver to the other have to be properly reconstructed and made consistent with the receiver
sub-domain. Such aspects are detailed analyzed by using the dam-breaking problem as test case.
INTRODUCTION
Many problems in ship hydrodynamics can be accurately and efficiently analyzed by assuming the water as
an inviscid and incompressible fluid in irrotational motion. On the other hand, several cases exist where the
potential flow theory may only partially help the physical investigation or where such model is not even able
to capture the fundamental mechanisms involved.
The water-on-deck problem is an example. In
this case one can use the potential flow theory to understand and predict the phenomenon, and to identify the
role played by many physical parameters. However the
most dangerous water-shipping events from the safety
and stability perspectives are connected with free-surface breaking, air cushioning and local water-structure
impacts. Only some of these phenomena can be handled
by a potential flow model. The whole water-on-deck
scenario is out of the BEM capabilities. For instance,
when a wave breaking event occurs circulation is created connected with the cavity closure caused by the
water-water impact. Non-zero circulation means that

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

vorticity is created. This occurs in the neighborhood of


the impact region. The vorticity originates due to the
different tangential velocities of the interacting portions
of water (see i.e. Dooley et al. 1997). Such vorticity
will be partially convected with the rotating cavity and
partially with the jet splash-up subsequent to the impact.
On a sufficiently long time scale, viscous effects will
matter. For a deeper understanding of the phenomenon,
two- and three-dimensional water-shipping experiments
have been recently performed by the authors (see i. e.
Greco 2001 and Barcellona et al. 2003). The tests highlighted the main features of the typical water on deck
phenomenon: (i) an initial plunging phase, quite localized in space and time, where the water enters the
ship in the form of a plunging jet hitting the deck; (ii) a
dam-breaking like behaviour, where no relevant breaking phenomena occur and the water propagates along
the deck similarly to the flow generated by a dam breaking; (iii) water-deck structures impacts, where the compact mass of water hits against obstacles on its way and
rises along them; (iv) a backward plunging phase, when
the raised liquid falls down under the gravity action
and forms a plunging jet hitting the underlying water;
and (v) the off-deck flow, where the water is eventually
pushed out from the deck and falls again into the sea.
The utility of a BEM solver for understanding some individual phenomena was confirmed. On the other hand,
the tests pushed toward a more general method able to
capture the complete water-on-deck picture. For a longtime investigation of the water shipping phenomena, a
rotational viscous two-phase (air and water) model has
to be considered. However, the regions where the potential flow theory cannot be applied are quite confined
near the ship. The vorticity created during the later
stages of the water-on-deck (i.e. due to the breaking of
a backward plunging jet) will be convected by the offdeck flow and eventually diffused going far from the
ship. This means its modeling becomes unimportant at
a sufficient distance from the vessel. From there on, the
flow evolution can be accurately described by an irrotational inviscid model.
Another example where the potential flow theory is not generally suitable is given by the roll motion
of a ship. This is a relevant problem for ship hydrodynamics and has been investigated both numerically
and experimentally. Extensive studies of the connected
phenomena have been carried out for instance by Yeung and coauthors (see i.e. Yeung et al. 1996 and Ye-

ung et al. 2000). In this case the potential theory could


be used in combination with an adequate model for the
vorticity shed during the motion. This implies some geometric assumptions and does not account for the viscous effects that eventually will appear. On the other
hand, the use of a Navier-Stokes (NS) solver to simulate
the roll motion of a ship in open water is challenging.
Some preliminary test cases have been performed asz
air
x

water

Potential flow

air

Potential flow

water

Potential flow

Figure 1: Roll motion of the section 3 of the DDG41


ship, scale factor 23.824. Top: mesh definition. Center: vorticity contour levels and enlarged view of the
velocity field near the lower tip of the cross-section
at a rotation angle
. Bottom: possible domain-decomposition strategy. Roll parameters: rotation
amplitude
and frequency
Hz.



 



suming that the flow is laminar. To limit the code heaviness a stretching has to be introduced both in the horizontal and vertical directions. However the stretching
cannot be particularly strong, that is
,
max

Copyright National Academy of Sciences. All rights reserved.



 

Twenty-Fifth Symposium on Naval Hydrodynamics



 

being
the outer and inner discretizamax and
tions, respectively, and a number less than two. If a
larger stretching is used, numerical problems arise connected with the solution of the Poisson equation for the
pressure. As a result, two factors exist working against
each other: the need of a sufficiently high resolution to
capture the vorticity generated during the ship rolling
and the need to resolve with a similar accuracy both the
inner region around the ship and outer areas far from
the vessel. An example of feasible stretched grid is
given in the top of figure 1 for the roll motion of a ship
cross-section (black line). The geometry and the motion
data refer to two-dimensional model tests carried out recently at INSEAN. In particular, figure 1 is related to the
section 3 (American nomenclature) of the DDG41 ship
(scale factor 23.824). Despite the fact we want to study
open water conditions the domain has been truncated
for computational limits reasons. A uniform grid with
, being the cross-section
draft, is used in an inner region around the body and a
stretching
with
is introduced
elsewhere. The resulting mesh is fairly able to capture
the evolution of the vorticity shed from the ship (the
center plot of the figure gives a snapshot of the vorticity contours). But an enlarged view of the velocity field
near the lower tip of the cross section (see center plot of
figure 1) highlights the need of a much higher resolution
in the body neighbourhood. To meet this requirement
and to capture properly the behaviour of the boundary
layer we should at least consider an inner mesh with
. Due to the stretching
limits this would increase the memory space of a factor 16 and the CPU time of a factor 16x4 with respect
to the used grid. Obviously if three-dimensional effects
are accounted for, the computational cost becomes substantially more expensive. On the other hand this can be
conveniently reduced by using physical considerations.
During the roll motion, vorticity of opposite sign is created. This may cause vortex pairs with nearly opposite
strengths that will travel far from the ship. Once away
from it, their effect on the ship pressure loads is however not significant. This means, also in the case of the
roll motion the region interested by vorticity and viscous effects relevant for the ship is rather localized near
the vessel. From an adequate distance from the vessel
on, the vorticity leaving the ship can be conveniently
neglected (dissipated numerically) and a potential flow
theory can be used to simulate the flow evolution.

   !#" $   %  


&('*) +
 ,-   $

   .#"*    




In the aim of developing a suitable and efficient


method to attack such types of problems, we have developed a domain-decomposition strategy. With this approach, for instance, the roll motion could be investigated by applying a Navier-Stokes solver within a limited region around the body. A potential flow model
could be used instead to analyze the rest of the fluid
domain (see i.e. bottom plot of figure 1). As a consequence, it would be feasible to achieve a much higher
resolution where needed. The present approach is the
result of previous research efforts (see i.e. Greco et al.
2002) where different domain-decomposition strategies
have been investigated. The fluid domain is split in two
(many) sub-domains where the problem is attacked by
different solvers. The solvers will exchange information across the common boundary portions. In our case,
these are characterized by an overlapping area (transmission boundary) where the involved solvers are contemporary used to solve the problem. This brings additional free parameters into the problem, that are (i)
the position and (ii) the extension of the transmission
boundary. In Greco et al. (2002) the DD strategy was
based on the use of a BEM and a NS solver combined
with a VOF method to capture the air-water interface.
The latter technique follows the evolution of the water
volume fraction in the domain and from this information reconstructs the free-surface configuration at any
time instant. The investigation confirmed the capabilities of the approach but highlighted the limits of the
used field solver that was a single-phase solver accurate
to the first order in space and time. Here the DD strategy
is applied by coupling a BEM method with a NS solver
combined with a Level-set (LS) technique to track the
free surface. The latter method follows the evolution of
the distance from the air-water interface and from this
information reconstructs the free-surface configuration
as the surface with
at any time instant. Both potential and field solvers are accurate to the second order
in space and time and simulate both water and air evolutions so that their interaction during air entrainment
phenomena can be investigated.

/0

In this paper the developed domain decomposition strategy is detailed described, the main challenges
related with the substantial differences between the used
solvers are discussed and the proposed solutions are reported. This is made by using the two-dimensional problem of a dam breaking as test case. In fact this is indirectly related to the water-on-deck phenomenon (see

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

i.e. Greco 2001). Moreover it is a well know problem


extensively studied both analytically, numerically and
experimentally. Therefore many reference results are
available in literature.
MATHEMATICAL MODEL
The present method is based on a Domain Decomposition (DD) strategy. A Boundary Element Method (BEM)
is used in the fluid regions with no fragmentation of the
free surface and negligible vorticity generation and viscous effects. In the fluid areas where the latter phenomena become relevant a Navier-Stokes (NS) solver
is introduced. This is coupled with a Level-set (LS)
technique to capture the free-surface evolution. In the
potential and viscous domains, both air and water are
described but with different assumptions. The aim is to
built up a composite solver accurate and efficient. In
both sub-domains no surface tension is considered and
the air is assumed incompressible like the water. The
latter assumption is not correct when air cavities sufficiently small are entrapped. However this simplification has been used during the development phase of the
present domain decomposition due to the smaller computational effort required. The strategy has been applied
to two-dimensional cases, but no restrictions exist for
the extension of the method to treat three-dimensional
flows. The main features of the two solvers used within
the DD and the adopted strategy are described in the
following sections.

with linear shape functions both for the geometry and


for the boundary data and with collocation points at the
edges of each element. This is combined with a second order Runge-Kutta scheme for the time integration.
For more details refer to Greco (2001). By using the
Mixed Eulerian-Lagrangian approach, at any time instant a Dirichlet-Neuman boundary value problem for
the velocity potential has to be solved in the water
domain. is known along the free surface while its normal derivative is constrained along the solid boundaries.
The continuity of the velocity potential is assumed at
the contact points between free surface and body. Once
the problem in water has been solved, the one in air
can be attacked. Enforcing the continuity of the normal
velocity and of the pressure across the air-water interface, the kinetic problem for is fully solved. This can
be successfully made in the case of exterior problems
(i.e. the water-on-deck problem). While, if an interior
problem has to be considered, the things become different (i.e. the dam-breaking problem inside an enclosed
tank). In this case, large pressure gradients occur in
reality near the air-water interface and they cannot be
recovered by enforcing explicitly the pressure continuity. Therefore a Neuman problem in the air domain has
to be settled. The pressure in the BEM sub-domain is
obtained from the Bernoulli equation. This requires the
time derivative of the velocity potential , evaluated in
the present study by solving a similar problem as for
(see i.e. Greco 2001).

132

NS-LS sub-domain
BEM sub-domain
Here a two-dimensional fluid is considered, evolving in
time according to the potential flow theory. The water is
assumed unaffected by the air. This is a reasonable approximation. In fact it is meant to be applied in fluid regions where air-cushioning does not occur and the free
surface can be modeled as a smooth simply-connected
surface. Therefore one can neglect the air-water coupling and assume that the heavier fluid (water) drives
the motion of the lighter one (air). This leads to a more
efficient and yet suitable method. In this sub-domain
the air-water interface is seen as a sharp surface across
which discontinuities of the tangential velocity component occur.
The problem is solved numerically by a BEM

Here a two-dimensional incompressible fluid is assumed


and turbulence effects are neglected. A Projection method is applied to the governing equations. This means
that the solution of the velocity field is the result of
a two-step procedure. In the first step, a guess solution guess is evaluated from the Navier-Stokes equations. In the second one, a correction corr is obtained
by enforcing that the total velocity
guess
corr
is divergence free. This leads to a Poisson equation
for the pressure, which represents the new governing
equation. The problem is solved numerically by a finite difference scheme combined with a a second order
predictor-corrector method for the time integration. A
Level-set technique is combined with the field method
to capture the air-water interface. Therefore the freesurface configuration is reconstructed in time by means

54

Copyright National Academy of Sciences. All rights reserved.

54

54
5 4 6
54

7 54

Twenty-Fifth Symposium on Naval Hydrodynamics

of a scalar function, Level-set function, representing the


local normal distance (with sign) from the interface. For
the details of the used method refer to Colicchio et al.
(2003). Within the LS technique the fluid variables (velocity, pressure and density) are smoothed in a narrow
(see sketch
region across the interface, say
in figure 2). To recover the sharp behavior of the air-

/98(:<;>= ?@=*A

smoothing

air
variables

air

wate

r int

erfa

ce

water
variables

Figure 2: Level-set technique.


Sketch of the
iso-contours delimiting the region across the air-water
interface where the smoothing is applied.

water interface should be very small. However it cannot be smaller than a threshold value, otherwise numerical instabilities arise. Typically
is a good
compromise. Special care has to be paid for the proper
smoothing of the variables inside the transition area.
This is because they are linked with each other nonlinearly through the governing equations and inconsistent smoothing can result in unphysical solutions. From
what has been said, inside the NS-LS sub-domain the
air-water interface is seen as a layer with finite thickness. Across it the variables slightly pass from their
definitions in water to the ones in air.

=CBEDF

To be consistent BEM and NS-LS solvers have


been chosen accurate to the same (second) order both in
space and in time. The main challenges of the present
domain decomposition strategy are related to the profound differences between the BEM and NS-LS freesurface definitions. To overcome such difficulties, the
Domain Decomposition (DD) step has to be accompanied by a proper Domain Composition (DC) step. In
the former the fluid domain is split into two (many)
sub-domains solved either by the BEM or by the NSLS solvers. They must interact with each other through
the common boundaries. In the latter step the information exchanged from one sub-domain to the other
must be properly converted to be consistent with the receiver solver. Since this aspect is a crucial element of
the present approach, we named our strategy DomainDecomposition Domain-Composition (DDDC) method.

To built up an efficient solution algorithm, the domaindecomposition strategy is applied both in time and in
space. The related details are described in the following
sections.
Spatial coupling
As far as the BEM can be used to study the problem
of interest, it will be applied. When applicable, this is
indeed the most efficient and accurate method to treat
free-surface flows. Let us assume a threshold time
as the time when it becomes useful to introduce the
domain decomposition strategy (for instance when the
air-water interface is going to break). From this time
instant on the domain is spatially split into two (many)
sub-domains. The problem in one of these is still solved
by the BEM, while the flow evolution in the other is described by the NS-LS solver (see sketch in figure 3). At
time the solution in the NS-LS sub-domain needs to
be initialized by the BEM. Once this is accomplished
the two methods will attack the problem in the corresponding sub-domains and will transmit each other
the required information through the overlapping region
(domain decomposition). The NS-LS solver gives to the

G

GH

Figure 3: Domain Decomposition Domain Composition method. Definition of the transmission region as
an overlapping area through which pressure and velocity data are exchanged. This coupling procedure has
been referred to as procedure b in Greco et al. (2002).

BEM both pressure and velocity components needed to


update and its normal derivative along the transmission boundary. There, also the new free-surface level
becomes available from the NS-LS solver. Similarly,
the NS-LS gets from the BEM pressure and velocity
components, together with the air-water interface level
and its local normal vector. Different sizes of the overlapping area have been tested. Its extension should be as

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

small as possible to reduce the code size but sufficiently


large to avoid that small errors close to the overlapping
region could amplify causing the instability of the couhas been found a
pling. Typically an extent of
good compromise. In this context, it is important that
the boundary and field solvers have similar discretizations near the transmission boundary.

%  

The pressure and velocity data cannot be given


directly from one solver to the other but must be properly converted according to the features of the receiver
solver (domain composition). This reconstruction has
to be applied both at time everywhere in the NS-LS
sub-domain and at any time
near the transmission boundary.

GH

GJIKGH

BEM composition step The only requirement for the


BEM is to recover (i) the interface as a zero-thickness
surface with (ii) sharp variation of the physical variables from one fluid to the other. The first request is
satisfied by taking as free-surface (at the transmission
boundary) the zero Level-set function (
). The

/K.

u,v,p
LSNS

with

5 4 065 4 7 5 4
N Q  N 7 N
Q!R
S NUT MVWV QZR 5U[3\^]
X Y 4
5 4 T @V@V E;`_badjcfehgig
BEM

corr

BEM

corr

(1)

has to be introduced for the velocity and pressure fields,


by applying a technique similar to the one used within
the Projection method. In this way the velocity field
used by the NS-LS solver is divergence free. (iv)
The smoothing process cannot be introduced symmetrically across the interface, as one would think consistent
with the smoothing of the density (see sketch 2). It has
to be consistent with the smoothing of the
function across the interface. This variable is quite relevant
for the robustness and validity of the numerical results
since it appears as a coefficient in the pressure Poisson
equation. Three different
smoothing are given in
figure 5. In all cases the largest
gradients are in the

54 

X

X X


BEM

water

by using the BEM information about the free-surface


configuration and its local normal vector. (ii) The sharp
BEM variables must be smoothed across the interface
consistently with the LS technique. (iii) The velocity
field has to be divergence free not only in the two separate fluid phases but also across the free surface. This
is not ensured by the BEM solution since it does not account for the air-water coupling. Therefore a correction
in the form

air

Figure 4: BEM composition step. Extrapolation procedure for the generic NS-LS variable (velocity compoand pressure ).
nents

5 ?ML

5 ?@L

X

second one is accomplished by sharpening the smooth


Navier-Stokes variables (velocity components
and
pressure ) through an extrapolation procedure in the
neighborhood of the interface (see example in figure 4).

Figure 5: Examples of
smoothing across the
air-water interface that could be used within Level-set
techniques. In the present study the smoothing corresponding to the dashed line has been applied.

NS-LS composition step The NS-LS requirements


can be listed as follows: (i) the Level-set function has
to be properly initialized everywhere at the time and
near the transmission boundary at any
. This will
allow to recover the air-water interface as a transition
layer with finite thickness. Such aim can be achieved

air region. This implies that, to be consistent with the


NS-LS solution, the smoothing of the BEM variables
has not to be enforced symmetrically across the interface. The smoothing area must be shifted toward the air
domain. In the present study we used the curve corresponding to the dashed line in figure 5 as
smoothing. This required a shift of
for the smoothing area.

GOIPG 

G

Copyright National Academy of Sciences. All rights reserved.

 =

X

Twenty-Fifth Symposium on Naval Hydrodynamics

The relevance of the composition step will be


described later in the paper by using the dam-breaking
problem as a test case.

TWO-DIMENSIONAL NUMERICAL STUDIES

Temporal coupling

The DDDC method has been tested by studying the water evolution subsequent to the breaking of a dam. This
is a well known problem, extensively studied for its
relevance in environmental and safety contexts. Moreover it is an interesting problem for ship hydrodynamics. The related flow is similar to the one developing
onto the ship deck during the most common type of
water shipping, as described above. To mimic the impact of the water against deck superstructures, a vertical
rigid wall is introduced downstream the initial dam and
the water-wall interactions during the impact are analyzed. This study is relevant to check the method capabilities in handling the green-water loading.

The exchange of data from one sub-domain to the other


is accomplished in time by means of a real coupling between the two solvers. The related algorithm is given
, the evolution
in figure 6. At any time instant

GlkmIZG 

t n> t0

BEM
tn

t n+1/2

first substep of the


RungeKutta scheme

DD
NSLS

DC
tn
DD

r

t n+1

second substep of the


RungeKutta scheme

DD
DC
DD

t n+1

corrector step

DC
t n= t

z
initial
water level

n+1

Figure 6: Domain Decomposition Domain Composition method. Time coupling algorithm.

Glk Glkn^o

from
to
is achieved by performing two intermediate time integration steps with both solvers. First,
the BEM evolves from
to
and makes available the required variables at the transmission boundary.
These are properly reconstructed (domain composition)
and given to the NS-LS solver. Then, the latter performs a trial time integration from
to
(predictor step) and makes available the information needed by
the potential solver. Once these have been adequately
converted (domain composition), the BEM can make its
time evolution from to
and release the final data
required by the NS-LS solver. These are reconstructed
(domain composition) and used by the field solver that
can perform the final time evolution from
to
(corrector step). At this stage everything is known at
both in terms of fluid variables and geometry. One
can then perform a new integration step where the BEM
uses the data (once reconstructed) given by the NS-LS
solver at the end of the previous time step.

G k G kn^oMpWq

Glk Glk
n3o

Glk Glkn3o

Glkn3o

  r

 D 
 r

NSLS
tn

st

The problem we analyzed is sketched in the left


plot of figure 7. A reservoir of water, high and
long, is protected by a rigid wall on one side (left)
and by a dam on the other one (right). Moreover it
is contained in a rectangular rigid tank high
and
long
. The potential solver is used to analyze the

DC

BEM
tn

t n+1

predictor step

Dam-breaking problem: DDDC features

Glk Glk
n3o

dam

vertical wall

BEM

NSLS

h
x

Figure 7: Dam breaking followed by the impact with


a vertical wall. Left: sketch of the studied problem.
Right: definition of BEM and NS-LS regions. The
free-surface configuration shown is the BEM solution
at
. This solution has been used to initialize the DDDC approach.

Sxw

Gvu  r Y   %

Syw

GH u  r Y   %

problem until
after the dam release.
Then the DDDC strategy is introduced, as shown in the
right plot of figure 7. The BEM sub-domain is conveniently restricted to the left side of the tank, while the
NS-LS solver is applied to study the right sub-domain.
In the simulations the viscosity has been set equal to
zero. Within the dam-breaking problem such parameter
matters for the local details of the flow but it is not relevant from the global point of view. Due to the chosen
value for the free surface is not steep when the DDDC
is initiated. Despite this fact the related flow conditions
imply a high sensitivity of the solution to the numerical
choices. The transmission boundary has been placed

GH

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

  rz    %

from the left tank wall. As a result, a thin


tongue of water enters the NS-LS sub-domain. Its dimensions are comparable with the thickness ( ) of the
smoothing region across the air-water interface. This
means we are treating a challenging coupling between
the two solvers.

=

If one applies naively the domain-decomposition


approach many numerical problems arise. Some will
destroy the solution, others will introduce localized errors more difficult to detect but relevant from phenomenological and dynamic points of view. To avoid such
errors a domain composition step is necessary. If the ve-

G u  r{  %

Figure 8: NS-LS Composition step. Enlarged view of


the air-water velocity field at
, when the
DDDC is started. The plotted flow area corresponds to
the dashed rectangular in the right plot of figure 7. Top:
solution given by the BEM. Also the streamlines are
shown. Center: solution after the symmetric smoothing
across the interface. Also the streamlines are shown.
Bottom: solution after an asymmetric smoothing across
the interface performed inside the two dashed lines.

locity field obtained by the BEM is directly given to the

NS-LS (top plot of figure 8), numerical errors are introduced related to the discrete representation of the velocity. These grow up in time and lead to unbounded oscillations. They are avoided once the velocity field given
by the BEM is smoothed across the interface (center
plot of figure 8). However the flow evolution highlights
the occurrence of small oscillations of the air-water interface and above all an unphysical behavior of the pressure field in a region close to the water tip front (top plot
of figure 9). The latter is a memory effect of the initial
pressure given by the BEM and can be avoided by enforcing divergence free velocities across the interface,
center plot of figure 9. The oscillatory behavior of the
air-water interface nearly detectable from the solution
can be finally eliminated by shifting the smoothing area
toward the air domain (bottom plot of figure 8). In this

G u  r|  } ~

Figure 9: NS-LS Composition step. Water-pressure


contour levels and air-water interface at
.
Top: unphysical pressure contours due to fictitious compressibility at the air-water interface. Also an oscillatory behavior of the free surface can be identified. Center: unphysical oscillatory behavior of the free-surface
due to a non proper smoothing at the air-water interface.
Bottom: final solution once the proper composition step
has been accomplished.

way the velocity field in the water region of the NSLS domain is more consistent with the one given by the
BEM solver. Moreover pressure and free-surface evo-

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

lutions recover the physical behaviors, bottom plot of


figure 9.

  rK#"  r  




The results here shown have been obtained with


. Sima NS-LS discretization
ilar resolution was used for the BEM elements. This
mesh corresponds almost to a convergent solution. Only
locally sensitivity of the chosen grid can be identified,
as discussed later. During the DDDC application the
time step was taken fixed and chosen equal to
, as imposed by the field method stability requirements. No special stability conditions have been introduced to account for the large free surface deformations
in the later flow evolution. However the numerical solution did not show any unstable behavior during such
stages.

Gvu  r|

 


Dam-breaking problem: results


Figure 10 shows the free-surface evolution after the dam
breaking. The DDDC method (dashed lines) is compared with the full NS-LS (solid lines) and with the full
BEM (circles) solutions. To be consistent with the potential solver, the viscosity has been set equal to zero in
the field methods simulations. From the sequence, the
results agree well during the whole flow evolution: before and after the water impact against the vertical wall,
during the water rise-up along the structure and its later
run-down, when a backward plunging is formed eventually hitting the underlying water. The further flow
evolution cannot be handled by the BEM. Differently
both the full NS-LS and the DDDC methods are able
to follow the breaking and splash-up phases. Figure 11
shows the free-surface snapshot at
. The
global agreement is still satisfactory although some local discrepancies can be detected. More in detail in the
shape of the splash-up jet, due to the closeness of the
transmission boundary in the DDDC method. Although
the latter should be just a virtual boundary we cannot
expect it to be perfectly transparent. Some differences
are also visible in the length of the thin layer of water
along the wall and in the shape of the cavity. These
local disagreements may be due to non-perfectly identical discretizations. On the other hand, the discrepancies connected with the film of water along the wall
are unimportant from the structural perspective since
inside the thin layer the pressure is practically atmospheric. Further, the discrepancies related to the cavity
are quite small and result just in slight differences in the

G u  r0 ~

r

s

D D

w Fr   
Gvu  r  ?  ? %  ? 

Figure 10: Dam-Breaking problem (


) and impact
with a vertical wall at
from the dam. Free surface configurations at
and
. Time increases from top to bottom. DDDC solution (dashed lines) is compared with full NS-LS (solid
lines) and full BEM (circles) results. To have consistent
solvers the viscosity has been set equal to zero in the
simulations.

 

structural loads. To the purpose, top plot of figure 12


gives the pressure evolution as evaluated along the vertical wall by the DDDC (squares) and the full NS-LS
(solid line) methods. The pressure has been recorded
at a location
from the bottom. It shows a

  }
}
}Fr

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

p(gh)

-1

NS-LS
DDDC
Zhou exp.

0.5

r
s


D D
  r w
vG u r   ~

Figure 11: Dam-Breaking problem (


) and impact
with a vertical wall at
far from the dam. Free
surface configuration at
. DDDC results
(dashed line) are compared with the full NS-LS solution
(solid line).
first sudden rise during the initial water-wall impact and
a second one during the water run-down phase. The
two results are in satisfactory agreement despite some
tiny differences during the air-cushioning phenomenon
(later stage of the pressure evolution). In the same plot
the experimental pressure by Zhou et al. (1999) is superimposed. In the tests a circular pressure gauge, with
above the bota diameter of mm, centered at
tom and with lowest area point , was used. The reasons for comparing the numerical results with experimental data centered at a different location can be found
i.e. in Greco (2001). Because no information was available about the exact timing of the experimental results,
they have been shifted in time to have the first pick coincident with the numerical one. The experiments are
consistent with the numerical ones although a different
post-breaking behavior can be detected. At this stage
the results are very sensitive to the shape of the air cavity and to the occurrence of three-dimensional effects
during the experiments. The horizontal force acting on
the vertical wall during the water-structure interaction
is given in the bottom of the same figure. The load
reaches a first peak during the initial water-wall impact
(
). After a short time the force starts
to rise again until the occurrence of another substantial
peak, due to the backward plunging hitting the underlying water (
). A secondary intermediate
peak can also be identified. This is due to the water impact with the tank roof. The DDDC and NS-LS methods predict the same magnitudes and durations of the
peaks and agree globally well with each other. A slight
delay in the roof impact can be detected for the DDDC
method, due to a different length in the water layer rising the wall. Moreover, similar small differences as for
the pressure are visible during the air cushioning.

w

&G B  D  u r 

w

GOB u r 

  
  r


0
1

t(gh)

1/2

s   r
D D
  r
  } }
} r

Figure 12: Dam-Breaking problem (


) and impact with a vertical wall at
from the dam. Top:
pressure evolution at a location
from the
tank bottom as obtained by DDDC (squares) and full
NS-LS (solid line) methods. The triangles give the pressure time history measured by Zhou et al. (1999) at
from the bottom
a pressure gauge centered at
and having
as lowest area point. Bottom: horizontal force evolution on the vertical wall as obtained by
DDDC (squares) and full NS-LS (solid line) methods.

  

Fr

Present investigation supplies the use of the DDDC strategy for handling complex free-surface flows involving breaking, air cushioning and impact phenomena. The efficiency properties of the method are also
confirmed. The DDDC results have been obtained by
reducing the CPU time of a factor two and the memory
space of a similar factor with respect to the full NS-LS
method.
CONCLUSIONS
A Domain Decomposition Domain Composition method
has been developed to study flows with large free-surface
deformations and breaking, leading to water-water and
water-structure impacts and air entrainment. The method
can also handle regions with vorticity generation and
viscous effects. It is based on the use of a BEM and a

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

NS solvers, the latter is combined with a LS method


to capture the free surface. The two solvers are accurate to the second order both in space and in time
and are fully coupled consistently with the related time
marching schemes. The resulting composite solver is
governed by the stability properties of the field method
and proved to be convergent under free-surface and grid
refinements. The main challenges connected with the
coupling of two quite different methods have been discussed and the proposed solutions have been described.
The effectiveness of the strategy both in terms of time
saving and of solution convergence has been studied
through the dam-breaking problem. Except for some
local discrepancies, the air-water evolution agreed with
full NS-LS solution during the whole phenomenon from
the dam release until the formation of a splash-up phenomenon, and with the full-BEM solution until the impact of the backward plunging jet with the underlying
water. The local and global structural loads evaluated
along a vertical wall downstream the initial dam are also
satisfactorily predicted. These represent a crucial factor
when CFD methods are considered. The results suggest
that the present DDDC strategy is a promising approach
to investigate complex problems in ship hydrodynamics.
As a next step we plan to include the effects of
air compressibility. These become relevant when air is
entrapped in cavities sufficiently small and should be
properly accounted for. To keep the idea of developing

N  X 
,-  %

valid for ideal gases during adiabatic processes will be


and
are reference values
applied. In equation (2)
for pressure and density (i.e. atmospheric pressure and
normal air density) and
.
Other important research that we plan to perform
is the development of an adequate and efficient dynamic
spotting of the domain-decomposition strategy. This
means to identify proper parameters and related threshold values to guide the initiation of the domain decomposition. Figure 14 shows two examples. In the first

Figure 14: Strategy to perform an efficient dynamic


spotting.
case (left), the key parameter is represented by the freesurface steepness. The DD can be switched on when
the curvature exceeds locally a threshold value. In the
second case (right), the key factor is the vicinity of the
air-water interface to a structure enclosing a cavity. The
DD can be switched on when the distance of the free
surface from the wall (excluding the water-wall intersection point) becomes smaller than a threshold value.
ACKNOWLEDGEMENTS

p = f( )

D /D t = 0

Present research activity is partially supported by the


Centre for Ships and Ocean Structures, NTNU, Trondheim, within the Green Water Events and Related Structural Loads project, and partially done within the framework of the Programma di Ricerca sulla Sicurezza
funded by Ministero Infrastrutture e Trasporti.

Figure 13: Strategy to account for the air compressibility in an efficient manner.
REFERENCES
an efficient and suitable method, we plan to model the
air compressibility only where necessary, that is inside
enclosed cavities (see figure 13). Elsewhere the air will
be treated still as an incompressible fluid like the water.
Therefore the Poisson equation will be solved for the air
pressure everywhere except for inside cavities. There,
the equation
(2)

NKN X   X )

Barcellona, M., M. Landrini, M. Greco, and


O. Faltinsen. An Experimental Investigation on Bow
Water Shipping. Journal Ship Research 47(4), pp. 327
346, 2003, Dec.).
Colicchio, G., A. Colagrossi, M. Greco, and M. Landrini. Freesurface Flow After a Dam break: A Comparative Study. Proc. of 4th Numerical Towing Tank

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Symposium (NuTTS). Hamburg, Germany, 2001.


Colicchio, G., M. Landrini, and J. Chaplin. Level-set
modelling of the two-phase flow generated by a surface
International Workshop on Water
piercing body.
Waves and Floating Bodies. Le Croisic, France, 2003.

~ 2f

Dooley, B., A. Warncke, M. Gharib, and G. Tryggvason. Vortex ring generation due to the coalescence
of a water drop at a free surface. Experiments in Fluids 22(5), pp. 369374, 1997.
Greco, M. A Two-dimensional Study of Green-Water
Loading. Ph. D. thesis, Dept. Marine Hydrodynamics,
NTNU, Trondheim, Norway, 2001.
Greco, M., O. M. Faltinsen, and M. Landrini. Water Shipping on a Vessel in Head Waves. Proceedings
24 Symposium on Naval Hydrodynamics, technical
session: Slamming, Green Water and Capsizing, pp. 1
14. Fukuoka, Japan, 2002.

2f

Yeung, R., C. Cermelli, and S. Liao. Vorticity Fields


Due to Rolling Bodies in a Free Surface - Experiment
and Theory. Proceedings 21 Symposium on Naval
Hydrodynamics. Trondheim, Norway, 1996.

2f

Yeung, R., D. Roddier, L. Alessandrini, L. Gentaz,


and S. Liao. On Roll Hydrodynamics of Cylinders Fitted with Bilge Keels. Proceedings 23 Symposium on
Naval Hydrodynamics. Val de Reuil, France, 2000.

2f

Zhou, Z. Q., J. Q. D. Kat, and B. Buchner. A nonlinear 3-d approach to simulate green water dynamics
on deck. Piquet (Ed.), Proc. 7 Int. Conf. Num. Ship
Hydrod., pp. 5.11, 15. Nantes, France, 1999.

2f

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Krish Thiagarajan
The University of Western Australia, Australia
The pressure impact curves shown in the
overheads are more spread out and less localized than
the pressure spikes seen in classic dam break problem
(shown in Figure 12 of the paper). Can you explain
the physical reason for the differences?
AUTHORS REPLY

Figure 2: Figure 12 of the paper.

There are many reasons for the different


pressure profiles.
First, the problems are qualitatively similar,
but the flow conditions that determine the impact
against the vertical wall and the successive plunging
breaking are not the same.
Second, the two scales are different and this
is important during the air-entrainment phase.
Third, the shape of the of the plunging is
very different, in the dam-break case it is very well
defined and it is more energetic, in the considered
water on deck event, the amount of water involved in
the plunging formation is smaller and the successive
impact on the water is less intense.
Fourth, the time intervals plotted into the
two figures are quite different when referred to the
time scale of the impact phenomena.

Figure 1: Experimental (lines with symbols) and


numerical (continuous line) pressure evolution along
the wall for the water on deck problem.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Jinzhu Xia
Australian Maritime College, Australia
What is the perspective of coupling this
methodology to a seakeeping computation?
AUTHORS REPLY
The coupling, by itself, is well suitable for a
sea-keeping analysis. The challenge is not in the
coupling of the two different codes, but in the
individual development of each solver, both to treat
3D flows and to allow the free motion of the body.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics


St. Johns, Newfoundland and Labrador, CANADA, 8-13 August 2004

The numerical simulation of ship waves using cartesian grid


methods with adaptive mesh refinement
Douglas G. Dommermuth1 , Mark Sussman2 , Robert F. Beck3 , Thomas T. OShea1 ,
Donald C. Wyatt1 , Kevin Olson4 , and Peter MacNeice5
1

Naval Hydrodynamics Division, Science Applications International Corporation,


10260 Campus Point Drive, MS 34, San Diego, CA 92121
2
Department of Mathematics, Florida State University, Tallahassee, FL 32306
3
Department of Naval Architecture & Marine Engineering, University of Michigan,
209 NA&ME Building, 2600 Draper Road, Ann Arbor, Michigan 48109
4
University of Maryland at Baltimore County, Code 931, NASA/GSFC, Greenbelt, MD 20771,
5
Drexel University, Code 931, NASA/GSFC, Greenbelt, MD 20771

Abstract
Cartesian-grid methods with Adaptive Mesh Refinement
(AMR) are ideally suited for simulating the breaking of
waves, the formation of spray, and the entrainment of air
around ships. As a result of the cartesian-grid formulation, minimal input is required to describe the ships geometry. A surface panelization of the ship hull is used
as input to automatically generate a three-dimensional
model. No three-dimensional gridding is required. The
AMR portion of the numerical algorithm automatically
clusters grid points near the ship in regions where wave
breaking, spray formation, and air entrainment occur.
Away from the ship, where the flow is less turbulent,
the mesh is coarser. The numerical computations are implemented using parallel algorithms. Together, the ease
of input and usage, the ability to resolve complex freesurface phenomena, and the speed of the numerical algorithms provide a robust capability for simulating the
free-surface disturbances near a ship. Here, numerical
predictions, with and without AMR, are compared to experimental measurements of ships moving with constant
forward speed, including a vertical strut, the DDG 5415,
and a wedge-like geometry.

Introduction
Two different cartesian-grid methods have been developed to simulate ship waves. One technique (CLSVOF)
combines Level-Set (LS) techniques with Volume-ofFluid (VOF) methods to model the free-surface interface. The second technique uses a pure VOF formulation. The CLSVOF formulation uses Adaptive Mesh
Refinement (AMR) to resolve small-scale features in the
flow. The VOF formulation uses domain decomposition
without AMR. Both methods that are described in this
paper use the same panelized geometry that is required

by potential-flow methods to automatically construct a


signed-distance-function representation of the hull (see
Sussman & Dommermuth (2001)). The hull representation is then immersed inside a cartesian grid that is used
to track the free-surface interface. No additional gridding
beyond what is already used by potential-flow methods
is required. The CLSVOF formulation is used to investigate the flow around the DDG 5415, and the pure VOF
formulation is used to model the flow around a vertical
strut and a wedge-like geometry. In all cases, comparisons are made to experiments.
The Numerical Flow Analysis (NFA) code is meant
to provide a turnkey capability to model breaking waves
around a ship, including both plunging and spilling
breaking waves, the formation of spray, and the entrainment of air. Cartesian-grid methods are used to model
the ship hull and the free surface. Following Goldstein,
Handler & Sirovich (1993) and Sussman & Dommermuth (2001), a body-force method is used to enforce
a no-slip boundary condition on the hull. Based on
Colella, Graves, Modiano, Puckett & Sussman (1999),
the ability to impose free-slip boundary conditions is also
provided. A surface representation of the ship hull is
used as input to construct a three-dimensional representation of the ship hull on a cartesian grid. The interface capturing of the free surface uses a second-order
accurate, VOF technique. At each time step, the position of the free surface is reconstructed using piecewise planar surfaces as outlined in Rider, Kothe, Mosso,
Cerutti & Hochstein (1994). Based on Iafrati, Olivieri,
Pistani & Campana (2001), the in-flow and out-flow
boundary conditions use a body-force technique to enforce a uniform stream with no free-surface disturbance
ahead of and behind the ship. A second-order, variablecoefficient Poisson equation is used project the velocity onto a solenoidal field thereby ensuring mass conservation. A preconditioned conjugate-gradient method
is used to solve the Poisson equation. The convective

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

interface. These challenges are not unlike those facing


experimentalists in the laboratory and in the field. In
particular, techniques are required to analyze unsteady
effects, including the formation of bubbles and spray.
Here, we propose various statistical approaches. For example, two approaches are proposed for analyzing the
free-surface elevation as predicted by VOF formulations.
The first technique reconstructs the free surface using the
average of the volume fractions over time. The second
technique takes the mean and variance over time of the
zero-crossings throughout a column of fluid. The first
technique is useful for predicting the mean surface elevation. In addition, it provides a prediction of air entrainment beneath the surface and droplet formation above the
surface. The second technique forms the basis for investigating the variance in the free-surface elevation. Generally speaking, high variance indicates regions where
either bubbles are entrained or droplets are shed. These
approaches and their nuances are discussed in greater detail in the results section.

terms in the momentum equations are accounted for using a slope-limited, third-order QUICK scheme as discussed in Leonard (1997). Based on the PARAMESH
suite of codes (MacNeice, Olson, Mobarry, deFainchtein
& Packer 2000), domain decomposition is used to solve
the field equations. PARAMESH controls data communication between blocks of grid points, and also between
computer processors. PARAMESH is written in Fortran
90. PARAMESH provides AMR capability, but here we
only illustrate the NFA code using uniform grid spacing
without adaptive meshing. (An AMR capability for the
NFA code is in progress.) On the Cray T3E, message
passing is accomplished using either the Cray SHMEM
library or MPI. The CPU requirements are linearly proportional to the number of grid points and inversely proportional to the number of processors. For the NFA code,
comparisons are made to measurements of flow around
a vertical strut (Zhang & Stern 1996)and a wedge-like
geometry (Karion, Waniewski-Sur, Fu, Furey, Rice &
Walker 2003).
Developed concurrently with the NFA code, another
code based on the Coupled Level set and Volume-ofFluid (CLSVOF) method has been developed for modelling free-surface flows in general geometries. The
CLSVOF code uses adaptive mesh refinement to compute multi-scale phenomena. Like the NFA code, the
CLSVOF code uses cartesian grid techniques to model
complex geometries. Also, like NFA, CLSVOF uses a
two-phase formulation of the air-water interface. Unlike the NFA code, which is based on PARAMESH,
the CLSVOF code is based on BOXLIB, which is developed by the CCSE group at Lawrence Berkeley National Laboratories. The strategy of BOXLIB is that
high-level adaptive gridding and parallel functions are
performed using C++ while numerical discretizations
of the Navier-Stokes equations are performed using a
FORTRAN code. The BOXLIB libraries take care of
all the dynamic gridding functions, whereas the user
only has to supply FORTRAN routines that operate on
fixed, uniform, rectangular grids. Please refer to the
work of Rendleman, Beckner, Lijewski, Crutchfield &
Bell (2000) for more information regarding BOXLIB.
For computation of incompressible flow on an adaptive
grid, it is not enough to insure that fluxes are matched
at coarse/fine grid boundaries. We must also compute a
composite projection step at each time step. A composite projection step insures that the pressure, velocity, and divergence-free condition, are satisfied across
coarse-fine grid boundaries. For details of our adaptive
implementation, we refer the reader to Sussman (2003b)
and the references therein. CLSVOF predictions are
compared to measurements of the flow around the DDG
5415 (see http://www50.dt.navy.mil/5415/).

Formulation
Consider turbulent flow at the interface between air and
water. Let ui denote the three-dimensional velocity field
as a function of space (xi ) and time (t). For an incompressible flow, the conservation of mass gives
ui
=0 .
xi

(1)

ui and xi are normalized by Uo and Lo , which denote the


free-stream velocity and the length of the body, respectively.
Following a procedure that is similar to Rider et al.
(1994), we let denote the fraction of fluid that is inside
a cell. By definition, = 0 for a cell that is totally filled
with air, and = 1 for a cell that is totally filled with
water.
The convection of is expressed as follows:
Q
d
=
,
dt
xj

(2)

where d/dt = /t + ui /xi is a substantial derivative. Q is a sub-grid-scale flux which can model the entrainment of gas into the liquid. Details are provided in
Dommermuth, Innis, Luth, Novikov, Schlageter & Talcott (1998).

CLSVOF and VOF formulations pose unique challenges associated with data processing of the free-surface

2
Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Let ` and ` respectively denote the density and


dynamic viscosity of water. Similarly, g and g are the
corresponding properties of air. The flows in the water
and the air are governed by the Navier-Stokes equations:
dui
dt

1 P
1
+
(2Sij )
xi
Re xj
ij
1
2 i3 +
,
Fr
xj

where Ri denotes the nonlinear convective, hydrostatic,


viscous, sub-grid-scale, and body-force terms in the momentum equations. uki and k are respectively the velocity components at time step k. t is the time step. P is
the first prediction for the pressure field.

= Fi

For the next step, this pressure is used to project the


velocity onto a solenoidal field. The first prediction for
the velocity field (ui ) is


1 P
ui = uki + t Ri
(8)
(k ) xi

(3)

where Re = ` Uo Lo /` is the Reynolds number and


Fr2 = Uo2 /(gLo ) is the Froude number. g is the acceleration of gravity. Fi is a body force that is used to impose
boundary conditions on the surface of the body. P is
the pressure. ij is the Kronecker delta symbol. As described in Dommermuth et al. (1998), ij is the subgridscale stress tensor. Sij is the deformation tensor:


1 ui
uj
Sij =
+
.
(4)
2 xj
xi

The volume fraction is advanced using a volume of fluid


operator (VOF):

= k VOFi uki , k , t
(9)
A Poisson equation for the pressure is solved again during the second stage of the Runge-Kutta algorithm:


1 P k+1

ui + uki

+
R
=
(10)
i
xi ( ) xi
xi
t

and are respectively the dimensionless variable densities and viscosities:


() = + (1 )H()
() = + (1 )H() ,

ui is advanced to the next step to complete one cycle of


the Runge-Kutta algorithm:



1
1 P k+1

k
uk+1
=
u
+
u
+
t
R

, (11)
i
i
i
i
2
( ) xi

(5)

where = g /` and = g /` are the density and viscosity ratios between air and water. For a sharp interface,
with no mixing of air and water, H is a step function. In
practice, a mollified step function is used to provide a
smooth transition between air and water.

and the volume fraction is advanced to complete the algorithm:




ui + uki k
, , t
(12)
k+1 = k VOFi
2

As discussed in Dommermuth et al. (1998), the divergence of the momentum equations (3) in combination with the conservation of mass (1) provides a Poisson
equation for the dynamic pressure:
1 P
= ,
xi xi

Details of the CLSVOF numerical time-integration


procedure are provided in Sussman (2003a).

(6)

E NFORCEMENT OF B ODY B OUNDARY C ONDITIONS

where is a source term. As shown in the next section, the pressure is used to project the velocity onto a
solenoidal field.

Two different cartesian-grid methods are used to


simulate the flow around surface ships. The first technique imposes the no-flux boundary condition on the
body using a finite-volume technique. The second technique imposes the no-flux boundary condition via an external force field. Both techniques use a signed distance
function to represent the body. is positive outside
the body and negative inside the body. The magnitude of
is the minimal distance between the position of and
the surface of the body. is calculated using a surface
panelization of the hull form. Greens theorem is used to
indicate whether a point is inside or outside the body, and
then the shortest distance from the point to the surface of
the body is calculated. Details associated with the calculation of are provided in Sussman & Dommermuth
(2001).

N UMERICAL T IME I NTEGRATION


Based on Sussman (2003a), a second-order RungeKutta scheme is used to integrate with respect to time the
field equations for the velocity field. Here, we illustrate
how a volume of fluid formulation is used to advance the
volume fraction function (see, for example, Rider et al.
(1994)). During the first stage of the Runge-Kutta algorithm, a Poisson equation for the pressure is solved:
 k


1 P

ui
=
+
R
,
(7)
i
xi (k ) xi
xi t
3

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

CLSVOF method

Free-slip conditions

In the CLSVOF algorithm, the position of the interface is updated through the level-set equation and the
volume-of-fluid equation. After, the level-set function
and the volume fractions have been updated, we couple
the level-set function to the volume fractions as a part
of the level-set reinitialization step. The level-set reinitialization step replaces the current value of the level-set
function with the exact distance to the VOF reconstructed
interface. At the same time, the VOF reconstructed interface uses the current value of the level-set function to determine the slopes of the piecewise linear reconstructed
interface. For more details of the CLSVOF algorithm, including axisymmetric and three-dimensional implementations, see Sussman & Puckett (2000).

In the finite volume approach, the irregular boundary (i.e. ship hull) is represented in terms of along
with the corresponding area fractions, A, and volume
fractions, V . V = 1 for computational elements fully
outside the body and V = 0 for computational elements
fully inside the body. Once the area and volume fractions
have been calculated, they are used in the Poisson equation for the pressure and in the projection of the velocity
onto a solenoidal field. Through the Poisson equation
and the projection operator, the component of velocity
that is normal to the ship hull is set to zero. This corresponds to imposing free-slip conditions on the hull form.
Details associated with the calculation of the area and
volume fractions are provided in Sussman & Dommermuth (2001) along with additional references.

E NTRANCE AND E XIT B OUNDARY C ONDITIONS


No-slip conditions

Entrance and exit boundary conditions are required


in order to conserve mass and flux. Two techniques are
considered. The first technique uses a body force, and
the second technique uses a special formulation of the
pressure.

The boundary condition on the body can also be imposed using an external force field. Based on Dommermuth et al. (1998) and Sussman & Dommermuth (2001),
the distance function representation of the body () is
used to construct a body force in the momentum equations. As constructed, the velocities of the points within
the body are forced to zero. For a body that is fixed in a
free stream, this corresponds to imposing no-slip boundary conditions.

Body-force method
Body forces are used in the momentum equations
(see Equation 3) and the convection equation for the volume fraction (see Equation 2) to force conservation of
flux and mass. For the velocities, a parallel flow with
(u, v, w) = (1, 0, 0) is forced at the entrance and exit.
For the volume fraction, the mean surface elevation is
forced to be zero at the entrance and exit. A similar procedure is used by Iafrati et al. (2001) in their level-set
calculations of two-dimensional breaking waves over a
hydrofoil. The body-force is prescribed as follows:

I NTERFACE C APTURING
Two methods are presented in our work for computing ship flows. Both methods use a front-capturing
type procedure for representing the free surface separating the air and water. The first technique is based
on the Volume-of-Fluid (VOF) method, and the secondtechnique is based on the Coupled volume-of-fluid and
level-set method (CLS).

Fi (x, t) = Fo T (x) (ui vi ) ,

(13)

where Fo is a force coefficient, vi = (1, 0, 0) is the desired velocity field at the entrance and exit, and T (x) is
a cosine taper that smoothly varies from one at the entrance or exit to zero inboard of the entrance or exit over
a distance Lf . The formulation for the volume fraction
is similar.

VOF method
In our VOF formulation, the free surface is reconstructed from the volume fractions using piece-wise linear polynomials and the advection algorithm is operator
split. The reconstruction is based on algorithms that are
described by Gueyffier, Li, Nadim, Scardovelli & Zaleski (1999). The surface normals are estimated using
weighted central differencing of the volume fractions.
A similar algorithm is described by Pilliod & Puckett
(1997). Work is currently underway to develop a higherorder estimate of the surface normal using a least-squares
procedure. The advection portion of the algorithm is operator split, and it is based on similar algorithms reported
in Puckett, Almgren, Bell, Marcus & Rider (1997).

Hydrostatic-pressure method
At the inflow boundary, the horizontal velocity is set
equal to the free-stream velocity and the normal pressure
gradient is zero. At all other side boundaries, the reduced pressure is zero and the velocity at the boundary
is extrapolated from interior grid cells. In our computations, we use the reduced pressure, Pr . We define
Pr = P ()g(z zo ), where zo is the static freesurface elevation. The resulting Navier-Stokes equations
4

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

in terms of Pr are
ui
1 Pr
(z zo ) g
ui
+ uj
=

t
xj
xi

xi

based on a NACA 0024 section. The numerical results


are compared to laboratory measurements that are reported in Zhang & Stern (1996). For the laboratory experiments, the chord length of the model was 1.2m long,
and the draft (1.5m) was sufficiently deep such that at the
bottom of the strut the effects of the free surface were
minimal. The Froude number based on chord length is
Fr = 0.55.

(14)

Recall that the density is expressed in terms of a step


function (see Equation 5). Substitution of the equation
for the density into the preceding equation gives

The length, draft, and depth of the computational doui


ui
1 Pr
(z zo )(1 )g H()
+ uj
=

(15) main normalized by chord length are respectively 4, 1,


t
xj
xi

xi
and 0.8. The height of the computational domain above
the mean water line normalized by chord length is 0.2.
The last term is discretized using the same second-order
The leading edge of the strut is located at x = 0 and the
technique used by Sussman (2003a) for the surfacetrailing
edge is located at x = 1. No flux boundary
tension term. The last term gives rise to a jump in the
conditions
are used on the centerplane of the strut (y =
reduced pressure of magnitude (z zo )(1 )g. By
0),
at
the
side
of the computational domain (y = 1.0),
forcing the reduced pressure to be zero at the walls, over
the
bottom
of
the computational domain (z = 0.8),
time, the water level at the walls relaxes to z = zo .
and the top of the domain (z = 0.2). Periodic boundary
conditions are used along the x-axis at x = 0.3725 and
I NITIAL T RANSIENTS
x = 3.6275. The three-dimensional numerical simulations used 512 128 128 = 8, 388, 608 grid points
Since VOF simulations are time accurate, there can
resulting in a grid spacing along each coordinate axis of
be problems with starting transients. As shown by Wexi = 0.0078125. The time step is t = 0.00125, and
hausen (1964) and others, unsteady oscillations can oc3001 time steps have been simulated, which corresponds
cur in the wave resistance, and by implication the surto 3.75 chord lengths. The number of sub domains along
face elevations, due to starting transients. There are also
the x, y, and zaxes are respectively 32, 8, and 8.
starting transients in the buildup of separation and the
512 CRAY T3E processors have been used to perform
boundary layer on the hull, but the viscous time constants
the numerical simulations. Each time step took approxiare significantly shorter than the wave resistance. The
mately 60 seconds per time step.
oscillations in the wave resistance occur at a frequency
Figure 1 compares numerical predictions to experequivalent to Uo /g = 1/4 and decay inversely proporimental
measurements. The numerical predictions are
tional to time. The decay rate is very slow and can lead
shown on the left side of the strut, and the experimental
to solutions that oscillate for relatively long times. This
measurements are shown on the right side of the strut.
can problematic if one is trying to reach steady state and
The color contours indicate the free-surface elevation.
also wants to minimize computer time. For computaRed denotes a wave crest ( = +0.15) and blue denotes
tions presented in this paper, a step function start of the
a wave trough ( = 0.15). In general, the agreement
velocity instantaneously jumping to the free-stream vebetween the numerical simulations and the experimenlocity has always been used. Step function starts are easy
tal measurements is very good. However, there are some
to initiate in the compute code, but they cause relatively
notable differences. For example, the numerical simulalarge transient oscillations. These very strong initial trantions show more fine-scale detail than the experimental
sients tend to weaken after the body has moved 10 body
measurements. This is because the experimental mealengths, but the weaker oscillations as predicted by Wesurements are time-averaged and the numerical simulahausen (1964) are still present. The effects of these trantions show an instantaneous snapshot of the free surface
sients are reduced by time averaging. We note that the
at t = 3.75. We also note that unlike the numerical simoscillations due to the starting transient can be mitigated
ulations, the measuring device that had been used in the
by reducing the severity of the startup from a step funcexperiments is only capable of measuring single-valued
tion to one that is much smoother and slower, which is
free-surface elevations. Another difference between nuan option that is currently being investigated.
merical simulations and experimental measurements occurs away from the strut where the numerical simulations
Results
show edge effects due to the smaller domain size that is
used relative to the actual experiments. Figure 1 illustrates that we are able to model the macro-scale features
NACA 0024 geometry
of the flow associated with the body interacting with the
The NFA code is used to simulate the flow around a
free surface.
surface-piercing vertical strut moving with constant forFigures 2 and 3 show details of the numerical simuward speed. The water plane sections of the strut are
5
Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Figure 1: Comparisons to Measurements for Strut Geometry.

lations and experimental measurements for two different


views. The numerical results are shown on the left side of
the strut and the experimental measurements are shown
on the right side of the strut. The red and green dots
along the sides of the strut denote experimental measurements of the free-surface profile. As before, the color
contours indicate the free-surface elevation. Red denotes
a wave crest ( = +0.15), and blue denotes a wave
trough ( = 0.15). Toward the rear of the foil, the
dots indicate the upper and lower bounds of the unsteady
rise and fall of the free surface due to flow separation.
Note that in Figure 2 and to lesser degree Figure 3, the
experimental contours off of the body do not appear to
agree with the experimental profiles on the body. This is
because the contour measurements off of the body could
not be performed too close to the body due to limitations
associated with the measuring device. In Figure 2, the
numerical simulations show the formation of a spilling
breaker and spray near the leading edge of the strut. Toward the rear of the strut, flow separation is evident. The
numerically predicted free-surface elevations agree well
with the profile measurements in both Figures 2 and 3.
The numerical simulations in Figure 3 illustrate that air
is entrained along the sides of the strut and in the flow
separation zone in the rear. Additional numerical simulations are in progress to establish the accuracy of the
numerical simulations.

6
Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Figure 2: Side Views Looking Down on Strut.

Figure 3: Side Views Looking Up on Strut.

7
Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

5415 geometry
The length, beam, and draft are respectively 5.72m,
0.388m, and 0.248m. The speed is 6.02 knots. Details
of the hull geometry,including the sinkage and trim, are
provided at http://www50.dt.navy.mil/5415/.
The length, width, and depth of the computational
domain normalized by ship length are respectively 2, 0.5,
and 0.5. The origin of our computational domain is taken
to be the point at which the unperturbed water intersects
the bow of the ship. The x coordinate at inflow is x = 0.5
and at outflow, x = 1.5. The height of the computational domain above the mean free surface normalized by
ship length is z = 0.5. Reduced pressure boundary conditions are used along the sides (y = 0.25) and back of
the computational domain (x = 1.5). The free-stream
velocity is imposed at the leading edge of the computational domain (x = 0.5) with zero pressure gradient.
No flux conditions are used at the top and the bottom of
the domain (z = 0.5). The CLSVOF formulation is
used to capture the free-surface interface. AMR is used
locally near the ship hull and the free surface.

Figure 4: AMR grid for 5415 at x = 0.1.

Three grid resolutions are considered: low, medium,


and high. The low resolution simulation consists of a
uniform mesh broken up into 64 rectangular grid blocks,
accounting for 2, 097, 152 cells. The mesh spacing is
= 0.0078125. The cpu time per time step is 376 seconds for the low resolution case. The low resolution simulation was run on 32 processors on an IBM supercomputer (AIX operating system).
The medium resolution simulation has 64 grid
blocks on the coarsest level and 148 grid blocks on the
finest level. There are 5, 324, 800 cells for the medium
resolution simulation. The cpu time per time step (32
processors) is 1300 seconds for the medium resolution
case.

Figure 5: Bow view of 5415

dictions and measurements is very good. However, the


experiments have fine-scale structure that is not present
in the numerics. Current research is focusing on improving resolution at the bow by using more levels of AMR.

The high resolution simulation has 64 grid blocks on


the coarsest level, 116 grid blocks on the medium level,
and 475 grid blocks on the finest level accounting for a
total of 17, 940, 480 cells. The grid spacing on the finest
level is = 0.001953125. The cpu time per time step
(64 processors) is 3000 seconds for the finest resolution
case.
Figure 4 illustrates the adaptive grid at x = 0.1
and t = 1.76. Blocks of grid points are clustered near
the ship hull and the free surface. We note that blocks of
grid points are added and deleted over the course of the
simulation depending on resolution requirements. In analyzing cells advanced per processor, the speed-up due to
adding processors or levels of adaptivity comes to about
70%. Figure 5 shows a perspective view of the bow.
The wave overturning that occurs at this Froude number
is clearly visible. Figure 6 compares numerical predictions to whisker-probe measurements at various positions
along the x-axis. In general, the agreement between pre8

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

(a)

(b)

(c)

(d)

(e)

(f)

Figure 6: CLSVOF predictions compared to whisker-probe measurements for the 5415. (a) x = 0.044. (b) x = 0.062. (c)
x = 0.080. (d) x = 0.098. (e) x = 0.115. (f) x = 0.133.

9
Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

resolution
coarse
medium
fine

grid points
320 128 96 = 3,932,160
480 192 144 = 13,271,040
640 256 192 = 31,457,280

resolution
coarse
medium
fine

grid spacing
0.00625
0.004166
0.003125

cpu time per time step (sec)


30.9
47.5
94.8

Table 1 Grid resolution.


Table 3 CPU speed.
resolution
coarse
medium
fine

sub-domains
20 8 6 = 960
30 12 9 = 3240
40 16 12 = 7680

processors
120
270
320

The mean surface elevations for each grid resolution


are shown in Figure 7. The flow is from right to left. The
color contours indicate the free-surface elevation. Red
denotes a wave crest ( = +0.025), and blue denotes a
wave trough ( = 0.025). The mean position of the
free surface is calculated from the volume fraction averaged over time from t = 4 to t = 6. Based on this time
average, the mean position of the free surface is defined
as the 0.5 isosurface. As grid resolution increases, the
bow wave becomes steeper. In addition, the trough in the
flow separation region at the corner of the wedge gets
deeper. The wave rays also become more distinct.

Table 2 Details of the domain decomposition.

Wedge geometry
The length and draft of the wedge geometry are respectively 35 and 3.5 feet. The entrance angle is 20 degrees. The speed is 6.0 knots. The Froude number is
Fr = 0.3021. The wedge geometry has a full bow based
on the Revelle hull form, and a narrow stern, based on
the bow of the Athena hull form. This enabled the model
to be towed in two different directions to investigate the
effects of fullness on the bow wave. Details of wedge geometry and the towing experiments are provided in Karion et al. (2003).

The correlation coefficient between the coarse and


medium resolution simulations is 0.94, and the correlation coefficient between the medium and fine resolution
simulations is 0.99. The rms differences between the
coarse and medium resolution simulations is 1.83103 ,
and the rms differences between the medium and fine
resolution simulations is 6.68 104 . This demonstrates that the prediction of mean quantities is converging. Increasing grid resolution also improves resolution
of small-scale fluctuations as shown in Figure 8.

The length, width, and depth of the computational


domain normalized by ship length are respectively 2, 0.8,
and 0.5. The height of the computational domain above
the mean free surface normalized by ship length is 0.1.
The bow is located at x = 0 and the stern is located at
x = 1. A reflection boundary condition is used on the
centerplane (y = 0) of the wedge. No-flux conditions
are used on the top (z = 0.1), bottom (z = 0.5), and
side (y = 0.8) of the computational domain. The flow at
the entrance (x = 0.6) and exit (x = 1.4) are forced
to be a parallel flows ((u, v, w) = (1, 0, 0)) with zero
free-surface elevations.

The rms surface fluctuations for each grid resolution are shown in Figure 8. The color contours indicate
the magnitude of the free-surface fluctuations. Red denotes the maximum rms fluctuations (
= 0.011), and
blue indicates regions where there are no fluctuations.
The rms fluctuations are calculated by taking the square
root of the variance of the vertical offset where the phase
changes from air to water. The regions where phase
changes occur include droplets of fluid above the mean
position of the free surface and bubbles of air beneath
the mean position of the free surface. The statistics are
calculated from t = 4 to t = 6. A histogram analysis indicates that phase changes are dominated by smallscale fluctuations on the mean position of the free surface. This corresponds to roughening of the free surface.
The greatest fluctuations in the free-surface elevation occur along the centerline of the wedge, in the flow separation region behind the corner of the wedge, and along
the front face of the bow wave. Comparisons of coarse,
medium, and fine resolutions show that fluctuations increase as the grid resolution increases. Interestingly, the
finest resolution simulation shows that the rms fluctuations increase in extent slightly off of the center plane on
the front face of the bow wave. Based on photographs
of the experiments (Karion et al. 2003), this is a region

Three different grid resolutions are used, corresponding to coarse, medium, and fine grid resolutions.
The details with respect to grid resolution are provided in
Table 1. The finest resolution is twice that of the coarsest.
The finest grid resolution is 0.003125 ship lengths. This
would correspond to 31cm for a 100m ship. In order to
resolve large-scale features associated with spray formation and air entrainment, we believe that grid resolutions
less than 10cm are required. Details of the domain decomposition are provided in Table 2, and the cpu time
per time step for each grid resolution are provided in Table 3. Based on these two tables, it can be shown that
the cpu time scales linearly with respect to the number of
grid points and the number of processors. The numerical
simulations have been run for 3001 time steps. The time
step for each simulation is t = 0.002.
10

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Figure 7: Mean surface elevation. (a) Coarse. (b) Medium (c)


Fine.

Figure 8: RMS surface fluctuations. (a) Coarse. (b) Medium


(c) Fine.

where the bow wave overturns. This effect is also evident


in the measurements shown in Figure 9.

color contours indicate the free-surface elevation. Red


denotes a wave crest ( = +0.03), and blue denotes a
wave trough ( = 0.03). QViz measurements are plotted on the left side of each figure. Snap shots of the free
surface at time t = 6 for each grid resolution are plotted on the right side of each figure. As grid resolution
increases, the fragmentation of the free surface also increases. We conjecture that the large-scale break up of
the free surface is dominated by inertial effects and that
the effects of surface tension are only important at the
very smallest scales. Work is currently in progress to test
this assertion.

The mean free-surface elevation for each grid resolution compared to laboratory measurements are shown
in Figure 9. Numerical predictions are plotted in the top
portion of each graph. Quantitative Visualization (QViz)
measurements are plotted in the bottom portion of each
graph. QViz uses a laser sheet to illuminate the free
surface. A video camera is used to capture snapshots,
which are then digitally processed. Additional details
of the QViz measurements are provided in Karion et al.
(2003). The color contours indicate the free-surface elevation. Red denotes a wave crest ( = +0.035), and
blue denotes a wave trough ( = 0.035). As before,
the mean position of the free surface is calculated from
the volume fraction averaged over time from t = 4
to t = 6. For these figures, 0.687 x 0.284
and 0.21 y 0.21. The correlation coefficients
between the measurements and the predictions for the
coarse, medium, and fine simulations are respectively
0.951, 0.954, and 0.957. Since the QViz instrument measures from the top down, we also consider the correlation
between the experimental data and the predictions of the
mean plus the rms fluctuations. In this case, the correlations improve to 0.950, 0.958, and 0.960 for respectively
the coarse, medium, and fine simulations.

Conclusions
With sufficient resolution, interface capturing methods
are capable of modelling the formation of spray and
the entrainment of air. Based on comparisons to other
VOF formulations that are not reported here, secondorder-accurate formulations such as those used in the
NFA and CLSVOF codes are desirable because firstorder schemes tend to inhibit wave breaking. A major benefit of our cartesian-grid formulations relative to
body-fitted formulations is that second-order VOF formulations are easier to develop.

Perspective views of the free-surface deformation


for each grid resolution are shown in Figure 10. The

In terms of future research, an AMR capability is


currently being developed for the NFA code. For our
11

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

second-order VOF formulation, a key issue is mass conservation and surface reconstruction along boundaries
where grid resolution changes. Various methods are
also being investigated to reduce initial transients. One
method slowly ramps up the free-stream velocity, which
is similar to how a towing-tank carriage operates. We are
also continuing development of techniques for processing VOF datasets to improve understanding and modelling of wave breaking.
Acknowledgements
This research is supported by ONR under contract numbers N00014-04-C-0097 and N00014-02-C-0432. Dr.
Patrick Purtell is the program manager. The second author is supported in part by the NSF Division of Mathematical Sciences under award number DMS 0108672
with Thomas Fogwell as program manager and by ONR
under contract number N00014-02-C-0543 with Judah
Goldwasser as program manager. The numerical simulations have been performed on the Cray T3E at the U.S.
Army Engineering Research and Development Center.
Figure 9: Comparisons to QViz Measurements for Wedge Geometry. (a) Coarse. (b) Medium (c) Fine.

References
Colella, P., Graves, D., Modiano, D., Puckett, E., & Sussman,
M., An embedded boundary / volume of fluid method
for free-surface flows in irregular geometries,
Proceedings of FEDSM99, 3rd ASME/JSME Joint
Fluids Engineering Conference, 1999.
Dommermuth, D., Innis, G., Luth, T., Novikov, E., Schlageter,
E., & Talcott, J., Numerical simulation of bow waves,
Proceedings of the 22nd Symposium on Naval
Hydrodynamics, Washington, D.C., 1998, pp. 508521.
Goldstein, D., Handler, R., & Sirovich, L., Modeling a no-slip
boundary with an external force field, J. Comp. Phys.,
Vol. 105, 1993, pp. 354366.
Gueyffier, D., Li, J., Nadim, A., Scardovelli, R., & Zaleski, S.,
Volume-of-fluid interface tracking with smoothed
surface stress methods for three-dimensional flows, J.
Comp. Phys., Vol. 152, 1999, pp. 423456.
Iafrati, A., Olivieri, A., Pistani, F., & Campana, E., Numerical
and experimental study of the wave breaking generated
by a submerged hydrofoil, Proceedings of the 23rd
Symposium on Naval Hydrodynamics, Nantes, France,
2001, pp. 746761.

Figure 10: Perspective views of wedge. (a) Coarse. (b)


Medium (c) Fine.

Karion, A., Waniewski-Sur, T., Fu, T., Furey, D., Rice, J., &
Walker, D., Experimental study of the bow wave of a
large towed wedge, Proceedings of the 8th International
Conference on Numerical Ship Hydrodynamics, Busan,
Korea, 2003.

12
Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Leonard, B., Bounded higher-order upwind multidimensional


finite-volume convection-diffusion algorithms,
W. Minkowycz & E. Sparrow, eds., Advances in
Numerical Heat Transfer, Taylor and Francis,
Washington, D.C., 1997, pp. 157.
MacNeice, P., Olson, K., Mobarry, C., deFainchtein, R., &
Packer, C., Paramesh : A parallel adaptive mesh
refinement community toolkit, Computer Physics
Communications, Vol. 126, 2000, pp. 330354.
Pilliod, J. & Puckett, E., Second-Order Accurate
Volume-of-Fluid Algorithms for Tracking Material
Interfaces, Technical Report LBNL40744, Lawrence
Berkeley National Laboratory, 1997.
Puckett, E., Almgren, A., Bell, J., Marcus, D., & Rider, W., A
second-order projection method for tracking fluid
interfaces in variable density incompressible flows, J.
Comp. Physics, Vol. 130, 1997, pp. 269282.
Rendleman, C., Beckner, V., Lijewski, M., Crutchfield, W., &
Bell, J., Parallelization of structured, hierarchical
adaptive mesh refinement algorithms, Computing and
Visualization in Science, Vol. 3, 2000, pp. 147157.
Rider, W., Kothe, D., Mosso, S., Cerutti, J., & Hochstein, J.,
Accurate solution algorithms for incompressible
multiphase flows, AIAA paper 950699.
Sussman, M., A second order coupled level set and
volume-of-fluid method for computing growth and
collapse of vapor bubbles, J. Comp. Phys., Vol. 187,
2003a, pp. 110136.
Sussman, M., A parallelized, adaptive algorithm for
multiphaseflows in general geometries, J. Computers
and Structures, submitted.
Sussman, M. & Dommermuth, D., The numerical simulation
of ship waves using cartesian-grid methods,
Proceedings of the 23rd Symposium on Naval Ship
Hydrodynamics, Nantes, France, 2001, pp. 762779.
Sussman, M. & Puckett, E., A coupled level set and volume of
fluid method for computing 3d and axisymmetric
incompressible two-phase flows, J. Comp. Phys., Vol.
162, 2000, pp. 301337.
Wehausen, J., Effect of the initial acceleration upon the wave
resistance of ship models, J. Ship Research, Vol. 7(3),
1964, pp. 3850.
Zhang, Z. & Stern, F., Free-surface wave-induced separation,
ASME J. Fluids Eng., Vol. 118, 1996, pp. 546554.

13
Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Joseph Gorski
Naval Surface Warfare Center, Carderock Division,
USA
The Cartesian based grid strategy the
authors present looks very attractive both from the
ability to quickly generate grids as well as the
improved accuracy possible for computing the free
surface.
Since no additional gridding beyond
potential flow methods is required have the authors
investigated how well the approach works for
predicting the boundary layer flow near a ship, which
can influence the wave field at the stern?
Also,
have the authors made a direct comparison of the two
methods, VOF and CLSVOF, or have any indication
that one may be better than the other?

AUTHORS REPLY
Thank you Joe for your comments.
As
formulated, our Cartesian grid method will not
resolve boundary layers along the hull. At the stem,
where turbulent break up is apt to occur, we would
have to incorporate a model into our formulation.
However, at the stern where there is a clean
separation, our formulation is adequate.
We are
currently in the process of comparing the merits of
VOF and CLSVOF.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Alessandro Iafrati
Istituto Nazionale per Studi ed Esperienze di
Architettura Navale, Italy
First of all, the authors should be
congratulated on producing such large amount of
results. I have two questions. The first one concerns
the entrance and exit boundary conditions: the model
you adopted to enforce a parallel uniform flow with
zero free surface elevation at entrance and exit looks
rather efficient. I wonder if you could give more
details about it. My second question concerns the
comparison between numerical and experimental
results shown on figures 1 and 2 of your paper: timeaveraged experimental measurements display
longitudinal striations which are not so evident from
the numerical results. Maybe, this is because the
latter are just instantaneous snapshot. Did you try to
establish the same comparison by using timeaveraged free surface profiles also for the numerical
data?
AUTHORS REPLY
Thank you Alessandro for your comments.
We force a uniform stream at the entrance and exit.
The x-component of velocity is set equal to negative
one, and the y- and z-components of velocity are set
equal to zero. The free-surface elevation is set equal
to its still-water level. The nature of the forcing is
very similar to the forcing that you have formulated
in your own research studies as noted in our paper.
In regard to our comparisons with experiments, we
have discussed this matter with experimentalists who
indicate that the capacitance wave probes that were
used in the experiments are prone to drift. This may
explain the striations that are observed in the
experiments.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
1.

The Qviz measuring technique is currently being


generalized so that it is possible to investigate
unsteady effects.

2.

Research is ongoing to investigate flows at lower


Froude numbers for a range of hull types.

3.

The CLSVOF formulation uses a free-slip


condition.
The VOF formulation as
implemented in NFA uses either a free-slip or
no-slip condition. The NFA results reported in
this paper use no slip. Your comments in regard
to implementing a mixed condition are
interesting and deserve further study.

4.

Blockage is always an issue in finite-size


computation domains. Matching to a spectral
solution in the outer domain is one way to
alleviate this problem. Alternatively, adaptive
mesh refinement is also attractive.

5.

We are primarily interested in time-accurate


computations.
As a result, our numerical
algorithm is optimized for explicit time-stepping
procedures. We note that VOF has stringent
Courant conditions that are hard to circumvent.
For our purposes, implicit time-stepping
procedures would smear the free surface too
much.

Kelli Hendrickson
Massachusetts Institute of Technology, USA
Firstly, I would like to thank the authors for
their work. All of them have contributed greatly to
the field of numerical Naval Hydrodynamics and this
work is an example of their continuing effort to delve
into the complex nature of even simulating high
Froude number flows about surface ships, much less
improve our understanding of the phenomena of
wave breaking, spray formation and air entrainment.
I have questions regarding both the
numerical method and comparisons with
experiments, so please bear with me if they appear
too detailed for this venue.
1.

2.

3.

4.

5.

Many of the comparisons to experiments are


quite good in terms of mean values such as the
regions where QVIZ data is available. Is there
available data to attempt comparison for the
unsteady components?
All of the results shown here are for fairly high
Froude number cases where experimental data is
difficult to obtain near the hull. Has there been
any validation effort for low Froude number
cases where experimental data (average and rms)
is easier to obtain?
It appears that the body boundary condition in
use for all of the results is a free-slip condition,
even though a no-slip condition is discussed in
the paper. Is the body force technique capable of
implementing a mixed condition which utilizes a
contact line model such that the effect of the
free-surface and wave breaking have on the body
boundary layer can be studied?
Do you feel the side-wall free-slip boundary is
affecting your results and if so, how difficult is it
possible to implement a condition which has less
blockage effect?
Because the method uses explicit time stepping,
the time step is relatively small. While the code
is fairly quick for the problem being attempted, I
imagine that a simulation to full steady state is
quite expensive in terms of computational time.
As the authors conjecture that the wave breakup
is more inertial dominated than surface tension
dominated, what would be the drawbacks to
using an implicit time integration algorithm to
increase the time step size?

AUTHORS REPLY
Thank you Kelli for your comments. We
address your questions point by point.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Stephen Scorpio
Johns Hopkins University, USA
I would like to thank the authors for an
excellent paper. What are your ideas for validating
the fine scale structure (e.g. ejection of water
droplets, air entrainment and free surface
fluctuations) predicted by your model?
AUTHORS REPLY
Thank you Steve for your comments. A
range of instruments are currently being developed to
capture the fine-scale structure that is present in ship
flows. In addition, different methods for processing
existing datasets are also being developed to capture
unsteady effects.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics


St. Johns, Newfoundland and Labrador, CANADA, 8-13 August 2004

Experimental Study of the Bow Wave of the R/V Athena I


Thomas C. Fu, Anna Karion, James R. Rice, and Don C. Walker
(Naval Surface Warfare Center, Carderock Division, USA)
ABSTRACT
Detailed measurements of the bow wave of the R/V
Athena I were made, at night, over the three-day
period 29-31 October 2003. The data was taken while
the ship traversed back and forth in St Andrews Bay,
Panama City, FL at ship speeds of 3.1, 4.6, 5.4, and
6.2 m/s corresponding to Froude numbers based on
waterline length (47 m) of 0.14, 0.21, 0.25, and 0.29,
respectively. Measurements were made using a laser
based optical technique. This data will be used to
characterize the breaking bow wave for a real ship in
real conditions, as well as to compare with tow tank
data of a large bow wedge (Karion et al (2003)). The
bow wedge is of similar geometry and was towed at
3.1 and 4.6 m/s. two of the same speeds as the R/V
Athena I. Utilizing these two data sets the differences
between field and tow tank testing will be
investigated, including scale effects and the effect of
ambient seas.
INTRODUCTION
While detailed measurements of bow waves have
been done previously (e.g. Ogilvie (1972), Miyata
and Inui (1984), Waniewski et al (2002), Roth et al
(1999), Karion et al (2003) and others), these test
have all been performed in tow tank facilities,
limiting the model size and wave amplitude. Except
for Karion et al (2003), these tests were performed in
relatively small tow tanks. More recently large
models (displacement = ~12,200 kg, Overall Length
=10.7m) have been utilized to study bow waves and
bow wave breaking by Karion et al (2003).
Large scale breaking has also been
examined by Coakley et al (2001) and Furey et al
(2003). In both of these works a large breaking wave
was generated from a submerged hydrofoil held at a
fixed angle of attack. Waves with peak to trough
amplitudes of ~0.5 m were studied. Though these
waves should be considered full-scale breaking
waves, they were spilling breakers more similar to
the trailing wave behind the transom of a ship (see

Figure 1) than the plunging wave associated with the


curl over of the bow wave of a flared hull ship (see
Figure 2)
Past, at sea studies of bow waves have been
mainly qualitative in nature, Ratcliffe (2003) for
example. Recently, measurements of bow wave spray
droplet size and velocity distributions have been
made on the R/V Revelle (Sur & Chevalier (2003)),
but measurements of the free-surface elevation of the
bow wave were not made. This paper will report the
results from field measurements of the bow wave of
the R/V Athena I. Quantitative measurements of the
bow wave elevation were made at several speeds
ranging from 3.1 to 6.2 m/s, in the relatively calm
water of St. Andrews Bay, Florida. There was
relatively little ship motion and the ambient seas
were characteristic of sea state 1, though there was
some slight wind chop (actual test conditions will be
discussed in the results section). Additionally,
comparisons between this data and breaking wave
data, obtained at NSWC for a bow wave generated
from a large wedge (Karion et al (2003) and Karion
et al (2004)) will be made, as well as a qualitative
comparison to a potential flow computation provided
by Donald C. Wyatt of SAIC, La Jolla, CA
TEST DESCRIPTION
The test spanned a three-day period from 29-31
October 2003. The ship left the dock at 2:00 AM,
local time, each morning and returned at 11:00 AM
the same day. Data was collected while the boat
transited back and forth between two stations located
approximately 2 nautical miles apart and oriented in a
North-Northeast direction. Ships headings were 60o
80o and 240o-280o .
Test Platform: The R/V Athena I (see Figure 3) is a
converted PG-84 class patrol boat built in 1969 and
converted to a research vessel in 1976. She has an
aluminum hull and an aluminum and fiberglass
superstructure. The ship owned by the U.S. Navy and
is operated by NSWC Carderock. Measurements and

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

a)
Figure 3: The R/V Athena I traveling at +10 m/s.
details for the R/V Athena are listed in Table 1.
More
information
can
be
found
at
www.code50.nswccd.navy.mil.

b)
Figure 1: a) Breaking wave over a submerged
hydrofoil, U=2.6m/s, chord=1.8m, depth of
submergence =1.8m, b) Flow behind the DDG-67,
USS Cole, U=10.3m/s.

a)

Instrumentation: A laser sheet quantitative


visualization method (QViz) was utilized to measure
lateral free-surface profiles at several axial locations.
The free-surface was illuminated by a laser light
sheet generated by a scanning mirror and imaged
using three black and white, progressive scan
cameras. In the current experiment, a laser light sheet
was generated from a scanning mirror and projected
from an optical rail system down to the waters
surface. The laser used was a diode-pumped, solid
state YAG laser, with an output of 2.5-3.0 W at 532
nm (Model MLM-0532 by Melles-Griot). The set-up
is shown in Figure 4. Two of the cameras faced aft,
while the third camera pointed forward. This was to
insure that if the wave blocked the aft facing
cameras view of the laser sheet, the forward facing
camera would still be able to obtain clear images.
The recorded digital images were then corrected for
distortion and calibrated. The corrected images were
then processed
Table 1: Description of the R/V Athena I

b)
Figure 2: Images of the bow wave of the DDG-67
USS Cole at 20 knots, a) bow on and b) side view.
Note the fluid sheet curl over characteristic of a
plunging breaker.

Length Overall

50.3 m

Length Waterline

47.0 m

Extreme Beam

7.3 m

Draft

3.2 m

Propulsion

Twin screw,
Twin diesel (low speed)
Gas Turbine, (high
speed)

Speed

6.5 m/s (diesel)


18 m/s (turbine)

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Camera 1

Camera 3

Spatial
Reference

Camera 2
a)

Camera 3

Spatial
Reference
Laser Sheet

Camera 2

Camera 1

b)
Figure 4: Image of the QViz set-up showing the 3
cameras and the spatial reference.

proved to be easier for the Captain than constant


speed through the water, due to the type of readout,
crude dial for speed through the water, digital display
for the GPS speed. So in order to maintain as
constant a speed as possible during a run GPS speed
was used. This issue will be resolved on future tests
by the addition to the bridge of digital knotmeter
readout. Because GPS speed was used, there is some
variability introduced by currents. Some attempt will
be made to characterize these currents by looking at
the differences between successive runs made at
opposite headings.
Test Conditions: Measurements of the bow wave
were made at four speeds, in head and following seas.
Data was collected in the relatively protected waters
of St. Andrews Bay, Florida to minimize ship motion
from ambient seas and so measurements could be
made in relatively calm conditions. Testing was
performed at night, which provided a stronger signalto-noise ratio in the QViz images. Specifically, data
was collected from 3:00 AM to 6:00 AM local time.
This time period coupled with the time of year (Oct
29-31), also made for calm test conditions. The
winds, which typically build as the day progresses,
were minimal (typical speeds 3.5 to 5.5 m/s), and it
was a new moon, so tidal excursions were
minimized. Data was collected at 3.1, 4.6, 5.4, and
6.2 m/s, corresponding to Froude numbers based on
water line length (47 m) of 0.14, 0.21, 0.25, and 0.29,
respectively. The test conditions were chosen to
cover a range of displacement speeds and also to
allow for comparison with tow tank results for a large
bow model of similar geometry to the R/V Athena
(Karion et al (2003)).
DATA PROCESSING

to provide the free-surface elevation in the image


plane of the camera. This technique has been used
extensively to measure free-surface elevations and
breaking waves (Fu et al (2003), Karion et al (2003))
and is described in Furey and Fu (2002).
The current system and its capabilities are
described in detail in Rice (2004). Figure 4b shows
the test set-up as deployed. Note the sea state. The
conditions shown in Figure 4b are representative of
the ambient conditions when data was collected. The
three cameras can be seen and also the light bar used
as a spatial reference in calibration images.
Ship motions were recorded from the ships
onboard GPS compass; a wave buoy recorded the
ambient seas; and the wind speed and direction were
measured using the ships onboard anemometer.
Maintaining constant GPS speed (speed over land)

The ship motion data from the GPS compass was


reduced to provide time-averaged heading, speed,
pitch and roll for each run. RMS values of each of
these quantities were also calculated to evaluate the
steadiness of the run. Figure 5 shows the heading,
speed, roll and pitch time series for a typical run and
the RMS values calculated for that run. It can be seen
that within in the accuracy limits (1.5 deg RMS Max.
Error) of the GPS compass (FURUNO SC-120) there
is minimal ship motion. This is important when
interpreting the results and in assessing the accuracy
of the free-surface elevation measurements. The
mounting pieces could be manually made to vibrate
and this would introduce measurement error.
Assessment and evaluation of the in-situ set-up, of
the QViz images, images from the video cameras
mounted on the set-up, showed minimal motion
relative to the ship.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

SPEED

Speed (m/s)

6.5
6.0
5.5
5.0
4.5
40.0

60.0

80.0
100.0
Time (sec)

120.0

ROLL
Roll (deg)

4.0
2.0
0.0
-2.0
-4.0
40.0

60.0

120.0

PITCH

2.0
Pitch (deg)

80.0
100.0
Time (sec)

1.0
0.0
-1.0
-2.0
40.0

60.0

80.0
100.0
Time (sec)

120.0

Heading (deg)

HEADING
360
330
300
270
240
210
180
40.0

60.0

80.0
100.0
Time (sec)

120.0

Figure 5: Typical time series of a) speed: mean=5.35


m/s, RMS =0.02, b) roll angle: mean=-0.86 degrees,
RMS=0.18, c) pitch angle: mean= -0.29, RMS=0.11,
and d) compass heading: mean=250.9 degrees,
RMS=0.58 for the R/V Athena I in St. Andrews
Bay.

Figure 6: Example of a QViz image of the R/V


Athena I bow wave at U=4.6 m/s, with the edge
detected by the data processing algorithm
superimposed in red (points) and yellow (smoothed
line).
As mentioned above, an attempt was made
to characterize the currents in the test region. By
looking at the average GPS speed and the average
speed through the water for each pair of successive
runs, the magnitude of the current in direction of boat
travel can be estimated. This analysis proved to be
less straightforward than anticipated. There proved
to be no consistent result from this analysis, i.e. each
pair of runs did not show a consistent bias in speed or
direction. This result most likely stems from
inaccuracies in the recording of the speed through the
water and in the lack of stationary statistics. Taking
into account all the data, one can characterize the
speed of the current as being approximately 0.3 m/s
and is viewed as an uncertainty in the reported speed.
QViz images were corrected for distortion
and calibrated using standard National Instruments
image processing subroutines. These corrected
images were then processed to provide free-surface
profiles. An image analysis program developed at
NSWCCD using National Instruments LabView
software with the Image Processing (Vision) toolbox
was used to extract the surface profile. Figure 6
shows a sample image with the processed profile
superimposed.
Free-surface images were recorded at 30
frames per second for one minute giving 1800
profiles. Each frame, for each speed and axial
position, was analyzed to generate time-averaged
profiles and the standard deviation for each location.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

a)

b)

c)

d)

c)

e)

Figure 7: Successive QViz images t =1/30th second, U=4.6 m/s and x=5.9 m from the bow stem.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

RESULTS
A sample of successive frames with the extracted
edge superimposed in red, for U=4.6 m/s at an axial
location 5.9 m from the bow stem are shown in
Figure 7. Note the wave crest amplitude changes by
approximately 0.15 m in ~ 1/6th of a second. To
characterize this unsteadiness of the free-surface
profiles, the mean and the standard deviation for each
speed and location were computed. Figure 8 shows
these mean profiles in red and the standard deviation
in blue for speeds of 4.6, 5.4, and 6.2 m/s at the same
axial location and the same general heading
As noted above 1800 individual profiles
were collected per run. At least two runs were
performed for each speed and axial position.
Comparing mean free-surface profiles for the same
location, for two separate runs, it was observed that
two mean profiles were very similar. The most
obvious difference was a slight amplitude offset, due
to a slight difference in speed. This difference was
much smaller than the unsteady fluctuations of the
instantaneous profile, where the difference between
runs was approximately 2 cm and the range of the
unsteady fluctuations was around 15 cm, so the large
fluctuations are not due to the way the ship is
operated, i.e. variations in speed or ship motions
while recording a data set.
Though the ambient conditions were very
calm, with only a slight wind chop from a 3.5 to 5.5
m/s breeze, there is still a great deal of fluctuation in
the free-surface elevation. Data for a given speed
was acquired at two headings, roughly equivalent to
head and following sea conditions. The effect of
heading on the breaking bow wave can be seen in
Figure 9. Figure 9 shows the average free-surface
profile for an axial position 5.3 m aft of the bow stem
for two successive runs at opposite headings. The
shape of the profile is markedly different. One can
see that on the southward run (red, heading = 255
deg.), which would be nominally a head seas
condition, the wave is smaller and the peak of the
profile is farther from the ship, than is the northbound
case (blue, heading = 65 deg.)). This difference is
not due to the direction of the swell, because there
was minimal swell present, but mostly due to the fact
that the wind direction is slightly off of the ships
track. The wind came out of nominally 30 degrees.
The QViz instrumentation was mounted on the port
side of the ship. So the system would then be on the
windward side on northbound runs and on the
leeward side during southbound runs. The slight
difference in speed through the water, due to the
currents in the bay should also be noted. This effect

was similar across the speed range, i.e. the wind


pushed the peak of the bow wave closer to the ship
on the windward side.

a) U=6.2 m/s

b) U=5.4 m/s

c) U=4.6 m/s
Figure 8: Mean surface profiles (red) and standard
deviation (blue) for a) 6.2 m/s, b) 5.4 m/s, and c) 4.6
m/s, at an axial position 5.9 m aft of the bow stem.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

standard deviation for R/V Athena data is ~0.1m and


0.01m for the tow tank. This was seen at all speeds
and is to be expected, due to the more variable
ambient conditions present in the field versus the
very calm conditions present in a tow tank. There is
also minimal ship motion and speed fluctuation in a
tow tank test, which would add to the free-surface
variability seen in the R/V Athena data, though
certainly this is not as large an effect as the ambient
conditions.

Figure 9: Effect of ships heading on the amplitude


of the bow wave for U= 4.6 m/s at x=5.3 m aft of the
bow stem.

The physical appearance of the bow wave


was very similar. This can be seen in Figure 12. The
most noticeable difference is not in the bow wave
itself, but the fact that in the field, after the bow wave
breaks, the presence of organic surfactants stabilizes
the white bubbly flow in the breaking zone and it is
carried downstream. While in the tow tank the
bubbles in the breaking region have a very short
lifetime and the surface convected bubbly layer is not

DISCUSSION
Figures 8 shows that the standard deviation in
amplitude of the profiles is only weekly dependent
upon speed. The magnitude of the variability was on
the order of 0.15 m. Because the magnitude of the
fluctuations was for a large part independent of speed
that would lead one to believe that ambient
conditions may be causing this variability. Looking
at the tail ends of the profiles away from the crest of
the bow wave the range of fluctuations is on the order
of 0.1 m. This is similar to the difference seen in
Figure 9. It may very well be that the wind chop
present was on the order of 0.1 m in amplitude and a
great deal of the variability present is due to the wind.
A more systematic assessment of the ambient seas
surface conditions has not been made to date. Video
and analysis of portions of the images further from
the ship would also help in characterizing the
ambient seas.
The objective of the test was to obtain
detailed free-surface measurements of the bow wave
of a full-scale ship. Data was taken in as calm as
possible field conditions to allow for comparison to
tow tank data.
Looking more closely at the
differences between field and tow tank conditions,
Figure 11 shows the time-averaged wave profile for
the Athena and for a large wedge model towed at the
same speed (4.6 m/s) and nominally the same axial
position. The wedge model is large (5 ft draft, 7 ft
max beam) and flared with a 20-degree entry angle,
making it generally similar to the R/V Athena. A
detailed description of the geometry is found in
Karion et al (2003). Figure 11 shows that the

a) R/V Athena, U=4.6 m/s. mean-red, std. dev. -red

b) Bow-wedge, U=4.6 m/s (Karion et al (2003)),


mean red, std. dev. - blue
Figure 11: Comparison of the standard deviation in
the free-surface profiles of the a) R/V Athena and a
b) large wedge (tow tank).

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

a) Bow Wedge, U=4.6 m/s


Figure 13: Contour map of the predicted and
measured (insert) free-surface elevation (cm)
(courtesy of D. Wyatt).

b) R/V Athena II at U=4.6 m/s


Figure 12: Images of the bow wave of a) a large bow
wedge and b) the R/V Athena II
present. The turbulent break up of the thin film on the
hull also is much more pronounced on the R/V
Athena than it is on the tow tank model. On the bow
wedge the thin film sheet does not break up until it
curls over and the bow wave breaks, but in Figure
12b the thin sheet on the Athena bow has already
begins to break up immediately.

a) Das Boot Prediction

One last comparison to be made is that of


the time-averaged free-surface measurements to the
results of a steady potential flow code. Donald C.
Wyatt using the Das Boot potential flow code
provided pre-test predictions. An example of his
results is shown in Figure 13. The potential flow
computation was for a steady, sea state 0 case, so
exact agreement is not expected, but comparing
Figure 14a (predicted) with Figure 14b (measured),
the amplitudes and shape are in general agreement.
b) Measurement
CONCLUSIONS
Data was collected to characterize the breaking bow
wave of the R/V Athena I, and to begin to investigate

Figure 14: Contour map of the a) predicted and b)


measured free-surface elevation (cm) of the R/V
Athena, U=4.6 m/s..

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

the effects of ambient conditions and the differences


between tow tank and field conditions.
These measurements demonstrated how
even relatively calm field conditions still introduce
unsteadiness and variability to the free-surface wave
field. Though they also showed there is some merit
and fidelity to steady CFD codes, in that the timeaveraged measurements seem to give good agreement
with the predictions, at least in this case where the
experiment was designed to provide data for the calm
water, minimal ship motion case.
Detailed measurements of the bow wave
were made using a laser-sheet visualization technique
demonstrating that measurements of this type, that
were in the past only possible in the laboratory, can
now be done in the field on full-scale ships.
ACKNOWLEDGEMENTS
This work would not have been possible without the
support of the Toby Ratcliffe (NSWC, Code 5200),
Media Lab (NSWC, Code 3830), Tom Broglio
(NSWC, Code 5300), the crew of the R/V Athena,
Don Wyatt (SAIC), Dr. Arthur Reed (NSWC, Code
5030) and Dr. L. Patrick Purtell (ONR). This work is
supported by ONR under contract number
N0001403WX20633. Dr. L. Patrick Purtell is the
program manager.
REFERENCES
Coakley, D. B., Haldeman, P. M., Morgan, D. G.,
Nicolas, K. R., Penndorf, D. R., Wetzel, L. B., and
Weller, C. S., Electromagnetic Scattering from
Large Steady Breaking Waves, Experiments in
Fluids, Vol. 30, 2001, pp. 479-487.
Fu, T.C., Furey, D., Karion, A., Mutnick, I., Rice, J.,
Sur, T., and Walker, D. Hydrodynamic
Measurements of a Steady Wave During Various
Breaking Conditions in the Circulating Water
Channel, Naval Surface Warfare Center, Carderock
Division, Hydromechanics Directorate R&D Report,
NSWCCD-50-TR-2003/012, March 2003, pp. 30.

of a Large Towed Wedge, 8th International


Conference on Numerical Ship Hydrodynamics,
Busan, Korea. September 22-25, 2003.
Karion, A., Fu, T.C., Rice, J., Walker, D., and Furey,
D., "Experimental Measurements of the Surface of a
Breaking Bow Wave, to be presented at the 25th
Symposium on Naval Hydrodynamics, St. Johns,
Newfoundland and Labrador, Canada, August 8-13,
2004.
Miyata, H. and Inui, T., Non-linear Ship Waves,
Advances in Applied Mechanics, Vol. 24, No. 1,
1984, pp. 215-288.
Ogilvie, T. F., "The Wave Generated by a Fine Ship
Bow", 9th Symposium on Naval Hydrodynamics,
1972, pp. 1483-1524.
Ratcliffe, T., "Photographic Visualization of the
Surface Wave Field and Boundary Layer
Surrounding the Research Vessel, R/V Revelle",
Naval Surface Warfare Center, Carderock Division,
Hydromechanics
Directorate
R&D
Report,
NSWCCD-50-TR-2003/003, March 2003, pp15.
Rice, J., Fu, T.C., Karion, A., Walker, D., and
Ratcliffe, T., "Quantitative Characterization of the
Free-Surface Around Surface Ships, to be presented
at the 25th Symposium on Naval Hydrodynamics, St.
Johns, Newfoundland and Labrador, Canada, August
8-13, 2004.
Roth, G.I., Mascenik, D.T., and Katz, J.,
Measurements of the Flow Structure and Turbulence
within a Ship Bow Wave, Physics of Fluids, Vol.
11, No. 11, 1999, pp. 3512-3523
Sur, T. W. and Chevalier, K., Field Measurements
of Bow Spray Droplets, ONR Final Report, ONR
Contract Number N00014-03-C-0105, September 30,
2003, pp31.
Waniewski, T. A., Brennen, C. E., and Raichlen, F.,
"Bow Wave Dynamics", Journal of Ship Research,
Vol. 46, No. 1, March 2002, pp. 1-15.

Furey, D.A. and Fu, T.C., Quantitative Visualization


(QViz) Hydrodynamic Measurement Technique of
Multiphase Unsteady Surfaces, 24th Symposium on
Naval Hydrodynamics, Fukuoka, Japan, July 8-13,
2002.
Karion, A., Sur, T., Fu, T.C., Furey, D., Rice, J., and
Walker, D., "Experimental Study of the Bow Wave

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION

at-sea data, but have not had a chance to remove any


effects of ship motion or ambient conditions.

Dane M. Hendrix
Naval Surface Warfare Center, Carderock Division,
USA
The difficulty in obtaining these detailed
measurements in the field for conditions that are of
practical interest and over extents likely to be helpful
to modellers is tremendous. I would like to thank the
authors for a very interesting paper and encourage
them to continue to examine the details of this data to
improve our understanding of the factors that control
the characteristics of breaking wave flows.
One question I have concerns the rather
dramatic difference due to ships heading shown in
Figure 9. I note that the ship speed and the wind
speed are roughly equal and that your heading is
nearly down wind. Do you think it likely that the
down wind pass is a condition where the ship is
moving in phase with the wind generated waves?
There might be a steady wave pattern around the ship
that is dependent on this particular condition and may
account for the difference in the average wave profile
observed.
AUTHORS REPLY
This is certainly possible and something we
will be looking at as we get further into the analysis.
The wind generated waves were minimal due to the
small fetch in the bay, but nonetheless, there was an
observable wind generated chop.
DISCUSSION
Dane M. Hendrix
Naval Surface Warfare Center, Carderock Division,
USA
My
second
question
concerns
a
characteristic of the flow that you may not have yet
had time to analyze in detail. In Figure 11, you show
that the average profile is much more variable for the
at-sea test than for the tow tank large bow wedge.
How would you compare the surface roughness of
the bow wave observed on Athena and on the large
bow wedge at the same speed at an instant in time?
AUTHORS REPLY
We have begun to look more in depth at the
rms fluctuations of the free-surface elevations
between the at-sea and tow tank data. We are
planning to compare the roughness spectra from the
two tests. We do observe increased variability in the

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Stephen M. Scorpio
Johns Hopkins University, USA
Does QViz require the presence of droplets
or bubbles to visualize the free surface?
AUTHORS REPLY
QViz does not require droplets or bubbles,
but it does require sufficient surface roughness to
spectrally scatter enough light that the surface can be
detected from the recorded images. This has not
been an issue in the field, but it does require the QViz
images be recorded at night. In the lab, we use
fluorescing dye to mark the free-surface when there
is insufficient scattering.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Patrick Purtell
Office of Naval Research, USA
What is the effect of surface roughness since
the fluctuations occur both in the tank and at sea?

AUTHORS REPLY
The surface roughness observed in the tank
is associated with the turbulence from the breaking
bow wave and some residual disturbance from the
previous run. The surface roughness seen in the field
is due to both the breaking wave turbulence and with
ambient roughness from wind driven chop and
ambient waves. We have made full-scale laboratory
measurements of the small scale roughness generated
by wind and see a significant increase in spectral
content at the wavenumbers associated with surface
roughness. The surface roughness does not seem to
alter the time averaged wave fields, but we will
continue our analysis and attempt to quantify or at
least identify the affects of the ambient conditions on
bow wave breaking.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Richard Lahey, Jr.
Rensselaer Polytechnic Institute, USA
Is it not possible to also measure the
velocity, turbulence and void fraction fields using
existing instrumentation (e.g., ). These type of salt
water data are badly needed by modelers.
AUTHORS REPLY
It is certainly possible to measure velocity,
turbulence and void fraction with existing
instrumentation at least at discreet points. Velocity
and turbulence can be measured with Acoustic
Doppler Velocimetry (ADV) probes in the lower void
fraction regions. And Eric Terrill at the Scripps
Institution on Oceanography, UCSD, did make void
fraction measurements of the transom flow region
during the May 2004, Athena effort, and Mory
Gharib, of the California Institute of Technology also
measured bubble size and velocity distributions. So
while it is certainly possible to measure velocity,
turbulence, and void fraction to go along with the
free-surface elevation and surface roughness, we
have not yet attempted to obtain a detailed
comprehensive salt water data set of all these
quantities for the same breaking wave feature. That
canonical data set is indeed one of the goals of the
field work and the authors do thank the modelers for
their encouragement and support.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Leonard Imas
Stevens Institute of Technology, USA
This paper describes a study in which fullscale wave breaking around a surface ship was
investigated. What makes this work unique is that
very little full-scale data on this phenomenon is
available in published literature. Congratulations
should be extended to the authors for undertaking an
experiment of this magnitude, especially one where
quantitative
measurements
were
obtained
successfully. A few comments on the discussion
presented in the paper:
1) On page 1, the authors present the different
speeds at which Athena was run vs. those at
which the wedge experiment was run. While
some of these are the same, in terms of scaling
they represent different Froude, Reynolds, and
Weber number combinations. It would have been
more useful to make these comparison using
these nondimensional parameters (using draft as
the length scale) rather than absolute speeds.
AUTHORS REPLY
The intent was to focus on the field
measurements and only make initial observational
comments on the comparison of the data to the wedge
experiment data. We will indeed carry out the kind of
specific analysis Len mentions. In fact, the draft at
the bow of the R/V Athena I is only 1.55m which is
very close to one of the three wedge drafts (1.5m,
1.1m, and 0.6m), so direct comparison and scaled
comparisons can be made.
DISCUSSION
Leonard Imas
Stevens Institute of Technology, USA
2) On page 4, there is a discussion regarding
characterization of near surface currents. It
should be noted that this is not a trivial task due
to the unsteady nature of Gulf of Mexico surface
currents near the free-surface. Much work has
been carried out in investigating this subject by
the off-shore petroleum industry in the framework of VIV problems.
AUTHORS REPLY
This has indeed been more complicated than
first anticipated. The data was taken in St. Andrews

Bay, with successive north-south runs spaced by


roughly 30 minutes. It was hoped that a bias would
be found between the GPS speed (speed over land)
and the EM Log (speed through the water), which
could be used as an estimate of the current. The
surface currents have indeed been found to be very
unsteady.
DISCUSSION
Leonard Imas
Stevens Institute of Technology, USA
3) Breaking wave time-space histories like those
shown in Figure 7 on page 5 could be useful in
validating the time-accuracy and free-surface
turbulence modeling fidelity of VOF and LS
uRANS and LES codes currently employed in
breaking wave studies.
4) On page 6, the authors highlight the (significant)
effect of wind directionality relative to ship
speed and heading has on breaking wave
kinematics. This is important to note as only a
very limited number of numerical studies on
wave breaking have accounted for wind effects.
The experimental results presented, to some
extent, bring into question the overall validity of
many numerical results.
5) On pgs. 7-8, there is a discussion regarding (i)
the bubbly region aft of the bow wave breaking
zone and (ii) the break-up of the bow (wave runup) thin film, namely the difference between the
physics observed in the Athena experiment and
the wedge experiment. While organic surfactants
will play a role in elucidating the differences in
(i) between the two experiments, the fact that the
Athena measurements/observations were taken
in
salt
water
while
the
wedge
measurements/observations were taken in fresh
water will have an impact of equal magnitude on
bubble transport in the breaking wave zone. The
differences in item (ii) are likely due to the
different Fr, Re, and We no.s combinations
representing each experiment. The thin sheet
break-up in this problem is closely related to the
dynamics and kinematics of the contact line
problem, which itself is highly dependent on Fr,
Re, and We no. (Imas 1998).
6) On page 8, the comparison to the potential flow
(Das Boot) panel method result should be treated
as qualitative since the tool employed in the
computation is not capable of explicitly

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

predicting wave breaking zones around a


surface-piercing body.
AUTHORS REPLY
This comparison was only intended to be
qualitative in nature. The point was to validate that
QViz was tracking the free-surface accurately, since
this was the first time the system has been used in the
field.
DISCUSSION
Leonard Imas
Stevens Institute of Technology, USA
7) On page 8, while I agree with the authors
conclusion that there is some merit to steady
CFD computations in the context of free-surface
wave predictions, the real issue here is what
application is the steady vs. unsteady CFD tool
being used for and whether the tool is
appropriate in that specific context. To answer
such questions, data from experiments such as
the one presented in this paper (e.g. Figures 7, 8,
11) will prove to be very useful.
REFERENCES
Imas, Leonard, 1998 Hydrodynamics Involving a
Free-Surface Body Juncture, Ph.D. Thesis, MIT,
Department of Ocean Engineering.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Tricia Waniewski Sur
Science Applications International Corporation, USA

DISCUSSION

I would like to ask the following questions.

Tricia Waniewski Sur


Science Applications International Corporation, USA

1.

3.

It is difficult to compare the free-surface


measured by QViz presented in Figure 14 with
that calculated by Das Boot presented in Figure
13. Is it possible to present these results similar
to Figure 12 in Karion et al. (2003) to facilitate
comparison?

Has any of the QViz data taken on the R/V


Athena May 2004 field experiment been
processed? Are there differences in free surface
elevations for the same operating conditions? Is
a quantitative assessment of wave breaking
parameters (similar to Karion et al., 2003)
planned?

AUTHORS REPLY
AUTHORS REPLY
This has been done and is in the final
version of the paper.
DISCUSSION
Tricia Waniewski Sur
Science Applications International Corporation, USA
2.

The discussion section describes differences in


the physical appearance of the bow waves in the
towing tank and in the field. It states, The most
noticeable difference is not in the bow wave
itself, but the fact that in the field, after the bow
wave breaks, the presence of organic surfactants
stabilizes the white bubbly flow in the breaking
zone and it is carried downstream. It is wellknown that more, smaller bubbles are created by
breaking waves in seawater. In addition, smaller
bubbles tend to be more stable than larger
bubbles. Is it possible that differences in bubble
number density and size distribution are the
primary reason for the difference in physical
appearance?

A sampling of the data has been analyzed.


The data for similar conditions for the two cruises
looks very similar. There are differences in freesurface elevation for similar operating conditions, but
this is mainly due to environmental conditions and
unsteadiness. As we get further into the analysis, we
will be able to examine this more. A quantitative
assessment of breaking parameters is planned for the
coming year.
REFERENCES
Karion, A., Sur, T. Waniewski, Fu, T.C., Furey,
D.A., Rice, J.R., and Walker, D.C., Experimental
Study of the Bow Wave of a Large Towed Wedge,
Proc. 8th International Conference on Numerical Ship
Hydrodynamics, 2003.

AUTHORS REPLY
The primary difference in the appearance is
the stability of the white bubbly region in saltwater.
This is due to differences in surface tension. The
smaller bubble size certainly contributes to the
bubble stability, but the visual differences I am
referring to are not differences in the appearance of
the white water regions, but the fact there are larger
areas of whitewater due to the differences in surface
tension.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics


St.Johns, Newfoundland and Labrador, CANADA, 8-13 August 2004

Numerical Simulations of Breaking Waves


around an Advancing Ship by an Unstructured NS Solver
T. Hino (National Maritime Research Institute, Japan)
comparison of computed and measured flow field data
for a ship model with a blunt bow. The method employs an unstructured grid for its larger flexibility to
handle complex geometry.
There are two major ways to treat free surface flows
with Navier-Stokes solvers. One is an interface fitting
method in which a free surface configuration is explicitly defined and tracked and a numerical grid is aligned
to a free surface. The other way is an interface capturing method. In this approach, an interface shape is
not tracked explicitly, instead, a scalar quantity, such as
the volume fraction of each cell or the level set function, is used to indicate the interface location. The interface fitting method offers a high accuracy because
the free surface conditions can be imposed accurately
in the exact location of an interface. However, difficulty arises when a free surface shape deforms largely
in such cases as very steep waves or breaking waves.
In case of spilling breakers, breaking wave models
(for example Rhee et al, 2002) are used to prevent the
breakdown of computation. On the other hand, the interface capturing approach can be used when the interface deformation is large, although the accuracy of
boundary conditions is not so good as the interface fitting method. Another advantage of the interface capturing approach, which is particularly attractive for an
unstructured grid method for a complex geometry, is
that it does not need re-gridding due to the free surface
movement. The VOF (Volume of Fluid) approach has
been applied to simulate breaking waves around a ship
(Azcueta et al, 1999). The present method adopts another interface capturing approach, that is, the level set
method (Sussman et al, 1994) for free surface modeling with one-phase flow mode (Hino, 1999).
In the following section, the brief outline of numerical procedure is given. It is followed by the description
of computational results and the comparison with measured data. Conclude remarks is presented in the last
section.

ABSTRACT
CFD (Computational Fluid Dynamics) methods for
solving Reynolds Averaged Navier-Stokes equations
are being used as a design tool in various fields of fluid
engineering. However, there remain many problems
which are not easily analyzed by standard CFD methods. Wave breaking which often appears in free surface
flows around an advancing ship is one of such examples. Its simulations are difficult because of the underlying complicated physics such as interface topology change, air-water interaction or energy dissipation.
Yet, analysis of wave breaking is of great importance
in ship hydrodynamics, since the wave breaking resistance plays a significant role in propulsive performance
of a certain class of ships. Therefore, it is desirable that
CFD methods for ship design have capability to simulate breaking waves. In this paper, the capability of the
up-to-date unstructured Navier-Stokes method which
is under development at National Maritime Research
Institute, Japan is examined for breaking wave simulations around a ship model. The method employs an
interface capturing scheme for free surface treatment
and it is expected that it can cope with large deformation of free surface shape.
INTRODUCTION
CFD (Computational Fluid Dynamics) methods
have reached the stage in which they are used as a design tool in various fluid engineering fields. However,
there remain many problems which are not easily analyzed by standard CFD methods.
Wave breaking often appears in free surface flows
around an advancing ship is one of such examples. Its
simulation is difficult due to the underlying complicated physics such as interface topology change, twophase flow interaction or energy dissipation. Analysis
of wave breaking, on the other hand, is of great importance in ship hydrodynamics, since wave breaking
resistance plays a significant role in propulsive performance of a certain class of ships. Therefore, it is desirable that CFD methods for ship design have capability
to simulate breaking waves.
In this paper, the capability of the up-to-date CFD
method which is under development at National Maritime Research Institute, Japan is examined by the

NUMERICAL PROCEDURE
Flow Solver
The flow solver used in this study is called SURF
(Solution algorithm for Unstructured RaNS with
FVM). It is under development toward a practical ship

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Spatial discretization is based on a cell-centered


finite-volume method. A computational domain is divided into unstructured polyhedral cells and flow variables (pressure, velocity and eddy viscosity) are stored
in the center of each cell. Cells whose shape are hexahedra, tetrahedra, prisms or pyramids can be used and
the combinations of these cells give greater flexibility
for handling complex geometry.
For the inviscid fluxes (convection terms and
pressure gradient terms), the second order upwind
scheme based on the flux-differencing splitting of Roe
(Roe, 1986) with the MUSCL approach is employed.
The viscous fluxes are evaluated by the second order
central scheme. Thus, the overall accuracy in space is
second order.
The backward Euler scheme is used for the time
integration. Local time stepping method is used, in
which time increment is determined for each cell in
such a way that the CFL number is globally constant.
The linear equations derived from the time linearization of the fluxes are solved by the Symmetric GaussSeidel (SGS) iteration.
The turbulence model used in the present investigation is the one-equation model by Spalart and
Allmaras(Spalart et al, 1994).

design tool. The governing equations are the threedimensional Reynolds averaged Navier-Stokes equations for incompressible flows. In order to couple pressure with a velocity field, artificial compressibility is
introduced into the continuity equation with expense
of time accuracy, which means that the present formulation is for a steady state solution. The final form can
be written as follows:
q (e ev ) (f f v ) (g g v )
+
+
+
=0
t
x
y
z
(1)
and
q = [ p u v w ]T
In the above expressions all the variables are made dimensionless using the reference density 0 , velocity
U0 and length L 0 . Pressure p is modified as
p = p +

z
F2

where p is
the original pressure and F is the Froude
number, U/ gL0 , with z being the vertical coordinate. By this modification of pressure, the gravitational acceleration term can be dropped from the zmomentum equations. The velocity components in the
(x, y, z) direction is expressed as (u, v, w).
The inviscid fluxes e, f and g are defined as

u
v
w
u2 + p
vu
wu

e=
uv , f = v 2 + p , g = wv
w2 + p
uw
vw

Free Surface Treatment


Free surface is an interface between air and water in
the present applications. Free surface conditions consist of kinematic and dynamic conditions and they are
implemented in the interface capturing framework.
The kinematic condition is the condition that fluid
particles on a free surface remain on an interface. This
is written in a mathematical form as follows:

where is a parameter for artificial compressibility.


The viscous fluxes ev , f v and g v are written as:

0
0
0

xx

v
xy v zx

ev =
xy , f = yy , g = yz
zx
yz
zz
where
ij = (

DH
H
H
H
H
=
+u
+v
+w
=0
Dt
t
x
y
z

(2)

where a free surface shape is defined as


H(x, y, z; t) = 0
In the present scheme, this kinematic condition is formulated based on the localized level
set method(Peng et al, 1999) which improves
the efficiency of the original level set approach
(Sussman et al, 1994) used in the previous version of
the present code (Hino, 1999).
The level set function is defined as the signed distance from the interface, i.e.,

in water
>0
=0
on the interface
(3)

<0
in air

1
ui
uj
+ t )(
+
)
R
xj
xi

R is the Reynolds number defined as U 0 L0 / where


is the kinematic viscosity. t is the non-dimensional
kinematic eddy viscosity.
Since a numerical procedure for the Navier-Stokes
equations are described in Hino, 1997 and Hino, 1998,
only the brief outline is given here and the treatment of
free surface is described in the following sections.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Since (x, y, z; t) = 0 defines the free surface shape,


the kinematic condition can be satisfied if the following equation is used to update :
D

=
+u
+v
+w
=0
Dt
t
x
y
z

(4)

In the localized version of the level set method, the


two parameters 1 and 2 where 0 < 1 < 2 are introduced. The signed distance function is rewritten as
d(x, y, z; t) and the definition of the level set function
is modified as

if d > 1
1
d
if |d| 1
=
(5)

1
if d < 1
Thus, the level set function is localized within the
bandwidth 21 from the interface. The transport equation (4) is also modified as

+ C() u
+v
+w
= 0 (6)
t
x
y
z
where C() is the cut-off function defined as
C() =

(|| )2 (2|| + 3 )
2
2
1
3

)
2
1

if || 1
if 1 < || 2
if || < 2
(7)

from the cell centered values with the second order accuracy in the physical space. The gradient of at the
cell center used in the extrapolation above is obtained
by the least squares method. Furthermore, when a cell
face is quadrilateral, the face is divided into two triangles and the one point quadrature is used for each
triangle, which guarantees the second order accuracy
in the physical space. The time integration is carried
out by the Euler backward scheme as in the same way
as for the momentum equations.
In order to avoid reflection of free surface waves in
the outer boundaries of a computational domain, the
wave damping method (Hino, 1999) is used.
There is a singular behavior of the interface in the
region close to a solid wall. The no-slip condition imposed on a solid wall prevents the interface movement
there, while the interface in the outer region moves due
to the fluid motion. It causes the unphysical large deformation of near a solid wall. The extrapolation
approach is employed here to remove this problem, in
which the value of for the cells close to the wall is
extrapolated from the outer cell. The selection of outer
cells on unstructured grids also needs special attention
(Hino, 1999).
The re-initialization of the level set is an important
step in the level set method, since the level set function
is no longer a distance function after the convection.
The re-initialization process can be done using the partial differential equation as in Sussman et al, 1994 or
Peng et al, 1999.
Flow Variable Extrapolation

in such a way that the update of is performed only in


the region where || < 2 .
The numerical solution method for Eq.(6) is similar
to the flow equations. The cell centered finite-volume
discretization applied for the cell i yields
Vi



i
+ C(i )
(i+j)/2 U(i+j)/2 = 0
t
j

Since most of ship hydrodynamics applications require a flow field of water region only, one-phase flow
approach is used, i.e., flow variables in the air region are extrapolated from a water region in such a
way that the dynamic condition on free surface boundary is satisfied. This method also has an advantage that it is not necessary to cope with large density difference between air and water. At this point,
the present method differs from the original level set
method(Sussman et al, 1994) where two-phase flow
approach is employed.
The dynamic free surface conditions can be approximated by the following two conditions. First, the velocity gradients normal to the free surface are zero.
Second, the pressure on the free surface is equal to atmospheric pressure. In order to satisfy the first condition, the velocity components are extrapolated in
the direction normal to the interface. Following the
localized level set method (Peng et al, 1999), this is
achieved by solving the following equation in the air

(8)

where
U(i+j)/2 ui Sx,(i+j)/2 + vi Sy,(i+j)/2 + wi Sz,(i+j)/2
Vi is the cell volume and j is the neighbor cells of the
cell i. The subscript (i + j)/2 denotes the cell face
between the cells i and j and (S x , Sy , Sz ) are the area
vectors of the cell face. Special care should be taken
in the construction of the flux (i+j)/2 U(i+j)/2 . When
the flow is uniform and parallel to the flat interface,
the flat interface should be preserved. To achieve this,
(i+j)/2 , the value of on the cell face, is extrapolated

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Free Surface

| i |

C
A

Table 1: Waterline coordinate of the


model(SRI-BBM)
x/L
y/L
x/L
0.00000 0.00000
0.50000
0.00198 0.024267 0.60000
0.00785 0.047895 0.65000
0.01747 0.070261 0.67414
0.03059 0.090776 0.71034
0.04685 0.10890
0.74655
0.06582 0.12415
0.78276
0.08702 0.13614
0.81897
0.10987 0.14453
0.85517
0.13378 0.14912
0.89138
0.15000 0.15000
0.92759
0.20000 0.15000
0.96379
0.30000 0.15000
1.00000
0.40000 0.15000

| j |
J

Figure 1: Free surface pressure condition.


region where < 0 for the pseudo time .

q = 0

||

(9)

Note that the quantity /|| is the unit vector


normal to the interface whose direction is from water
to air. In the region away from the interface where
is constant, /|| is replaced by the vector
(0, 0, 1).
The pressure boundary condition is written as
p=

h
on the free surface
F2

(10)

(zc /F 2 )(|i | + |j |) pj |i |
|j |

where
zc =

y/L
0.15000
0.15000
0.15000
0.14935
0.14694
0.14166
0.13346
0.12223
0.10771
0.089610
0.067359
0.037416
0.00000

A ship model used is called SRI-BBM (Ship Research Institute- Blunt Bow Model) which is designed
for investigation of wave breaking around a blunt
bow of simple geometry. Flow field measurements
for this model have been carried out at Ship Research Institute (currently, National Maritime Research
Institute)(Hinatsu et al, 2001). Fig.2 shows the geometry of the model whose length, width and draft are
2.0m, 0.6m and 0.6m, respectively. The waterline
shape is a semi-circle of radius r/L = 0.15 at the
bow followed by a parallel part of a half ship length
and a smooth curve to the stern-end. Waterline coordinates are shown in Table 1. The hull shape is wallsided down to z/L = 0.15 and the remaining part
is a lower-half of a body of revolution with the profile
being same as the waterline.
Fig.3 shows the views of the computational grid.
Since the geometry is simple, the grid is generated as
a structured grid of O-O topology and the data structure is converted into the unstructured grid format. The
number of cells are 128 68 80 in the streamwise,
girth and normal directions. The grid points are clustered to the body surface and are densely distributed
in the interface region. A solution domain is on the
port side of the ship assuming symmetry of flow field
and its size is 2 x 4, 2.5 y 0 and
2.2 z 0.3, while a ship is placed at 0 x 1.
The averaged spacing adjacent to the solid wall is
3.16 106 .
As a flow condition, Froude number and Reynolds
number, based on a ship length, are set 0.3834 and

where atmospheric pressure is assumed to be zero and


h is the z-coordinate of the interface. For an air cell
which is next to a water cell, pressure is extrapolated
in the following way. Suppose that the cell i is the air
cell for which the pressure must be extrapolated and
the cell j is the neighboring water cell. From the definition, || is the distance to the interface with i < 0
and j > 0. The interface is supposed to be located
between the cell centers i and j. Thus pressure at the
cell i is extrapolated as
pi =

blunt bow ship

|i |zj + |j |zi
|i | + |j |

is the z coordinate of the point C on the interface (see


Fig.1) and z i and zj are the z coordinate of the center
of cells i and j. See Hino, 1999 for detail. In case that
an air cell has several adjacent water cells, the pressure
value is obtained by taking the average of the extrapolated values from each water cell. In the remaining air
region, pressure is extrapolated using Eq.(9).
RESULTS
Condition of Computation

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

0.4
0.3
0.2
y/L

0.1
0.0

-0.1
-0.2
-0.3
0.00

0.25

0.50
x/L

0.75

1.00

Waterline
0.4
0.3

Figure 3: Partial view of computational grid.

0.2
z/L

0.1

and wave contours between experiment and simulation. In these figures, it is observed that general features of a wave field have been captured by the present
computation. Bow wave forms an arc-like shape. It
is followed by the steep trough behind it. The second
wave crest is present slightly aft of a midship. However, there are some differences in the details of wave
configurations. A bow wave in the picture shows a turbulent surface due to spilling beaker, while a free surface shape in the computation is smooth except for the
small hump ahead of a bow. The grid resolution may
not be sufficient enough to capture breaking and/or
small-scale turbulent motions. Furthermore, the preset
steady state solution method cannot simulate unsteady
flows properly. The second wave crest in the experiment is breaking with air bubbles, which is also beyond
the capability of the present code which employs onephase flow approach. However, the computed wave
front of the second crest shows discontinuity, which
indicates wave breaking.
The same trend is seen in the comparison of wave
profiles along a ship hull shown in Fig.7. From the
fore-end to the bottom of wave trough, the simulated
wave profile agrees very well with the measured one,
where wave breaking is not present. The second wave
crest with strong breaking is not simulated well, although the weak discontinuity of a free surface shape
can be observed. The computed wave crest height is
smaller than that in the experiment. Note that in the experiment this region is a mixture of water and air as can
be seen in Fig.5, while in the computation flow field is

0.0

-0.1
-0.2
-0.3
0.00

0.25

0.50
x/L

0.75

1.00

Profile
Figure 2: Geometry of ship model.
3.4 106 , which corresponds to the experimental condition.
Flow Field Solution
Computation continued up to 3,000 iteration steps
with the CFL number of 5. At this stage, the resistance
and the number of cells within water still show fluctuations with the iteration step. Since the actual flow field
is unsteady due to strong wave breaking, it is reasonable that the present steady state solution method failed
to get convergence. In the following, the snapshot flow
data at 3,000-th step is analyzed to examine the capability of the present method. Time accurate simulation
is underway and the result will be presented in the presentation.
Fig.4 shows hull surface pressure and velocity vectors on a center plane together with wave profile.
Figs.5 and 6 shows the comparison of wave patterns

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Measured

Y/L

0.25

-0.25

-0.5

0.25

0.5

0.75

X/L

Computed

Figure 4: Computed flow field.

h/L

Figure 6: Comparison of wave contours. Contour interval h/L is 0.005.


0.08
0.06
0.04
0.02
0
-0.02
-0.04
-0.06
-0.08
-0.1

Measured
Computed

0.2

0.4

0.6

0.8

X/L

Experiment
Z
Y

Figure 7: Comparison of wave profiles along a ship


hull.

cations of a steep profile are different from each other.


Note that the computed wave profile in the stern region
is not stationary.
The magnified view of the second wave crest on the
plane at y/L = 0.16 (1% of ship length away from
the ship hull) is shown in Fig.8. Velocity vectors and a
wave profile exhibit wave breaking. It can be stated
that the present method has a capability to simulate
breaking waves although the magnitude of braking is
under-predicted.
The hull surface pressure distribution is compared
in Fig.9. Again, the comparison show good agreement
from the fore-end to the wave trough. In the mid ship
and aft part, the distribution patterns differ in the same
way as the wave profiles. The region of high pressure
beneath the second crest is narrower in the computation than in the experiment, which correspond to the
lower wave height in the computation.
Finally, the velocity and the wave profile in front of
a bow and a center plane (Plane A) and on a plane with
45 degrees from a center plane (Plane B) as shown

Simulation
Figure 5: Comparison of wave patterns.

assumed to be filled with water. The computed stern


wave shows very steep profile which corresponds to
the breaking wave in the experiment, although the lo-

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

0.4

0.05

0.2
z/L

0.0

Plane A

-0.05

-0.2

Plane B
r/L

-0.1

-0.4
-0.25

-0.15

0.00

0.25
x/L

0.50

0.75

1.00

-0.2
0.4

0.5

0.6

Figure 10: Location of velocity measurement sections.

0.7

measured velocity magnitude beneath a free surface


is reduced, this suggest energy dissipation with the
wave breaking/turbulent free surface effect in the experiment. These features are not simulated well in the
computation. The computation also cannot reproduce
the vortical motion beneath the free surface in front of
a body which is seen in the experiment. The same tendency is observed in the comparison of Plane B.
In order to clarify the reason for discrepancies, grid
refinement tests should be carried out first. In addition
to that, effects of the steady state formulation and one
phase flow approach should be examined. These are
subject to future investigations.

Figure 8: Breaking waves near the midship at y/L =


0.16.

0.2

0.1

Z/L

-0.1

-0.2

-0.3
0

0.25

0.5

0.75

X/L

Measured

CONCLUDING REMARKS
The capability of the up-to-date CFD method which
is under development at National Maritime Research
Institute has been examined by the comparison of computed and measured flow field data for a ship model
with a blunt bow.
In general, wave breaking is not simulated well
by the present method, although the simulated flow
field agrees reasonably well with the experimental one
where wave breaking is not present and the breaking in
the midship region is captured to some extent.
Poor grid resolution in the interface region may be
one reason for this discrepancy. However, there seem
to be more subtle limitations in the present method
for simulating breaking waves . Steady-state solution methods like the present method may not simulate
breaking waves which are essentially unsteady. Twophase flow approach may be necessary for simulations
of strong wave breaking. Especially if the air trap plays

0.2

0.1

Z/L

-0.1

-0.2

-0.3
0

0.25

0.5

0.75

X/L

Computed
Figure 9: Comparison of pressure distribution on a
ship hull. Contour interval Cp is 0.1.
in Fig.10 are compared in Figs.11 and 12. On the
center plane (Plane A), the wave profile in the experiment is less steep than in the computation and the

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

PLANE A

a certain role in the breaking process, two-phase flow


approach becomes essential.
Further investigations should be focused on these
subjects to improve the present methods capability of
wave breaking simulations.

0.12
0.1
0.08

Z/L

0.06
0.04

ACKNOWLEDGMENT

0.02
0

The author would like to thank staffs of Center for


CFD Research, National Maritime Research Institute,
for their valuable discussions and suggestions.

-0.02
-0.3

-0.2

-0.1

X/L

Measured
0.12

REFERENCES

0.1
0.08

Rhee, S.-H. and Stern, F.,RANS Model for Spilling


Breaking Waves, J. Fluids Engineering, Vol.124, 2002,
pp.424 432.

Z/L

0.06
0.04
0.02
0

Azcueta, R., Muzaferija, S., Peric, M., Computation of


Breaking Bow Waves for a Very Full Hull Ship, Proc.
Seventh Int. Conf. Numerical Ship Hydrodynamics, 1999,
pp. 6.2-1-6.2-11.

-0.02
-0.3

-0.2

-0.1

X/L

Computed

Sussman, M., Smereka, P. and Osher, S., A Level Set Approach for Computing Solutions to Incompressible TwoPhase Flow, J. Comput. Phys., Vol.114, 1994, pp.146159.

Figure 11: Comparison of velocity distributions on a


center plane y = 0. Velocity components are (u, w).
PLANE B

Hino, T., An Interface Capturing Method for Free Surface


Flow Computations on Unstructured Grids, J. of the Soc.
Naval Archit. Japan, Vol.186, 1999, pp.177 183.

Z/L

0.1

Hino, T., A 3D Unstructured Grid Method for Incompressible Viscous Flows, J. of the Soc. Naval Archit.
Japan, Vol.182, 1977, pp.9 15.

0.05

-0.2

-0.1

Hino, T., Navier-Stokes Computations of Ship Flows on


Unstructured Grids, Proc. of the 22nd Symp. on Naval
Hydro., 1998.

R/L

Measured

Roe, P.L., Characteristic-Based Schemes for the Euler


Equations, Ann. Rev. Fluid Mech., Vol. 18, 1986, pp.337
365.

Z/L

0.1

Spalart, P.R. and Allmaras, S.R., A One-Equation Turbulence Model for Aerodynamic Flows, La Recherche
Aerospatiale, No.1, 1994, pp.5 21.

0.05

-0.2

-0.1

Peng, D., Merriman, B., Osher, S., Zhao, H. and Kang,


M., A PDE-Based Fast Localized Level Set Method, J.
Comput. Phys., Vol.155, 1999, pp.410-438.

R/L

Computed

Hinatsu, M, Tsukada, Y., Fukasawa, R. and Tanaka,


Y.,Experiments of Two Phase Flows for the Joint Research Proc. SRI-TUHH mini-Workshop on Numerical
Simulation of Two Phase Flows, Ship Research Institute,
2001.

Figure 12: Comparison of velocity distributions on a


plane of 45 degrees from a center plane. Velocity components are (u t , w), where ut is the component tangential to the measurement section.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Arthur Reed
Naval Surface Warfare Center, Carderock Division,
USA
The author notes discrepancies between the
waves ahead of the bow by computation and
measurement. I note that the computations fail to
capture the stagnation flow ahead of the bow. This is
seen in the experiments as a predominantly vertical
flow on the centerline immediately ahead of the bow.
Stagnation flow is a steady phenomena that should be
captured by the computations, even with a simple
free surface breaking code. An examination of why
the stagnation flow is not captured might help to
resolve the differences between the computations and
experiments.
AUTHORS REPLY
Fig. A1 shows the computed streamlines in
the center plane in front of the bow. The present
computation captures the feature of the stagnation
flow although the computed stagnation point is
higher than the experimental one. In Fig.A2, the
front views of the surface pressure distributions are
compared between the computation and the
measurement. Again, the high (stagnation) pressure
zone in the computation is located higher than in the
measurement. This difference is most likely due to
the poor grid resolution in the vertical direction,
although the verification is needed.
0.08

Z/L

0.06

0.04

-0.1

-0.05

X/L

Fig.A1 Computed streamlines on the center plane.


Cp=1
0.1
0.08
0.06

Measured

Computed

0.04

Z/L

0.02
0
-0.02
-0.04
-0.06
-0.08
-0.1
-0.12
-0.2

-0.1

0.1

0.2

X/L

Fig.A2 Front view of surface pressure distributions


in the bow.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics


St. Johns, Newfoundland and Labrador, CANADA, 8-13 August 2004

Experimental Measurements of the Surface of a Breaking


Bow Wave
Anna Karion, Thomas C. Fu, James R. Rice, Don C. Walker and Deborah A. Furey
(Naval Surface Warfare Center, Carderock Division, USA)
A BSTRACT
A prismatic wedge was towed in fresh water in the David
Taylor Model Basin at the Naval Surface Warfare Center,
Carderock Division (NSWCCD), generating a large bow
wave. Towing speeds ranged from 0.7 to 4.6 m/s, and
drafts ranged from 0.6 to 1.5 m. These conditions correspond to Froude numbers from 0.2 to 1.4 (based on
model draft, D). In addition to the variations in draft and
speed, two different bow geometries were investigated:
one with a 20 degree bow entrance angle, 20 degree flare,
and sharp leading edge, and one with a 40 degree bow entrance angle, no flare, and rounded leading edge. Measurements of free-surface elevations near the bow were
made using a laser imaging technique. Some results
from the experiment have been presented in a previous
paper (Karion, et al. (2003)). The current paper presents
further analysis of the experimental data. Fluctuations
on the free surface are quantified, and characteristics of
the breaking region are studied in more detail. Information about the breaking region can aid those attempting
to model wave breaking with either advanced numerical
techniques or with simpler empirical methods.

rough breaking. Turbulence and velocity measurements


in two-dimensional breaking ocean waves were conducted by Melville, et al. (2002) on a larger scale, and on
a model-scale bow wave by Roth, et al. (1999). Surface
roughness measurements, similar to those presented
here, were presented by Walker, et al. (1996) for a
two-dimensional steady breaking wave. The experiment
described in this paper was designed to characterize the
breaking region of a large three-dimensional bow wave
with energetic breaking.
In the current experimental study, two different bow geometries and leading edges are used to generate bow waves under otherwise identical conditions.
A laser-sheet based imaging technique is used to measure free-surface heights for a variety of speeds and
three different model drafts. Results for averaged freesurface profiles and maximum wave height scalings from
this experiment have been presented in a previous paper
(Karion, et al. (2003)). This paper presents free-surface
fluctuations and the location and size of breaking regions
for the various conditions of the experiment, as well as
surface roughness measurements for two conditions.
E XPERIMENTAL SETUP AND INSTRUMENTATION

I NTRODUCTION
Bow wave dynamics have been the subject of analytical,
experimental,
and numerical
research
in
the
past
(including
Ogilvie (1972), Miyata and Inui (1984), Wyatt (2000),
Roth, et al. (1999),
Waniewski, et al. (2002),
and
others).
At present, a large variety of computational methods exists for modeling freesurface flows near surface ships (Wyatt (2000),
Sussman and Dommermuth (2001), and many others).
As numerical methods have grown more sophisticated,
developers have begun to attempt to model the breaking
bow wave. Experimental data in this breaking region
is needed for validation of these advanced numerical
techniques. While previous experiments have also investigated bow waves (Ogilvie (1972), Roth, et al. (1999),
Miyata and Inui (1984), Waniewski, et al. (2002)), most
were conducted with smaller models and did not exhibit

A large bow wedge model (Model No. 5605) was towed


in the Deep Water Towing Basin at the David Taylor Model Basin of the Naval Surface Warfare Center,
Carderock Division. This fresh-water basin is 15.5 m
wide and approximately 575 m long. During the experiment, the depth of the basin varied between 6.3 m (when
the model was at the 1.5 m draft) and 5.4 m deep (at the
0.6 m draft). The water level was lowered to reach the
different drafts for the experiment, while the model itself
remained at a fixed height on the carriage.
Model No. 5605 is a large prismatic wedge with
dimensions as indicated in Figure 1. One end (referred
to as the fine bow) has a 20 degree entrance angle and is
flared at 20 degrees, with a fine leading edge (0.16 cm
radius of curvature). The other end (referred to as the
full bow) has a 40 degree entrance angle and is straightsided, with a rounded leading edge (1.27 cm radius of
curvature). For the first part of the experiment, the model

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Figure 1: Diagram of wedge model with dimensions (in meters). The fine bow is on the right in the figure, and the full bow
on the left.

Table 1: Matrix of run conditions for which there was bow


wave breaking for the fine bow.

D (m)
0.6
0.6
1.1
1.1
1.5
1.5
1.5
1.5

U (m/s)
2.0
2.9
2.6
3.9
2.0
3.1
3.9
4.6

Fr
0.8
1.2
0.8
1.2
0.5
0.8
1.0
1.2

was towed with the fine bow forward, and for the second
part, the model was mounted in the opposite direction
(with the full bow forward). Thus, measurements of the
free surface were made for both configurations. Figure 2
shows an example of the bow wave generated by the full
bow (in the photograph, the model is running from left to
right).
The run conditions are summarized in Table 1
for the fine bow and Table 2 for the full bow. In the
tables, D represents the draft (the immersion depth of the
model, measured from the free surface at zero velocity),
U the towing velocity, and Fr the Froude number. The
Froude number is defined as follows:
U
Fr = ,
gD

Figure 2: Example of the bow wave generated by the full bow


at D = 1.1 m and U = 2.6 m/s. The model is painted black
with a white grid. The vertical grid spacing is 30.5 cm, and the
horizontal spacing is 30.5 cm if measured along the centerline
of the model. The small horizontal marks at the top of the grid
are the top of the model.

Table 2: Matrix of run conditions for which there was bow


wave breaking for the full bow.

D (m)
1.1
1.1
1.1

(1)

where g is the coefficient of gravitational acceleration.


The conditions shown in these tables are those for which
the rough breaking region was analyzed. The experiment
included additional run conditions at slower speeds but

Copyright National Academy of Sciences. All rights reserved.

U (m/s)
1.7
2.6
3.1

Fr
0.5
0.8
1.0

Twenty-Fifth Symposium on Naval Hydrodynamics

down the basin. Images were collected for 2 seconds at


30 frames per second at each of approximately 180 locations along the hull (every 2.54 cm) for the fine bow, and
115 locations for the full bow.

Figure 3: Diagram of laser sheet and camera setup, viewed


from above. In this case, the model would be moving from left
to right. When the full bow was forward, the traversing system
was mounted on the full bow.

there was no breaking in the bow wave, so data from


those conditions is not presented here.
Quantitative Visualization and Data Analysis
A non-intrusive optical technique, Quantitative Visualization (QViz), has been developed to measure the freesurface deformations occurring in regions commonly inaccessible using more traditional measurement methods,
i.e. near wake flows, bow waves and breaking waves.
Using a laser sheet and video camera, the QViz system illuminates the surface of interest and collects digital images representing instantaneous cross sections of
the spray envelope and surface profiles (further details
can be found in Furey and Fu (2002)).
In the current experiment, a laser light sheet
was projected onto the water surface perpendicular to the
side of the hull. The laser used was a diode-pumped,
solid-state YAG laser, with an output of 2.53.0 W at
532 nm (Model MLM-0532 by Melles-Griot), transmitted through two fiber optic cables to cylindrical lenses
mounted inside laser probes, producing two laser sheets.
The laser sheets were next to each other, in the same
plane, generating effectively one sheet with two times
the width. A standard video camera was mounted facing aft, collecting images of the laser sheet reflection
off the water surface. The camera and fiber optic laser
probes were mounted to a rail system that was attached
to a traverse, which was in turn mounted on the model
itself (see Figure 3). The two fiber optic laser probes
formed a laser sheet approximately one meter wide. The
camera was mounted approximately 1.5 meters forward
of the laser sheet and angled down towards the water surface. The traversing system was automated and coupled
with the data acquisition software so that once the desired number of images was acquired, the laser sheet and
camera system moved aft together and the camera began
acquiring images at a new position. Thus, depending on
the run time, many positions were covered on each pass

Digital images from the video camera were collected at 30 frames per second using a National Instruments frame-grabber board and a personal computer. An
image analysis program was developed at NSWCCD
using National Instruments LabView software with the
Image Processing (Vision) toolbox using built-in edgedetection routines to extract the surface profile information. For the present analysis, the breaking region of
each image was analyzed separately, and the resulting
water line determined as a function of time. The image size was 640480 pixels, covering a viewing area
of approximately 1.51.2 meters. Thus the lowest possible uncertainty (approximately 1 pixel) was equal to
0.25 cm. Distortion due to camera placement and viewing angle was corrected using an image of a calibration
grid with equally spaced points and a calibration algorithm in National Instruments LabView IMAQ Vision
software package. The largest error in the system was
introduced by the camera placement and the calibration
method. For example, error in the calibration would result if the grid was not held perfectly square and at the
correct distance relative to the camera. It is estimated
that the total error on the free-surface elevations is 1cm.
However, this is an error affecting the determination of
the absolute location of the free surface. The relative
error in the free-surface measurement from one frame
to the next (affecting the fluctuations of the surface presented here) is much lower, estimated to be at the pixel
error value of 0.25 cm. Future effort to reduce the error
on these types of measurements is focusing on placing
the cameras on automated pan-and-tilt units, so that their
orientation is exactly known, and the calibrations can be
performed in a more controlled environment.
The images collected in this experiment were
taken with an interlaced camera, so that each image was
acquired in two fields, 1/60 of a second apart in time. In
an interlaced image, the first field is composed of the odd
pixel lines and the second the even. For the current analysis, these fields were separated to minimize the blurring
of the moving surface (the shutter speed of the camera
was also at 1/60 second). The effective vertical resolution for each image was thus halved to 240 pixels, and
the number of images doubled. The LabView software
used to perform this separation of the interlaced fields
simply interpolated to regenerate the intermediate pixels. Thus, at each location, the two seconds of data at 30
frames per second, each composed of two fields, resulted
in 120 images for analysis.
An example of an image that has been analyzed
in this way is shown in Figure 4. The red line superimposed on the surface of the wave is the edge that has been

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

0.6

Fine Bow Wave Profile, D = 1.5 m, FrD = 1.0, U = 3.9 m/s (7.5 kt), X = 2.6 m

0.4

0.26

0.2

z (m)

0.25
0

0.24

z (m)

0.2

0.23

0.4

0.22
0.6
0.2

0.2

0.4

0.6
y (m)

0.8

1.2

0.21

1.4

0.2
0.7

Figure 4: Example of a QViz image of the bow wave generated


by the full bow, with the edge detected by the data analysis
algorithm superimposed in red. The speed of this run is 2.6 m/s,
and the draft is 1.1 m.

extracted using the analysis. The sides of the image seem


distorted because the image has been processed using a
calibration grid to correct for the distortion due to camera placement. The model appears on the left side of the
image.

0.75

0.8

y (m)

0.85

0.9

0.95

Figure 5: An example of the mean free-surface location in the


breaking region, in the red points, with the blue dots showing
the band one standard deviation away. The standard deviation
is calculated for each point based on an analysis of 120 frames
over two seconds.

R ESULTS
D = 1.5 m Fr = 1.0

Surface Fluctuations
The Quantitative Visualization (QViz) data was used to
determine the variation of the free surface with time.
Two seconds of data images were collected at each fixed
laser sheet location; these were analyzed to determine the
fluctuation of the free surface. Surface fluctuations were
determined only on the breaking region of the bow wave;
that is, in the region beginning where the bow sheet first
impinged back onto the free surface. In this area, the
laser light sheet was scattered very effectively due to the
turbulent, multiphase nature of the flow. Therefore, the
region analyzed was that where the image was clearly
and significantly brighter than the surrounding background. The image was thresholded at a high pixel value
that differed based on the condition, but was usually at
an approximate value of 200 out of 255, and then the
edge-detection algorithm was executed to find the topmost edge. Because this region had high contrast levels,
the image processing revealed relatively accurate measurements of the free surface. The non-breaking parts of
the free surface, although previously analyzed in a mean
fashion (Karion, et al. (2003)), were not clear enough for
a frame-by-frame analysis.

median RMS (meters)

0.015

0.01

0.005
raw data
smoothed data

0
0

0.5

1
1.5
2
x (meters from stem)

2.5

Figure 6: Standard deviation of the free-surface profile as a


function of distance along the hull, x, for the fine bow at Fr =
1.0, D = 1.5 m, and U = 3.9 m/s.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

The median of the standard deviation along the


y dimension is shown as a function of x, the distance
along the hull, for the fine bow at D = 1.5 and Fr = 1.0
in Figure 6. The magnitude of the fluctuations varies significantly along the hull with no apparent trend, for this
and the other conditions. It was observed, when analyzing the fluctuations from one run to another and correlating them to the time of day of the run, that the fluctuation magnitude was significantly lower for runs performed when the basin water was its calmest (for example, for the first run of the day or after a long break).
The results show that fluctuations in the free surface of
the wave are strongly dependent on the initial conditions.
During the experiment, these initial conditions were not
tightly controlled. Although the wait time between runs
was roughly constant at 30 minutes, some runs were performed when the water was significantly calmer, such as
the first run of the day or after a long break in runs to
make an equipment repair. There was also an effort made
not to run a fast speed immediately following a slowerspeed run, but this was not always possible due to tight
scheduling.
Figure 7 shows the median value of the standard deviation (taken over all positions x) as a function
of draft Froude number. There is considerable scatter in
the data, due to the reasons explained above. Figure 8
shows the same data normalized by the maximum average wave height for each condition. This figure shows
that in general, the magnitude of the surface fluctuations
scales with the maximum wave height, and varies from
between two and six percent.
Surface Roughness Measurements
Wavenumber spectra have been constructed by performing a fast-fourier transform of each detected edge (i.e.
the free-surface profile for each frame). The goal of performing these transforms was to establish a spectrum of
the turbulent length scales in the breaking region of the
wave, and to determine whether the spectrum changes
with the wave conditions, such as Froude number. Spectra were only calculated for the full bow because the surface of the breaking region of the fine bow was too short

0.03

fine bow, D = 0.6 m


fine bow, D = 1.1 m
fine bow, D = 1.5 m
full bow, D = 1.1 m

0.025

RMS (m)

0.02

0.015

0.01

0.005

0
0

0.5

Fr

1.5

Figure 7: Median standard deviation of the free-surface profile,


as a function of draft Froude number. The error bars represent
the standard deviation for each point.

0.1

fine bow, D = 0.6 m


fine bow, D = 1.1 m
fine bow, D = 1.5 m
full bow, D = 1.1 m

0.08

RMS / Zmax

The free-surface location was determined for


each of the 120 images at a certain laser sheet position,
and its mean calculated as a function of y position along
the wave, where y is defined as the distance from the
models centerline, in the direction perpendicular to the
hull (i.e., along the laser sheet). The standard deviation
of this mean was also calculated as a function of location
in the sheet, y. An example of the result is shown in Figure 5 for the fine bow at Fr = 1.0. The band around the
mean free-surface profile indicates the magnitude of the
fluctuations at each location along the free-surface profile.

0.06

0.04

0.02

0
0

0.5

FrD

1.5

Figure 8: Median standard deviation of the free-surface profile,


normalized by the maximum wave height, as a function of draft
Froude number. The standard deviation for each point ranges
from 0.005 to 0.02.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Full Bow, D = 1.1 m

10

FrD = 0.8, X = 0.50 m


Fr = 1.0, X = 0.50 m
D
Fr = 0.8, X = 0.75 m
D
Fr = 1.0, X = 0.75 m

Power density

10

10

10

10

10

10

10

10

wavenumber (rad / m)

Figure 9: Wavenumber spectrum for the breaking region of


the bow wave for the full bow at two velocities (2.6 m/s, or
Fr = 0.8, and 3.1 m/s, or Fr = 1.0) and at two locations along
the hull.

in length, usually ten centimeters or less. The straightsided bow with its rounded leading edge produced a wide
breaking wave, as contrasted with the narrow, plunging
breaker of the fine bow. The reflection of the laser sheet
off of the wide breaking region of the wave provided a
free-surface profile that was long enough to allow for
performing the FFT (see Figure 4).
The resulting wavenumber spectra are shown in
Figure 9 for two different speeds (2.6 m/s and 3.1 m/s)
at two different locations (0.5 m and 0.75 m from the
stem). The draft is 1.1 meters. The wavenumber results
show that the power spectrum at the higher wavenumbers, or shorter wavelengths, does not vary significantly
with Froude number. This result is consistent with
results shown for a two-dimensional spilling breaker
by Walker, et al. (1996). It appears that the two cases
at higher speeds have slightly more energy at lower
wavenumbers than the two cases at lower speeds. However, this phenomenon would need to be studied experimentally in more detail in the future to make a firm conclusion. These somewhat limited results do support the
conclusions of Walker, et al. (1996), however.
Extent of Breaking
The extent of breaking occuring in the bow wave for each
condition was determined from the QViz images. Essentially, for each image, the region in which the laser
reflected very brightly off of the surface was defined as
the breaking area. This was the area in which the laser
reflected off of the multiple air-water interfaces of the
wave. Although this determination is somewhat subjective, it was found that there was a very clear contrast in

Figure 10: Contour plot of free-surface height with the location of breaking superimposed in black. This run is for the fine
bow, and has D = 1.5 m, U = 3.9 m/s (7.5 kt) and Fr = 1.0.

the images between this region and the rest of the free
surface, and effort was made to be consistent. This was
the same region that was used for the surface fluctuation
and surface roughness analyses described above.
Figures 10 and 11 show contour plots of the free
surface with the breaking area, as defined above, superimposed in black, for the fine bow and full bow, respectively. It should be noted here that the contour plots were
also generated using the same laser sheet data, but analyzed in a different manner (see Karion, et al. (2003) for
more contour plots and details regarding this analysis).
The mean width of the breaking region in the direction perpendicular to the hull (i.e., in the plane of the
laser sheet) for the fine bow is shown in Figure 12 as a
function of draft Froude number. Figure 13 includes the
values for the full bow on the same plot, with a new scale.
The breaking width has been normalized by the characteristic length for the wave, U 2 /g. This normalization relates the width of breaking to the wavelength, (2)U 2 /g.
Figure 12 illustrates that the extent of breaking does increase with increasing Froude number, even when normalized, indicating that the fraction of the wavelength
that exhibits rough breaking on the surface is not constant. Figure 13 shows that the hull shape makes a very
significant difference, as the width of the breaking region is much larger for the full bow than for the fine bow.
The extent of breaking for the full bow seems to decrease
from Fr = 0.8 to Fr = 1.0. This may be at least partially
due to a transition in the form of the wave between these
two conditions. The bow wave transitions from a spilling
breaker that breaks violently ahead of the stem (for Fr =
0.8) to a plunging breaker that begins breaking behind
the stem, once the bow sheet impinges on the free surface (for Fr = 1.0).

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

fine bow, D = 0.6 m


fine bow, D = 1.1 m
fine bow, D = 1.5 m
full bow, D = 1.1 m

0.9
0.8

breaking

/ (U /g)

0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
0

0.5

FrD

1.5

Figure 13: Width of the breaking region for the fine and full
bow, normalized by U 2 /g, as a function of draft Froude number.

Figure 11: Contour plot of free-surface height with the location of breaking superimposed in black. This run is for the full
bow, and has D = 1.1 m, U = 2.6 m/s (5 kt) and Fr = 0.8.

0.15

0.1

breaking

/ (U2/g)

fine bow, D = 0.6 m


fine bow, D = 1.1 m
fine bow, D = 1.5 m

0.05

0
0

0.5

Fr

1.5

Figure 12: Width of the breaking region for the fine bow, normalized by U 2 /g, as a function of draft Froude number.

C ONCLUSIONS
Data from laser-sheet visualization of the bow wave of a
large towed wedge has been analyzed with the goal of understanding features of the turbulent surface of the breaking wave. The large size of the model was necessary to
produce a wave large enough to exhibit characteristics
of waves in the field, such as spray and bubble generation and turbulent, multiphase breaking. The model was
towed at three different drafts, various speeds, and with
two different bow shapes, yielding a unique data set for
comparison with computations.
Three different results are presented that will
aid researchers modeling breaking bow waves. The magnitude of the free-surface fluctuations is presented, serving as a measure related to the turbulent energy of breaking. The surface roughness on the face of the breaking
wave is presented for cases in which the wave was wide
enough to enable a spectrum analysis. The wavenumber
information supports previous work in this area, although
not enough of the data was able to be analyzed to be conclusive. Lastly, the area and location where breaking occurs is shown; the extent of the breaking, normalized by
U 2 /g, increases with draft Froude number and is significantly larger for the straight-sided (full) bow than the
flared (fine) bow. The data and conclusions here can be
used to aid in the development of breaking models in the
future.

ACKNOWLEDGEMENTS
This research is supported by ONR under contract numbers N0001403WX20225 and N0001403WX20728. Dr.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

L. Patrick Purtell is the program manager.


R EFERENCES
Furey, D.A. and Fu, T.C., Quantitative Visualization (QViz)
Hydrodynamic Measurement Technique of Multiphase Unsteady Surfaces, 24th Symposium on Naval Hydrodynamics,
2002.
Karion, A., Sur, T.W., Fu, T.C., Furey, D.A.,
Rice, J.R. and Walker, D.C., Experimental Study
of the Bow Wave of a Large Towed Wedge,
8th Intl Conference on Numerical Ship Hydrodynamics,
2003.
Melville, W.K., Veron, F. and White,, C.J., The velocity field
under breaking waves: coherent structures and turbulence,
Journal of Fluid Mechanics, Vol. 454, 2002, pp. 203233.
Miyata, H. and Inui, T., Non-linear ship waves,
Advances in Applied Mechanics, Vol. 24, No. 1, 1984,
pp. 215288.
Ogilvie, T. F., The Wave Generated by a Fine Ship Bow,
9th Symposium on Naval Hydrodynamics, 1972, pp. 1483
1524.
Roth, G.I., Mascenik, D.T. and Katz, J., Measurements of
the flow structure and turbulence within a ship bow wave,
Physics of Fluids, Vol. 11, No. 11, 1999, pp. 35123523.
Sussman, M. and Dommermuth, D. G., The numerical simulation of ship waves using cartesian-grid methods, 23rd Symposium on Naval Ship Hydrodynamics, 2002,
pp. 762779.
Walker, D.T., Lyzenga, D.R., Ericson, E.A. and Lund, D.E.,
Radar backscatter and surface roughness measurements for
stationary breaking waves, Proc. R. Soc. Lond. A, Vol. 452,
1996, pp. 19531984.
Waniewski, T.A., Brennen, C.E. and Raichlen, F., Bow Wave
Dynamics, Journal of Ship Research, Vol. 46, No. 1, 2002,
pp. 115.
Wyatt, D.C., Development and Assessment of a Nonlinear Wave Prediction Methodology for Surface Vessels,
Journal of Ship Research, Vol. 44, No. 2, 2000, pp. 96107.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION

the breaking region. Are the wavenumber


spectra shown in the Figure from a single frame
or averaged over multiple frames? Is it possible
to estimate typical time scales as well from the
QViz images?

Trish Sur
Science Applications International Corporation, USA
I would like to ask the following questions.

AUTHORS REPLY
1.

Typical results for the Quantitative Visualization


(QViz) technique for the bow wedge experiment
are presented in Figures 10 and 11 of the present
paper and in Figures 5 through 12 of a previous
companion paper (Karion et al., 2003). These
figures show a region close to the bow wedge for
which QViz is unable to obtain measurements of
the free surface elevation. Similarly, there is an
outboard limit to the QViz measurement area.
For the QViz measurements used in the present
analysis, did the breaking areas have missing
portions? If so, how would this affect the
characteristics of the free surface fluctuations
reported herein?

AUTHORS REPLY
There are two reasons that data is missing
from the contour maps in Figures 10 and 11. The
first is that the light was shadowed behind the crest of
the wave, keeping the camera from viewing the area
behind the wave crest. The breaking regions were at
the wave crest and were not affected by this
shadowing. The second, the outboard limit to the
measurement area, does however apply to the
breaking areas as well. As the measurement system
and camera moves aft, the crest of the wave moved
out of the cameras field of view.
Visual
observations indicate that often by this point the
rough breaking had dissipated and was spread out
over a wide area, but it is possible that some breaking
roughness was there and not measured. It is not clear
how the breaking regions aft and outboard of the
measurement area would affect the characteristics of
the free surface fluctuations. Figure 6 illustrates that
there is no clear trend in the fluctuations as a function
of distance downstream of the stem. Therefore, there
is no reason to believe that the characteristics of the
breaking region outside the measurement region is
different from those of the region measured (at least
within our accuracy).
DISCUSSION

The wave number spectra are averaged over


all the frames for a particular run, which is to say for
2 seconds. In the presentation at the conference I will
show additional results averaged over 25 positions
for each speed. The answer to the second question is
unfortunately no, it is not possible to estimate typical
time scales. The QViz data images were acquired at
a standard 30 frames per second, and the fluctuations
occur at a higher rate than this. Fluctuations slower
than 15 frames per second (to satisfy the Nyquist
sampling criterion) could be resolved but it was
found that the fluctuations are faster than the
sampling rate in this case. A high-speed video
system would be necessary to capture fluctuation
time scales.
DISCUSSION
Trish Sur
Science Applications International Corporation, USA
3.

The present paper alludes to using this


characterization of the breaking region to help
develop breaking models. To your knowledge,
has this been done and/or are there any specific
plans along these lines?

AUTHORS REPLY
I am not aware of anyone developing
breaking models based on our data (although there
are others who have developed their own models
based on acceleration or slope). However, there are
plans to use this data and other tow-tank data to
generate a breaking model at NSWC Carderock.
REFERENCES
Karion, A., Sur, T. Waniewski, Fu, T.C., Furey,
D.A., Rice, J.R., and Walker, D.C., Experimental
Study of the Bow Wave of a Large Towed Wedge,
Proc. 8th International Conference on Numerical Ship
Hydrodynamics, 2003.

Trish Sur
Science Applications International Corporation, USA
2.

Figure 9 of the present paper defines the length


scales associated with the surface roughnesses in

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Stephen M. Scorpio
Johns Hopkins University, USA
I would like to thank the authors for making
progress on a challenging measurement. Would it be
possible to make a quantitative comparison with
Walkers 2D breaking wave measurements by
measuring surface fluctuations normal to the
breaking wave crest? Walker suggested a model
spectrum where the spectral roll-off was related to
the width of the turbulent plume. Does your data
support this hypothesis?
AUTHORS REPLY
We do not believe it is possible to use the
data from this experiment to quantify the
wavenumber spectrum in the direction perpendicular
to the breaking wave crest. The laser sheet was
oriented perpendicular to the hull itself, and thus the
instantaneous wave profiles are only known in this
direction. Sequential x positions were one inch apart
and could not be used to reconstruct a continuous
surface in any arbitrary direction.
(The second part of this question is
answered in our reply to Dr. Ericsons second
discussion question.)

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Richard Leighton
General Dynamics, USA
What ancillary measurements were made?
To expand on that, for validation purposes, there are
more robust measurements for comparison. For
example, the pressure along a transverse cut on the
hull. Without a robust and independent baseline
measurement, the detailed comparisons are suspect.
AUTHORS REPLY
Unfortunately, there were no additional
measurements made during this experiment other
than those reported in the given paper and in Karion
et al. (2003). These include laser sheet wave height
data and high-speed video of spray droplets as well as
extensive video.
REFERENCES
Karion, A., Sur, T., etc. 8th International Symposium
on Numerical Hydrodynamics, September 2003.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION

AUTHORS REPLY

Leonard Imas
Stevens Institute of Technology, USA

The curves shown in figure 5 are actually the


time-averaged wave height (in the center, in red),
with the lines above and below the curve showing the
band one standard deviation away from this average
in either direction. Therefore, the standard deviation
is not actually time-averaged. The wave profile is
time-averaged, and the standard deviation gives an
idea of how much the wave profile fluctuates with
time.

This paper describes a study in which a


systematic study to investigate wave breaking due to
the motion of a surface-piercing body was carried out
by performing a large-scale tow-tank experiment.
What makes this work unique is that (a) draft-based
Froude, Reynolds, and Weber number combinations
were varied parametrically to investigate breaking
wave kinematics and (b) free-surface topology during
the wave-breaking events was studied. Very little
data of this kind is available in published literature.
Congratulations should be extended to the authors for
undertaking an experiment of this magnitude,
especially one where quantitative measurements were
obtained successfully. A few comments on the
discussion presented in the paper:
1) On page 1, it would be more consistent to
acknowledge that the computational research
tools cited are not unique in their capabilities;
namely, commercially available computational
tools are also fully capable of modelling wave
breaking but at the same time, they too require
further validation. Data such as that presented
here is very useful in this respect.
AUTHORS REPLY
We certainly agree and only cite certain
methods as examples of some of the tools that
researchers are currently using.
Commercial
computational tools are also rapidly becoming
sophisticated in their treatment of wave breaking and
multiphase flows.

DISCUSSION
Leonard Imas
Stevens Institute of Technology, USA
3) Regarding Figure 5 (pg. 4), would any additional
information be evident if these results are plotted
as a 3D surface where the axes are z, y, Fr?
Namely, is the behavior self-similar?
AUTHORS REPLY
The wave profiles at a given position are
different enough that a 3D surface plot would not be
appropriate.. However, the following two figures
each show three wave profiles (with standard
deviation bands) at the same position. The first
figure shows profiles at the same Froude number, but
different drafts, and the second shows profiles at
three different Froude numbers but the same draft.
All are quite different.

DISCUSSION
Leonard Imas
Stevens Institute of Technology, USA
2) On page 4, in Figure 5, the standard deviation
curves which are shown are actually timeaveraged. It would be of interest to see how
these curves appear frame by frame over the two
second period for the different Froude numbers
at which the model was tested. Namely, how
much information is removed by performing this
averaging?

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

AUTHORS REPLY
Unfortunately, the data set from this
experiment only allowed for a wavenumber analysis
of the full bow data at two Froude numbers, 0.8 and
1.0. The figure below illustrates the average wave
number spectrum for each case (averaged over x
location). The figure further illustrates that the
curves differ in the low wavenumber region, but it is
unclear how this effect might change at a different
Froude number condition.
2

10

Fr = 0.8
Fr = 1.0

4) The effect of ambient conditions on the level of


surface fluctuations is discussed on page 5. This
is important to note as assigned free-stream and
ambient turbulence levels are often taken for
granted when free-surface CFD computations are
performed.
AUTHORS REPLY

Power density

10

10

10

10

10

It is unfortunate that during the course of the


experiment the amount of ambient surface fluctuation
was not specifically measured before each run. This
makes it difficult to quantify the effect of ambient
conditions on the fluctuations of the bow wave. The
experiment only serves to show that there is certainly
an effect, which was noted in the experiment by
differences in fluctuation magnitudes based on the
time of day of a given run.

10

10

10

wavenumber (rad / m)

Similar conclusions to those reached based on the


results presented (in Figures 10 13), on extent of
wave breaking were recently obtained using VOF
simulations (Imas 2004).
DISCUSSION
Leonard Imas
Stevens Institute of Technology, USA

DISCUSSION
Leonard Imas
Stevens Institute of Technology, USA
5) As the authors point out, surface fluctuations are
important as a measure related to turbulence
energy in breaking. Using this measured quantity
for CFD code validation should prove useful as at
present time, there is limited understanding
about the accuracy of uRANS and LES turbulence
models performance in the breaking wave zone.
6) The surface roughness measurements presented
in Figure 9 essentially indicate that the breaking
wave region is spectrally broad-banded. It would
be useful to produce this curve for a broader
range of Froude numbers, both above and below
1.0. Namely, do the differences in the curves
become larger or smaller at lower wavenumbers?

To conclude, I have found the data measured


by the authors as part of this experiment very useful
for purposes of code validation on two separate
occasions now and would like to thank them again
for sharing their data and commend them for the high
quality of their work.
REFERENCES:
Imas, Leonard, Simulation of Free-Surface Flow
Around the DTMB Wedge Body, presentation given
at the ONR Ship Wave Breaking and Bubble Flow
Review, La Jolla, California, March 2004.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Eric A. Ericson
Johns Hopkins
Laboratory, USA

University

Applied

Physics

The breaking bow wave measurements and


analysis presented in this paper address important
issues regarding free-surface hydrodynamics.
Detailed measurements such as these are critical to
understanding and predicting the generation of
surface wakes by ships.
A couple of issues came to mind as I read
the paper. The first concerns the overturning bow
sheet. From the paper, I get the impression that the
bow sheet is not present for all run conditions and
that its presence impacts the form of the breaking
bow wave. Can the laser sheet measurements be used
to characterize the bow sheets in any way? In
particular, it would be interesting to identify the
location where the bow sheet impinged back onto the
free surface, quantify its distance from the hull, and
study its impact on the following breaking crest.
The second issue concerns the low
wavenumber roll-off of the surface elevation spectra
shown in Figure 9. As the authors note, these spectra
have forms similar to those observed for the
stationary breaking waves presented by Walker et al.
(1996). The spectra from Walker et al. had a low
wavenumber roll-off determined approximately by
2 divided by the plume length, where the plume
length is the distance from the breaking crest to the
toe of the turbulent region on its forward face. Can
the plume lengths of the breaking bow waves be
estimated from the laser sheet measurements? If so, it
may be interesting to see if the same low
wavenumber roll-off dependence applies.

The location at which the bow sheet impinges on


the free surface is somewhat difficult to
determine with sufficient accuracy to be useful.
It can probably be estimated, however, by
relating it to the point where the multiphase
nature of the breaking begins that is, where the
laser light sheet begins to be reflected very
brightly in the images. This is probably the point
of initial splash-up of the wave, which probably
coincides or is at least close to the point of
impingement by the bow sheet. This location
can be found in Figure 10 at the forward tip of
the shaded area indicating the extent of breaking.
In this example, the impact of the jet probably
occurs close to x = 1 meter from the stem and y
= 0.5 meters.
2) Although the plume length specifically has not
been determined, the average width of the
breaking region has, and is shown in Figures 12
and 13 of the paper. This quantity, defined in
the paper as the region where the laser reflects
brightly of the multiphase flow in the breaking
region, is defined in the direction of the laser
sheet. The plume length perpendicular to the
crest can be determined by multiplying the
avergage breaking width reported in the paper by
the cosine of the angle between the wave crest
and the laser sheet (for both Froude numbers it
was measured to be close to 40 degrees). The
plume length varies along the crest of the wave
(or x position along the wedge), as does the peak
frequency of the wavenumber spectrum.

AUTHORS REPLY
1) The authors believe that the bow sheet exists for
all the breaking conditions of the fine bow. That
is, the breaking wave generated by the fine bow
was a plunging breaker, characterized by a bow
sheet riding up the side of the hull, overturning,
and impinging back on the free surface. The
current paper only addresses conditions in which
the bow wave was breaking. There were
certainly experimental conditions that did not
exhibit the bow sheet, but they are not
considered to be breaking for the purpose of this
paper, because whitewater was not observed in
the laser sheet. The full bow, in contrast, did not
always exhibit a bow sheet because the bow
wave was sometimes more of a spilling wave
than a plunging wave with a clear bow sheet.

The figure above shows the frequency at which


the peak occurs for each x position for each of
the two Froude numbers. The x axis shows the
plume length at that position, which is the
breaking width along the laser sheet after the
cosine correction. The dashed line in the figure
is the equation y = 2/x, following the

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

suggestion of the discussor. It seems to match


the data very well, a fact which suggests that the
dominant wavelength is equal to the length of the
plume.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics


St. Johns, Newfoundland and Labrador, CANADA, 8-13 August 2004

An Experimental Investigation of Breaking


Bow Waves Simulated with a 2D+T Technique
Mostafa Shakeri, Xinan Liu, Sven Goll and James H. Duncan
(University of Maryland, United States of America)
has to be large as well and, of course, the model must
be operated at the same Froude numbers as the full
scale ship. The second major issue that must be
addressed in order to obtain realistic bubble
populations is that the experiments must be performed
in salt water. It is well known that in salt water air
entrainment processes produce many more small
bubbles than in fresh water, Monahan and. Zeitlow
(1969), Monahan et al. (1994), Slauenwhite and
Johnson (1999).

ABSTRACT
The purpose of this experimental study is to explore
the physics, surface profiles, flow fields, and bubble
distributions in breaking bow waves of high-speed
ships.
These laboratory experiments must be
performed at relatively large scale and in salt water to
insure that the bubble size distributions are similar to
those in full scale ships. In order create these large
waves, a two-dimensional deformable wave maker is
being used. The profile of the wave maker at each
instant in time (t) is set to match the vertical section of
a ship model at a streamwise location that is a distance
Ut from the stem, where U is the ship model speed.
The wave maker has a draft of 0.91 m and an
equivalent full beam of 2.83 m. It is capable of speeds
that simulate 30 knots full scale speed for a ship with a
draft-to-length ratio of 23.11. The wave maker and
the wave tank are built to be used with salt water. In
this paper, measurements of the profiles of the
breaking bow wave in fresh water are presented for
two cases with equivalent full scale ship speeds of 20
and 25 knots.

In this paper, the above problems are addressed by


studying experimentally the dynamics and bubble
generation in simulated breaking bow waves. These
waves are generated by a method known as 2D+T in
which a two-dimensional wave maker moves
horizontally and deforms in a manner so that its surface
approximates the intersection of the profile of the
three-dimensional ship hull and a fixed vertical plane
as the ship moves in a direction normal to the plane.
Tulin and Wu (1996) have shown through numerical
calculations that the 2D+T method produces waves that
are quite similar to those of 3D ships. In the remainder
of this section, we show a simple comparison of
experimental data to illustrate these similarities. Dong
et al. (1997) made an experimental investigation of free
surface flow around a 3D ship model using a PIV
technique. The model was 3.05 m long and the model
speed ranged from 0.914 m/s to 2.44 m/s. Dong et al.
also photographed the surface flow patterns using a
camera that was mounted on the towing carriage.
Figure 1(a) (taken from their paper) shows an extended
exposure photograph of the bow wave system around
the ship model at a Froude number (based on model

INTRODUCTION
Breaking bow waves in the flow around high-speed
ships are of great practical and scientific interest. The
primary goal of the current experiments is to explore
the dynamics of air entrainment by breaking bow
waves and to provide information on the bubble
populations and velocity fields created by these
breakers at positions as far downstream as the stern of
a 3D ship model. In these experiments, two major
issues must be taken into account in order to achieve
realistic bubble populations. First, the breaking bow
waves must be large enough to make the effect of
surface tension small relative to the available kinetic
energy thus allowing the creation of small bubbles,
similar to those created in the large waves of full scale
ships. In order to generate large waves, the ship model

length Fr = U /

gL , where U is the model speed, g

is the acceleration of gravity and L is the model length)


of 0.334. The water surface was marked with
Aluminum powder. Since the wavelength of the bow
wave is short, the breaker is strongly affected by

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

surface tension and appears as a spilling breaker with


no plunging jet.
As yet, there are no complete experimentally
determined 2D+T wave profiles for direct comparison
with the work of Dong et al. However, some
information can be gained using the experimental data
on short wavelength unsteady two-dimensional
breakers given in Duncan et al. (1999). The breakers
were generated mechanically by a plunging wave
maker and forced to break far from the wave maker. In
these experiments, crest profile histories of the
breakers were measured using a photographic
technique that employs a high-speed digital camera, a
laser light sheet and fluorescent dye (see the following
section for more details on this measurement
technique). A sequence of crest profiles from a weak
spilling breaker is shown in Figure 2. This wave
breaking event is also dominated by surface tension
and occurs without the formation of any plunging jet.
The data is plotted in a manner that gives the 2D+T
equivalent of the wave pattern on the starboard side of
a ship model as it moves vertically down in the figure.
In spite of the fact that this wave was not created with
any hull displacement in the vicinity of the breaker (as
it would have been in a 2D+T simulation), many
features in the surface pattern are qualitatively similar
to those found in the photograph by Dong et al. (1997).
In particular, the leading edge of the breaking zones
have the same shape and the curved trajectories and
number of the ripples are quite similar in the two cases.
While not proving the accuracy of a 2D+T simulation,
this comparison gives credibility to the similarity of the
two flows.

Figure 2. 2D+T representation of wave profiles of a


two-dimensional spilling breaker generated by
dispersive focusing. From Duncan et al. 1999}

Experimental Details
The experiments are being carried out in a wave tank
that is 14.80 m long, 1.15 m wide, and 2.20 m deep
with a water depth of 1.83 m. The 2D+T wave maker
was built to simulate the 5415 model that was used in
tests at the Naval Underwater Warfare Center,
Carderock. (Other ship profiles can be simulated as
well by modifying the computer control software.)
Unfortunately, it is not possible to simulate the bow
bulb in the 2D+T experiments so this feature was
removed from the wave maker profile sequence. The
beam to draft ratio of the 5415 model is 3.11 and the
length to draft ratio is 23.11. The draft of the 2D+T
wave maker is 0.91 m. Thus, the equivalent beam and
length of the 3D model simulated by the 2D+T wave
maker are 2.82 m and 21.03 m, respectively.
A schematic drawing of the wave maker is shown in
Figure 3. The wave maker is powered by four
servomotors, which drive four vertically oriented
shafts. Each shaft drives a toothed pulley, which drives
a tube via a rack-and-pinion-like system. The drive
tubes, in turn, drive horizontally oriented drive plates

Figure 1. Bow wave pattern around a ship model


with a length of 3.05 m, Fr=0.334. From Dong et al.
(1997).

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

that are as wide as the tank (1.14 m) and guided along


the tank walls by tracks. The edges of the plates
farthest from the pistons are attached to the skin of the
wave board via hinges. Position sensors and motor
tachometers are used in a computer-based feed back
control system to achieve the desired motion of each
piston. Most of the wave maker is submerged inside
the tank and the wave maker frame is bolted to the
bottom and sidewalls of the wave tank. The wave
maker is made of appropriate materials to resist
corrosion due to salt water.

Wave board

Servo motors
Keel bar
Figure 4. A photograph of the wave maker in the fully
extended position (midships).
In order to examine the accuracy of the wave maker,
two measurements are shown in Figure 5 and 6. Both
measurements were done at a full scale equivalent ship
speed of 25 knots, Fr= 0.346. The wave maker motion
lasts 2.25 seconds at this equivalent speed. Figure 5
is a plot of the measured horizontal position of each
drive tube versus time along with the desired position
data for each tube taken from horizontal cuts of the
hull of the 5415 model. The drive tube position data
was measured with the linear position sensors used for
feedback for the wave maker control system (see
Figure 3). Note that measured and desired data are in
very close agreement. Figure 6 is a plot of the shape of
the wave board at various times and the shape of the
5415 hull profiles at the corresponding streamwise
positions along the length of the model. The wave
board position measurements were taken from
photographs the wave maker during its extension.
Note the excellent agreement between the two sets of
data with the exception of the bottom of the first
profile where the top of the bulbous bow can be seen in
the 5415 profile data.

Position
sensors

Figure 3. A schematic drawing of the wave maker.


The wave board bends over the top of the keel bar
assembly. The keel depth of the model is at the top of
the keel bar.
The waves are generated by the motion of the flexible
wave board that is driven by the four drive tubes. The
wave board, which spans the width of the wave tank, is
constructed from interleaved thin stainless steel plates
of various lengths. The plates are slotted and riveted
together in such a way that does not allow any straightthrough flow of water from one surface of the wave
board to the other. The stainless steel plates are thin
enough to bend elastically under the differential action
of the pistons. Each piston is attached to a different
layer of stainless steel via hinges so that as the pistons
move out at different speeds, the changing distance
between the hinge points is accommodated by the
stainless steel plates sliding relative to each other.
The effective keel of the model is created by the fixed
keel bar over which the wave board bends and slides,
see Figure 3. A photograph of the wave board in the
tank and in the fully extended position is shown in
Figure 4.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

fluorescent dye and upon illumination by the laser light


sheet the dye glows with a greenish-yellow color. Also
attached to the instrument carriage is a high-speed
digital camera that was operated at 250 pictures per
second. Each picture consists of an array of 512 by
512 pixels with eight-bit gray levels. The glowing line
seen in each image is the intersection of the light sheet
and the water surface and its position and shape yield
the crest profile in each image via digital image
processing techniques. The carriage is mechanically
connected to the top drive tube of the wave maker so
the camera and light sheet move at the time varying
speed of the upper part of the wave board.

60

Offset (inches)

50
40
30
20
10
0
0

50

100

150

200

250

300

Time (arbitrary units

Figure 5. Position versus time of the four drive tubes of


the 2D+T wave maker plotted along with the
corresponding positions for the 5415 model
(essentially horizontal plan cuts of the hull profile).
The two sets of four curves are in very close agreement
and are therefore indistinguishable in the plot.
Profile Near
stem

High-speed movie
camera

Mirror

Laser ligh
sheet

Profiles near
Mid ship
Carriage moves
with top drive plate

60

50

z(in)

40

1.82 m

30

Water
mixed with
fluorescene
dye

20

10

10

20

30
y(in)

40

50

60

Figure 7.
apparatus.

Figure 6. Hull profiles of the 5415 model at various


streamwise sections (solid lines) and corresponding
measured positions of the wave board of the 2D+T
wave maker (data points).

1.15 m

Water surface profile measurement

Results
A sequence of six photographs from a single run with
the wave maker is shown in Figures 8 (a) through 8 (f).
The sequence covers the first 452 ms of the
experimental run. This run has an equivalent full scale
ship speed of 25 knots (Fr = 0.346) and is completed in
2.25 s. The image widths are about 50 cm. In Figure 8
(a), the wave maker has just started to move and the
water surface is nearly flat. The wave board is vertical
at the left side of the image. The bright nearly
horizontal line on the left at the bottom of the image is
the intersection of the light sheet and the water surface.
After 20 ms (Figure 8 (b) the water surface has begun
to rise up the wave board. At t=144 ms, Figure 8 (c),
the water surface has risen much further up the wave
board and a jet directed horizontally to the right is

The first step in the experiments is to measure the


profile histories of the wave system generated by the
2D+T wave maker. Important characteristics of the
profiles of the breaking waves include the crest height,
wave steepness, and the plunging jet thickness and
velocity just prior to impact with the front face of the
wave. These measurements are taken photographically
with the system shown in Figure 7. Illumination is
provided by an Argon Ion Laser that is mounted
outside the tank. The laser beam is directed by a series
of mirrors to the instrument carriage upon which is
mounted a system of mirrors and lenses to convert the
beam into a light sheet that is oriented vertically along
the center plane of the tank. The water is mixed with

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

beginning to form. The jet begins to fall toward the


water surface at about t = 280 ms, see Figure 8 (d). As
the jet falls air is either entrapped in the top or captured
as the jet impacts the water surface, see Figures 8 (e)
and (f).

Figure 8 (c). t = 144 ms. The jet begins to form.


Figure 8 (a). t = 0 ms. The wave board has barely
begun to move.

Figure 8 (d), t = 280 ms. The jet begins to fall due to


gravity.

Figure 8 (b). t = 20 ms. The water begins to rise on


the wave board.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

The maximum water surface height at the last profile


processed is about 300 mm.

Figure 9 Profiles of the wave crest generated at the


bow for an equivalent full scale speed of 25 knots. The
total run time is 2.25 seconds and the time between
water surface profiles is 4 ms.

Figure 8 (e). t = 400 ms. Air can be seen in the jet tip.

The time histories of the height of the water contact


line on the wave board and the horizontal speed of the
plunging jet were recorded for runs with 2.8-second
(Fr=0.2755) and 2.25-second (Fr=0.346) durations
these runs correspond to 20 knot and 25 knot
equivalent full scale speeds, respectively. The contact
line height data for the 2.8-second and 2.25-second
runs are shown in Figures 10 and 11. respectively.
The maximum heights of the contact lines are 18 cm
and 25 cm for the 2.8-second and 2.25-second runs,
respectively.

20
18
16
Contact line height (cm)

Figure 8 (f). t= 452 ms. The jet begins contact with the
front face of the wave.
Data from image sequences such as the one from
which the samples were taken for Figure 8 are being
processed to obtain the history of the wave crest
profile. A sample data set is shown in Figure 9. The
plot contains about 100 profiles from a run with a 2.25second duration. The time between profiles is 4 ms. It
can be seen that the water surface rises quickly up the
wave board and flattens across the top at about the
same time that a horizontally moving jet is formed.

14
12
10
8
6
4
2
0
0

0.2

0.4

0.6

0.8

1.2

Time (seconds)

Figure 10. Contact line height versus time for 2.8second run.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

30

Conclusion

Contact line height (cm)

25

These preliminary experiments demonstrate the


feasibility and important features of bow wave
generation with the 2D+T technique. In this
technique, it is possible to generate very large energetic
waves in a facility filled with salt water. The waves
are at least qualitatively similar to bow waves
generated by 3D ship models and the relationship
between these two wave systems will be explored
further in the future. One of the main goals to be
addressed in future work is the relationship between
the geometric characteristics of the bow waves prior to
jet impact and the turbulent fluid motions and bubble
populations after jet impact. Flow fields will be
measured with particle image velocimetry and bubbles
will be measured with various techniques including
shadow graphs and in-line holography. It is hoped that
this data can be used in numerical models to predict
empirically the air entrainment characteristics of the
breaking bow wave of high-speed ships.

20
15
10
5
0
0

0.2

0.4

0.6

0.8

1.2

Time (seconds)

Figure 11. Contact line height versus time for 2.25second run.
Another interesting quantity is the horizontal speed of
the jet as shown in Figure 12. Data for both the 2.8second run and the 2.25-second run are shown in the
figure along with the velocity of the top drive tube
which is located 15.24 cm above the undisturbed water
level. The velocities are non-dimensionalized by the
maximum velocity of the top drive tube ( U max ) in

Reference

each case and time is multiplied by U max . As can be

Tulin M. P. and M. Wu, 1996, Divergent bow


waves, Proceedings of the 21th Symposium on Naval
Hydrodynamics, Trondheim, Norway, June 1996.

seen from the figure, the horizontal speed of the jet


was as high as 1.7 times U max and still rising when the
last data points were taken for the jet speed.

Dong R. R., J. Katz, T. T. Huang, 1997, On the


structure of bow waves on a ship model, Journal of
Fluid Mechanics, Vol. 346, pp. 77-115.

1.8
Channel 1 velocity, 20 knots
Jet horizontal velocity, 20 knots
Channel 1 velocity, 25 knots
Jet horizontal velocity, 25 knots

1.6

Duncan J. H., H. Qiao, V. Philomin and A. Wenz,


1999, Gentle spilling breakers: crest profile
evolution, Journal of Fluid Mechanics, Vol. 379, pp.
191-222.

1.4

Velocity/U

1max

1.2

0.8

Monahan E. C. and C. R. Zeitlow, 1969, Laboratory


comparisons of fresh-water and salt-water whitecaps,
Journal of Geophysical Research, Vol. 74, pp. 69616966.

0.6

0.4

0.2

20

40

60
Time*U1max (inches)

80

100

Monahan E. C., Q. Wang, X. Wang, and M. B.


Wilson, 1994, Air entrainment by breaking waves: A
Laboratory assessment, Aeration Tech. ASME Fluids
Engineering Division, Vol. 187, pp. 21-26.

120

Figure 12. Horizontal velocities of the jet tips and the


top drive tube as a function of time for the 2.8-second
and 2.25-second wave maker motions..

Slauenwhite D. E. and B. D. Johnson, 1999, Bubble


shattering: Differences in bubble formation in fresh
water and seawater, Journal of Geophysical Research,
Vol. 104(C2), pp. 3265-3275.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Alan Brandt
Johns Hopkins University, USA
Is there an expected scale to the intermittent
fingering noticeable at the leading edge of the
spilling breakers? A scale could be associated with
the eddy structure in the turbulent region.
AUTHORS REPLY
We are planning to rotate the laser light
sheet and camera by 90 degrees about a vertical axis
so that we can measure the surface structures along
the cross stream length of the jet. We agree that there
will probably be a relationship between the dominant
length scale of these surface structures and the length
scales of the turbulent velocity field in cross stream
planes, at least just after jet impact with the front
wave face. However, we are not sure that the PIV
measurements that are planned for the underlying
flow will be sufficient to determine the turbulent
length scales in the cross stream plane. In any case,
we will attempt to address this issue.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Richard Lahey, Jr.
Rensselaer Polytechnic Institute, USA
To make good local void fraction
measurements one can use a rake of tubes with quick
closing ball valves at either end. At a given time all
valves are closed capturing the local air volume (i.e.,
void fractions). This device works well in high void
fractions where optical devices wont work. The
authors are encouraged to investigate the use of these
devices.
AUTHORS REPLY
Thank you for your interesting suggestion.
We will explore the possibility of this measurement
technique. Though we can see the advantages of this
measurement technique in statistically stationary twophase flows, we will need to address difficulties
caused by the fact that in the breaking wave the highvoid fraction zone is in a region of space that is
initially above the mean water level in the tank and
the high-void fraction zone is moving rapidly in
laboratory coordinates.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics


St. Johns, Newfoundland and Labrador, CANADA, 8-13 August 2004

Field Measurements of Bow Spray Droplets


T. Waniewski Sur and K. Chevalier
(Science Applications International Corporation, USA)
ABSTRACT
The bow spray is a very prominent hydrodynamic feature of the flow around naval combatant ships; it must be
included as part of computational fluid dynamic models
if a complete hydrodynamic prediction is desired. Modelling the generation and evolution of large quantities of
droplets over a wide range of scales requires specialized
procedures, and full scale field measurements of bow
spray are needed for model development and validation.
This paper presents full scale measurements
of bow spray droplets from a field experiment led by
the Scripps Institution of Oceanography (SIO) aboard
the R/V Roger Revelle on a transit from Lyttelton,
New Zealand to Pago-Pago, American Samoa in February/March 2002. SAIC deployed a high-speed digital
video camera to examine the generation of spray droplets
by the breaking bow wave over a region extending from
zero to six meters aft of the forward perpendicular. Ship
forward speeds ranged from 1.0 to 7.7 m/s, and sea state
conditions ranged from sea state zero to three. Ship
motion data was collected and the flow around the R/V
Roger Revelle was documented by nine standard digital
video cameras throughout the field experiment.

namic feature, it continues to be extremely challenging


to model. Not only is the overturning shape of the bow
wave free surface highly three-dimensional and complex,
but the flow is also multiphase in nature. Portions of the
bow wave sheet thin down to the point where it ruptures
into myriad spray droplets. Other portions of the bow
wave remain intact, curl over, and impinge on the free
surface entraining air bubbles into the flow around the
ship. At Science Applications International Corporation
(SAIC), an ongoing computational ship hydrodynamics
effort focuses on these modelling challenges. Several
years ago a simple bow spray model (described in a later
section herein) was developed as part of this effort; however, suitable experimental data was not available for its
validation even though the study of sprays has been an
active research area.

INTRODUCTION

Interest in sprays arises from many practical applications related to power, propulsion, heat exchange,
and material processing. There are also natural occurrences of sprays. A recent review of the science and
technology of droplets and sprays may be found in Sirignano (1999). Experimental investigations have been and
continue to be useful for exploring the physics of spray
generation and evolution. Dai et al. (1997) studied turbulent primary breakup of annular wall jets and correlated drop sizes at the onset of turbulent primary breakup
by equating the surface energy required to form a drop
to the kinetic energy of an eddy of corresponding size.
Sarpkaya and Merrill (1999) also focused on the physics
of ligament and drop formation at the free surface of liquid wall jets, flowing over smooth and sand-roughened
plates. They found that the root mean square value of
v 0 was quite close to the initial ligament ejection velocity, thereby relating the ligaments to the internal structure of the flow. Turbulent breakup of the liquid sheet is
widely accepted as the main source of spray generation;
however, it should be noted that some groups observed
instabilities in the liquid sheet which also contribute to
spray droplet generation (see, for example, Shroff and
Liepmann (1997)).

The breaking bow wave has been of interest to the


naval hydrodynamics community for the past several
years. Even though there has been a significant amount
of research dedicated to understanding this hydrody-

While these and other laboratory experiments


were carefully performed and documented, resulting empirical correlations cannot be incorporated into a full
scale numerical model without uncertainty because the
scaling of the spray generation process is unknown. Vi-

Not only do the high-speed images provide detailed records of full-scale bow wave break up into spray
droplets, but image processing yielded a quantitative
characterization of the spray including droplet sizes and
velocities. Variations in spray droplet size and velocity
with ship Froude number and with distance aft of the forward perpendicular were observed. It was also observed
that the bow wave and spray generation were quite unsteady even for the mildest sea state conditions. Correlations between the pitch angle in a ship-fixed coordinate
system and the spray sheet elevation showed that the ship
motion relative to the incoming waves is a key parameter
in determining the extent of the bow spray sheet.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Figure 1: Photographs of the bow flow at twenty knots at two different scales: the USS Cole (left) and David Taylor Model Basin
model 5415 (right).

sual inspection of model and full scale flows at the same


Froude numbers (see Figure 1) reveals simple Froude
scaling is inadequate. There are significant differences
in the appearance of bow spray; the bow sheet does
not break up into spray droplets in the towing tank. It
is likely that the spray generation process scales with
both Reynolds and Weber numbers in addition to the
Froude number. In addition, ambient conditions (wind
and waves) strongly effect the spray generation process
in the field. Seawater also contains surfactants which
can accumulate and reduce the local surface tension by
as much as 0.02 N/m (Kaiser et al., 1988). For these
reasons, it is important to obtain full scale field measurements of bow spray characteristics and this paper
presents the first attempt (to the authors knowledge) to
use high-speed photography and image processing techniques to study full scale bow flows in the field.

Figure 2: R/V Roger Revelle with approximate location of


SAIC camera boom marked by the blue line. The green lines
are the forward and aft boom control lines; they passed through
mooring holes and were secured by two cam cleats at the common topside location.

The high-speed monochrome video camera


used by SAIC (Photron FASTCAM-PCI 2K/16m) had
three main components: (1) a small camera head, 160
(W) x 330 (D) x 180 (H) mm, (2) a full-size standard
PCI board, and (3) a 16 m long cable which connected
the camera to the board. The camera had an adjustable
frame rate from 30 to 2,000 frames per second. Its imaging sensor was a square pixel, 250 Hz progressive scan
CCD and the scan area was 4.8 (H) x 3.6 (V) mm, or
1/3 inch. Full resolution, 512 by 480 pixels, could be
achieved up to frame rates of 250 frames per second. It
had a recording capacity of 512 Mb, or 2,176 images at
full resolution. The video was recorded using two Cmount lenses: (1) a 75 mm fixed focal length lens or (2)
a 100 mm zoom lens at full zoom.

R/V ROGER REVELLE FIELD EXPERIMENT


The R/V Roger Revelle field experiment was conducted
on a transit from Lyttelton, New Zealand to Pago-Pago,
American Samoa in February/March 2002. Ship forward
speeds ranged from 1.0 to 7.7 m/s, and sea state conditions ranged from sea state zero to three. Various instruments were deployed by SIO and photographers from
the Naval Surface Warfare Center, Carderock Division
(NSWCCD), extensively documented the flow around
the ship using nine digital video cameras (Ratcliffe et al.,
2003). Ship motions were recorded using a six-degreeof-freedom motion package (Seatex MRU 5) installed on
the R/V Roger Revelle. SAIC deployed a high-speed
digital video camera to examine the generation of spray
droplets at locations along the bow wave ranging from
zero to six meters aft of the forward perpendicular.

In the field experiments, the high-speed video


camera was housed in a protective enclosure with an optical glass window and mounted on an adjustable boom
that facilitated camera deployment across a large range
of positions. The boom was located on the port side of
2

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

the R/V Roger Revelle, situated so that when the boom


was hanging vertical it was adjacent to the forward perpendicular (see Figure 2). Three adjustable lines controlled the boom position. Regulation of the forward
and aft control lines determined the fore/aft angle of the
boom with respect to vertical. Similarly, the outboard
angle of the boom was controlled by the outboard line,
which was arranged with a 2:1 purchase system in order
to facilitate the adjustment of the boom. Additional lines
were used to control the attitude of the camera housing.
All of the control lines went to a common topside location to permit adjustment by a single person.

Figure 5: Illustration of the high-speed video camera and lens.

Images of Calibration Target

A typical camera/boom deployment positioned


the camera approximately 6.9 m away from the spray
generated by the breaking bow wave (see Figure 3). Using a 100 mm zoom lens, the field of view was approximately 26 cm horizontal by 25 cm vertical resulting in
a resolution of 1.9 pixels/mm. Since this field of view
did not encompass the entire bow wave, the camera was
repositioned to look at different sections of the bow wave
for a given forward speed. The camera was also calibrated by recording images of a stationary grid target located at a known distance from the camera lens. The grid
was imaged on the main deck of the R/V Roger Revelle.
A tuffcam, a rugged 30-frames-per-second video camera, provided by NSWCCD photographers was mounted
on top of the high-speed video camera housing. Since
it had a larger field of view, it was used to aim the highspeed video camera and to simultaneously record images
with GPS time code. It documented the overall bow flow.

Before each group of high-speed video runs of the bow


spray droplets, images of a grid distortion target (Edmund Industrial Optics) were taken on the main deck of
the R/V Roger Revelle. The white reflective Mylar target
is 206 by 280 mm, and a 150 by 150 mm grid of black
dots is printed on it. The dot diameter is 30.13 mm
(similar in size to the spray droplets), and the dot center spacing is 60.0008 mm. The target was oriented in
a vertical plane, perpendicular to the high-speed video
camera line of sight, for each working distance. A threedimensional model of the high-speed camera boom and
the R/V Roger Revelle hull was used to determine the
working distances from measurements of the control line
lengths . The images from the calibration target runs
were used to determine the field of view for each working
distance even though this can be calculated using similar triangles to relate the field of view to the working
distance, lens focal length, and CCD sensor size. An illustration of this relationship is included as Figure 5. For
both lenses the agreement between the measured and calculated values was good.

Over the ten-day experiment, more than 100


high-speed video image sequences were recorded and
typical images are shown in Figure 4. The ship forward speed for these image sequences ranged from 1.0 to
7.7 m/s (2 to 15 knots), but the majority of the sequences
were recorded at 6.2 m/s (12 knots).

Images of Spray Droplets


For each run, the AVI file was viewed first to determine
whether or not the images had the contrast and focus required for successful image processing. If the contrast
and focus were good, then the TIFF image files were processed. The first fifty TIFF files in each image sequence
were used to select the best grayscale threshold value for
the image sequence so that the water appear white and
the air would appear black. Then the remainder of the
sequence was processed in consecutive groups of fifty
due to memory limitations. The main image-processing
steps were to:

IMAGE PROCESSING
Both the tuffcam and the high-speed video were used to
make measurements of bow spray characteristics. The
tuffcam video was used to estimate elevations of the bow
wave/spray sheet by scaling the image from distinctive
markings on the R/V Roger Revelle ship hull. These
videos also highlight the unsteady nature of the bow
wave due to the at-sea conditions. The high-speed image
sequences were processed using custom software written
in MATLAB (The MathWorks, Inc.) together with the
Signal Processing and Image Processing toolboxes. The
key image-processing steps for the calibration and spray
droplet images are presented, and a discussion of the error associated with these techniques follows.

1. Import fifty images into the MATLAB workspace.


2. Use grayscale threshold to convert image from
grayscale to binary.
3. Filter binary images to remove noise using medfilt algorithm.
3

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Figure 3: Forward looking view of a typical deployment of the SAIC camera boom and high-speed video camera aboard the R/V
Roger Revelle (left). Aft looking view of a deployment with a large outboard offset (right).

Figure 4: Part of a typical raw image sequence (eight frames out of 2,176) from the high-speed video camera of the spray droplets
generated by bow wave break up. The sequence reads from the left to the right and from the top to the bottom; the elapsed time in
between successive frames is 0.4 seconds. The field of view of the camera was approximately 26.5 cm (H) by 24.8 cm (V).

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

4. Label objects, groups of connected on-pixels, using


four-connected neighborhoods.

Qualitative Description of Bow Spray Generation


The generation of bow spray is a complex process that
depends on many parameters including turbulence, gravity, surface tension, liquid sheet geometry, velocity distribution in the sheet, surface shear, roughness of the
contact surfaces, temperature distribution in the jet, pressure fluctuations within and outside liquid sheet, acoustic excitation, external flows, intentionally imposed disturbances, and foreign particles (Sarpkaya and Merrill,
1999). Both the tuffcam and the high-speed video enhance the observations from the field. In particular, they
provide detailed records of full-scale bow wave breakup
into spray droplets.

(a) Discard objects with an area less than four pixels.


(b) Discard objects with an area greater than 25%
of total image area.
p
5. Calculate the apparent diameter, 2 area/, for the
remaining set of objects.
6. Track the objects from one image to the subsequent
image.
(a) The centroid of the object in the subsequent
image must be located within a certain neighborhood around the centroid of object in the
original image. The size of the neighborhood
is determined by averaging the magnitude of
the velocity of several object pairs selected by
the user for each group of fifty images.

The primary mechanism of bow spray droplet


generation is the breakup of the bow sheet. In the highspeed video, the distortion of the bow sheet into ligaments that elongate until the tip pinches off to form a
droplet was observed. This type of breakup has been described in detail by Sarpkaya and Merrill (1999). The
high-speed video also showed instances where the bow
sheet became extremely thin and appeared to shatter.
First, holes form which perforate the sheet. The holes
grow rapidly leaving threads at the edges which then
break up into drops. This seems similar to descriptions
given by Dai et al. (1997). Both types of bow sheet
breakup are seen in the images presented in Figure 4. In
windy conditions, the airflow assists bow sheet breakup,
i.e. atomization.

(b) The object in the subsequent image must have


an area within 10% of the object area in the
original area.
(c) Any non-unique object pairs are discarded.
7. Calculate velocity by dividing distance travelled by
the time between two successive video images.
Error Analysis

The secondary mechanism of bow spray generation is the impact of the bow wave jet on the free surface, and the amount of spray droplet generation (and
air entrainment) depend quite critically on the degree of
breakup of the bow wave jet before impact. The free
surface itself is disturbed and may be rough and foamy.
After the primary impact, there may be additional impacts similar to a skipping stone, but these produce fewer
droplets. Finally, droplets may also be produced by
droplet splitting and by bubbles bursting at the free surface.

The greatest source of error in this technique arises from


the depth of field uncertainty. The depth of field refers to
a range that begins in front of the working distance and
extends behind it; if an object is located in this range it
will appear in-focus in the image. An illustration is included as Figure 5. For each of the SAIC high-speed
video runs, the depth of field was determined using a
software lens calculator (MiDAS 2.0 by Xcitex, Inc.)
created from accumulated camera and lens manufacturer
literature. Then, the depth of field was used to calculate
upper and lower limits on droplet size and velocity measurements. For example, at a working distance of 6.9 m
(SAIC high-speed video runs 71 and 72), a 5.3 mm apparent diameter droplet could appear as large as 5.5 mm
if it was at the near edge of the depth of field or as small
as 5.2 mm if it was at the far edge of the depth of field.

Quantitative Results
Image processing of the high-speed videos yielded quantitative results; the following sets of figures present examples of typical results from processing a single run as
well as groups of related runs. The first set of figures,
Figures 6, 7, and 8, show typical results from processing a single run, SAIC run 72 where the ship forward
speed was 6.2 m/s (12 knots). In addition to the results
shown here, the image processing programs also calculate velocity direction, droplet eccentricity, and droplet
size-velocity maps. For this run, the camera was located
at (x, y, z) = (2.20, 6.62, 4.47) meters, with respect

RESULTS AND DISCUSSION


The image processing and analysis yielded results which
were both qualitative and quantitative in nature; this section presents a summary. In addition, the high-speed
video was correlated with the ship motion data using
GPS time code; these results are also presented here.
5

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

3.5
3
1000*probability function

350

number of droplets per image

300
250
200
150

4
5
time (s)

1.5

0.5

300 400 500 600 700


apparent droplet velocity (cm/s)

800

900

1000

Figure 6 presents the number of droplets per image as a function of time and demonstrates the unsteady
nature of the bow spray. Figure 7 shows the numerical
frequency (probability density function) of the apparent
spray droplet diameter. The number of droplets is normalized such that the area under the curve connecting
the data points is equal to one. Recall from the Image
Processing section
that the apparent droplet diameter is
p
defined as 2 area/. A total of 148,410 droplets were
counted. The bin size for these distributions was 1 mm;
for example, the fraction of spray droplets with an apparent diameter between 2 and 3 mm is represented by the
filled circle at 2.5 mm on the abscissa. The mean spray
droplet apparent diameter for this run was 5.3 mm. Based
on spray droplet distributions reported in the literature
(see, for example, Sarpkaya and Merrill (2001) and Sallam et al. (1999)), a more Gaussian-shaped distribution
was expected. The shape of these size distributions indicate that there are probably a number of smaller droplets
that could not be captured with a high-speed video camera having a resolution of 512 by 480 pixels.

1
1.5
apparent droplet diameter (cm)

200

to an origin at the forward perpendicular on the design


waterline. A three-dimensional model of the high-speed
camera boom and the R/V Roger Revelle hull was used
to determine this position from measurements of the control line lengths. The R/V Roger Revelle stem is visible
in the video from this run.

2.5

0.5

100

Figure 8: A typical bow spray number frequency distribution


of the spray droplet velocity for SAIC run 72. The ship forward speed was 6.2 m/s (12 knots). The mean value of this
distribution is 2.7 m/s and the standard deviation is 1.4 m/s.

Figure 6: Number of bow spray droplets identified per image


as a function of time for SAIC run 72. The ship forward speed
was 6.2 m/s (12 knots). The sampling rate was 250 Hz. The
dotted line indicates the mean value.

probability function

1.5

0
0

50

0
0

0.5

100

0
0

2.5

2.5

Figure 7: A typical bow spray numerical frequency distribution of the apparent spray droplet diameter for SAIC run 72.
The ship forward speed was 6.2 m/s (12 knots). The mean value
of this distribution is 5.3 mm.

Figure 8 shows the numerical frequency (probability density function) of the spray droplet velocity for
the same run as is presented in Figure 7. It is extremely
important to note that the velocities reported here were
measured in the focal plane of the camera. Although the
spray droplets also move in and out of the focal plane,
6

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

locations along the wave. Figure 10 illustrates a general


decrease in droplet apparent diameter as the distance aft
of the forward perpendicular increases.

mean apparent droplet diameter (mm)

6
5.5

RUN 72
RUN 71

Correlation of High-Speed Video and Ship Motion


Data

4.5
RUN 79
RUN 77

Ship motion data was collected throughout the R/V


Roger Revelle field experiment by a motion reference
unit (Seatex MRU 5). A small portion of the ship motion
data which corresponded to a set of high-quality SAIC
high-speed video camera runs was provided by SIO.
The motion package included three angular rate sensors
and three linear accelerometers that were strapped-down
(as opposed to gyro-stabilized) to the platform frame.
As a result, measurements were made in the platforms
frame of reference. Transformation of the measurements
to the earths frame of reference is involved, but welldocumented in the literature (see, for example, Thwaites
(1995) and Edson et al. (1998)).

RUN 78

RUN 75
RUN 74

3.5

RUN 80

RUN 73
RUN 76

3
RUN 81

2.5
0

3
4
distance aft of FP (m)

Figure 10: Mean spray droplet apparent diameter as a function of distance aft of the forward perpendicular (FP) for runs
71-81 (black). The average diameter for all of the runs at a
given location is also shown (red). The error bars on the mean
values indicate upper and lower bounds based on spray droplet
position within the image depth of field.

Our initial hypothesis was that the bow spray


would correlate with the total vertical displacement of
the bow in an earth-fixed coordinate system. A procedure outlined by Edson et al. (1998) was used to calculate the total vertical displacement at the bow; however,
it was found that the result was highly dependent on the
time constant chosen for the high-pass filters. These filters could not be eliminated from the procedure because
there was an obviously non-physical low frequency component with an amplitude of greater than 10 m. It was
possible to select a time constant for the high pass filter to remove this low frequency component; however,
the filter may also have removed physically meaningful
frequency components. In addition, video observations
suggested that the motion of the ship relative to the incoming waves was important in determining the extent
of the spray sheet, but incoming wave height was not
measured. Because the total vertical displacement of the
bow could not be calculated with confidence and the observations suggested that the relative motion between the
ship and the waves may be better correlated with the bow
spray events, the correlation results presented here use
the pitch angle data in a ship-fixed coordinate frame.

this movement could not be measured with the existing


high-speed video system. The distribution has a mean
value of 2.7 m/s and standard deviation of 1.4 m/s.
Figure 9, present mean spray droplet size and
velocity as functions of Froude number
based on ship
draft, F rd . It is defined as F rd = U/ gd where d is the
ship draft. The runs selected for these figures were from
similar locations, each field of view was within 1.2 meters of the bow stem. Sixteen runs are included in these
figures: runs 69 and 70 at 3.1 m/s (6 knots), runs 60,
61, 62, 64 and 66 at 4.1 m/s (8 knots), runs 47 and 48
at 5.1 m/s (10 knots), runs 71, 72 and 75 at 6.2 m/s
(12 knots), runs 53 and 54 at 7.2 m/s (14 knots), and
runs 56 and 57 at 7.7 m/s (15 knots). Table 1 shows the
effects of the depth of field for the working distances employed by the runs included in these figures. The mean
spray droplet apparent diameters and velocities were averaged for all of the runs at a given Froude number. The
figures show a decrease in the mean spray droplet diameter and an increase in the mean spray droplet velocity
with Froude number; linear curve fits were added to the
figures to emphasize these expected trends. At higher
Froude numbers the breakup of the bow wave is more
energetic and produces droplets that are smaller and have
greater velocities.

The correlation results are summarized in Figure 11 for SAIC runs 7779. Figure 11 show the number
of droplets identified in each high-speed video image as a
function of time; the high-speed video runs begin at zero
seconds. In addition, about sixty seconds of pitch angle
data is also shown where positive pitch angle values indicate bow down motion. Finally, a sixty second interval
of the bow spray elevation as measured from the tuffcam video was added. Each frame from this video was
stamped with the GPS time code and was used to synchronize the measurements. Both the pitch angle and the
spray elevation data is periodic; furthermore, bow down

Figure 10 demonstrates the variation of mean


droplet size with location along the bow wave. Eleven
runs, Runs 71 through 81, are included in this figure.
Each was performed at a ship forward speed of 6.2 m/s
(12 knots), yet the field of view was located at successive
7

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Mean Diameter (mm)

6
5.5
5
4.5
4
0.4

0.5

0.6

0.7

0.5

0.6

0.7

0.8

0.9

1.1

0.8

0.9

1.1

Mean Velocity (m/s)

4
3
2
1
0
0.4

Frd

Figure 9: Mean spray droplet apparent diameter (top) and velocity (bottom) as a function of Froude number based on ship draft,
F rd . In both figures, the error bars on the mean values indicate upper and lower bounds based on spray droplet position within the
image depth of field. Linear curve fits are shown by the dotted lines.

run
47, 48, 53
54, 56, 57
60-62, 64, 66, 69-72
75

working distance
(m)
5.94
5.21
6.91
3.84

near edge
(m)
5.79
5.10
6.71
3.77

far edge
(m)
6.10
5.32
7.12
3.90

depth of field
(cm)
31
22
41
13

max.
(mm)
5.23
5.06
5.24
5.15

min.
(mm)
4.97
4.84
4.94
4.98

Table 1: Spray droplet size error bounds based on location within the depth of field for a given working distance.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

motions seem to correlate well with the increases in bow


spray elevation. Cross-correlation results for pitch angle and spray elevation for SAIC runs 7779 are shown
in Figure 12. The cross-correlations appear quite structured, the local maxima occur at regular intervals. In fact,
all of the correlations have eleven local maxima over a
sixty second interval which indicates that the two signals
have a common frequency of 0.2 Hz. The absolute maximum occurs between 0.5 and 1.0 seconds in the figures;
this slight apparent phase difference in the two signals is
likely due to slight synchronization errors.

full-scale; the main conclusions in this area follow.


The primary mechanism of bow spray generation
appears to be the breakup of the bow sheet. Breakup
occurs as the bow sheet: (1) distorts into ligaments
that elongate until the tip pinches off to form a
droplet and/or (2) thins until it ruptures or shatters
into myriad droplets
Increases in the ship Froude number cause the mean
spray droplet apparent diameter to decrease and the
mean spray droplet velocity to increase because the
breakup of the bow wave into spray droplets is more
energetic.

CONCLUSIONS

Increases in the distance aft of the forward perpendicular along the bow wave appear to cause the
mean spray droplet apparent diameter to decrease.

A high-speed video camera was deployed on the 2002


R/V Roger Revelle field experiment to record images of
the bow spray; a multitude of high-quality images were
obtained. These images were processed using custom
software yielding both qualitative and quantitative results. The conclusions from this experimental work fall
into two main areas. The first area is an assessment of the
high-speed video measurement technique and the second
area is a description of the bow spray at full-scale.

The ship motion relative to the incoming waves is a


key parameter in determining the height of the spray
sheet. Since incoming wave data was not available
for this experiment, correlations between pitch angle in ship fixed coordinate system and the spray
sheet elevation were used instead.
It is important to note that these results are from an initial
experiment employing a unique research technique, and
they should be interpreted and used with care.

High-Speed Video Measurement Technique


To the authors knowledge, these were the first attempts to use high-speed video measurement techniques
to study full-scale bow flows in the field. At the time of
the experiment, the best commercial off-the-shelf highspeed video camera was purchased. This camera was
deployed successfully in the field and recorded highquality (excellent contrast and focus) images. Furthermore, spray droplet velocity distributions appear reasonable when compared with the ship forward speed; therefore, frame rate, field of view, and other camera parameters were appropriately selected. However, a camera
with higher resolution and additional memory is strongly
recommended for future experiments. Also, simultaneous recording of GPS time code with the high-speed
video images is recommended to facilitate synchronization with other shipboard measurement systems. In addition, a motorized lens to allow electronic control and
readout of zoom, focus, and aperture is recommended.
An electronically controlled pan and tilt would also improve the ease of use of this system in the field. Moreover, it would significantly improve the repeatability of
different camera positions and views.

FUTURE RESEARCH
Future directions for this work include both field experiments and numerical modelling efforts.

Field Experiments
Additional field experiments were conducted on the R/V
Athena in Panama City, FL in October 2003 and are
planned for May 2004. Table 2 presents a comparison of geometric characteristics for a DDG-51, the R/V
Roger Revelle, and the R/V Athena. High-speed video
of the bow spray was recorded. These experiments provided an opportunity to collect calm water data for a
full-scale naval combatant and to improve the high-speed
video system according to the suggestions described in
the Conclusions section. The calm water conditions are
important for several reasons. First, experimental measurements are more repeatable in calm, protected water than in the open ocean. Second, current prediction
methodologies for breaking waves (computational fluid
dynamics and scaled model-scale data) are limited to
steady conditions; therefore, calm water full-scale data
is required for their validation and further development.

Bow Spray
This experimental work provides a detailed description
of the breakup of the bow wave into spray droplets at
9

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Length (m)
Beam (m)
Length-beam ratio
Block coefficient
Prismatic coefficient
Bow half angle (deg.)
Flare angle (deg.)

DDG-51
142 (waterline)
18.0 (waterline)
7.9
0.521
0.627
10
10-20

R/V Roger Revelle


83.2
16.0
5.2
0.545
0.608
19 (at waterline)
26

R/V Athena
47.0
7.0
6.9
0.47
0.641
10
10-25

Table 2: Comparison of ship characteristics.

picted in Figure 13. At the time this model was developed, suitable experimental data was not available for its
validation.
There are some numerical models documented
in the literature which are related to the bow spray problem; however, they cannot be directly applied. Modelling the generation and evolution of large quantities of
droplets over a wide range of scales requires specialized
procedures. For example, the approach that Sussman and
Dommermuth (2001) used for studying spray sheets is
only suitable for micro-scale analysis of droplets; it is
not possible to model the flow around an entire ship with
such detail. Numerical models for spray droplet breakup
computations such as the Taylor Analogy Breakup model
(ORourke and Amsden, 1987) and the surface wave
instability model (Reitz, 1987) are also too detailed.
Complementary to the numerical simulations are simpler approaches. Sellens and Brzustowski (1985) used
the maximum entropy formalism to predict the drop
size distribution in a spray resulting from the breakup of
a liquid sheet, but the average droplet size must already
be known. Novikov and Dommermuth (1997) presented
a statistical description of droplets in a turbulent spray
connected with the description of turbulent dissipation.
Formulas for characteristic droplet sizes and corresponding probability distributions are obtained; however, these
contain unknown constants that can only be determined
empirically.

Figure 13: Bow wave (blue) of surface ship model 5415 (gray)
with envelope of spray sheet (white) as calculated by previous
SAIC bow spray model.

Numerical Modelling
As a result of this experimental work, there is a unique
collection of bow spray measurements for different geometric and flow conditions that will be helpful in bow
spray model development and validation. The development of an empirical spray model that will interface with
an existing base flow code has been initiated by Dr. Sur
at SAIC. Several years ago, a bow spray model was developed at SAIC as part of an ongoing effort in computational ship hydrodynamics. It used the bow wave
predicted by Numerical Flow Analysis (NFA), a mature
Large-Eddy Simulation (LES) code written by Dr. Douglas Dommermuth of SAIC, for surface ship model 5415.
Spray source points were located along the contact line
of this bow wave along the ship and then along the cusp
of the overturning bow wave. A size distribution similar
to that presented by Novikov and Dommermuth (1997)
was assigned to the spray droplets and they were ejected
from the source points with velocities related to the base
flow velocity. The spray droplets were tracked as they
interacted with an approximate solution for the air flow
around the ship. The results of the calculation are de-

ACKNOWLEDGEMENTS
This research is supported by the Office of Naval
Research under contract numbers N00014-97-C-0345,
N00014-01-C-0295, N00014-02-C-0283, and N0001403-0105. Dr. L. Patrick Purtell is the program manager.
The authors also wish to acknowledge Mr. John Kuhn
of SAIC, Dr. Eric Terrill , Ms. Lisa Lelli, and Professor
Ken Melville of the Scripps Institution of Oceanography,
and Dr. Thomas Fu, Mr. Billy Boston, and Mr. Martin
Sheehan all of the Naval Surface Warfare Center, Carderock Division for their contributions to this work.
10

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

REFERENCES

simulation of ship waves using cartesian-grid methods,


Proceedings of the Twenty-Third Symposium on Naval
Hydrodynamics, Office of Naval Research, 2001,
pp. 762779.

Dai, Z., Hsiang, L.-P., and Faeth, G., Spray Formation at the
Free Surface of Turbulent Bow Sheets, Proceedings of
the Twenty-First Symposium on Naval Hydrodynamics,
Office of Naval Research, 1997, pp. 490505.

Thwaites, F.T., Development of an Acoustic Vorticity Meter


to Measure Shear in Ocean-Boundary Layers, Doctoral
dissertation, 1995, Massachusetts Institute of
Technology, Woods Hole Oceanographic Institution.

Edson, J.B., Hinton, A.A, Prada, K.E., Hare, J.E., and Fairall,
C.E., Direct Covariance Flux Estimates from Mobile
Platforms at Sea, American Meteorological Society,
1998, pp. 547562.
Kaiser, J.A.C, Garrett, W.D., Ramberg, S.E., Peltzer, R.D., and
Andrews, M.D., WAKEX 86; A Ship Wake/Films
Exploratory Experiment, NRL Memorandum Report
6270, 1988, Naval Research Laboratory.
ORourke, P. J. and Amsden, A. A., The TAB Method for
Numerical Calculation of Spray Droplet Breakup,
Society of Automotive Engineers, paper number 872089,
1987.
Novikov, E.A. and Dommermuth, D.G., Distribution of
droplets in a turbulent spray, Phys. Rev. E, Vol. 56,
No. 5, 1997, pp. 54795482.
Ratcliffe, T., Boston, W.S. and Sheehan, M., Photographic
Visualization of the Surface Wave Field and Boundary
Layer Surrounding the Research Vessel, R/V
REVELLE, NSWCCD-50-TR-2003/003, March 2003,
Naval Surface Warfare Center, Carderock Division, West
Bethesda, MD.
Reitz, R.D., Modelling Atomization Processes in
High-Pressure Vaporizing Sprays,
Atomization and Sprays Tech., Vol. 3, 1987, pp.
309337.
Sallam, K.A., Dai, Z., and Faeth, G. M., Breakup of
Turbulent Liquid Jets in Still Gases,
Proc. AIAA Fluid Dynamics Conference, AIAA paper
number 99-3759, June 1999.
Sarpkaya, T., and Merrill, C.F., Spray formation at the free
surface of liquid wall jets, Proceedings of the TwentySecond Symposium on Naval Hydrodynamics, Office of
Naval Research, 1999, pp. 796808.
Sellens, R.W. and Brzustowski, T.A., A Prediction of the
Drop Size Distribution in a Spray from First Principles,
Atomization and Spray Technology, Vol. 1, 1985,
pp. 89-102.
Shroff, S. and Liepmann, D., Spray Generation over Curved
Surfaces, In Proceedings of the Fluids Engineering
Division Summer Meeting, American Society of
Mechanical Engineers, 1997.
Sirignano, W. Fluid Dynamics and Transport of Droplets and
Sprays, Cambridge University Press, 1999.
Sussman, M. and Dommermuth, D.G., The numerical

11

Copyright National Academy of Sciences. All rights reserved.

100

20

10

10

20

2
40

30

100

50

0
30

20

10

10

20

2
40

30

100

50

0
30

20

10

10

20

pitch (deg)

0
30

pitch (deg)

200

pitch (deg)

number of droplets

number of droplets

number of droplets

Twenty-Fifth Symposium on Naval Hydrodynamics

5
40

30

time (s)

Figure 11: SAIC high-speed video run 77 (top), 78 (middle) and 79 (bottom). The blue line indicates the number of droplets per
high-speed video image; the high-speed video runs begin at time zero seconds. The green line indicates the pitch angle where
a positive value represents bow down motion. Finally, the red line indicates the elevation of the spray sheet measured from the
tuffcam video.

amplitude

6000
4000
2000
0
30

20

10

10

20

30

20

10

10

20

30

20

10

10

20

30

amplitude

5000
4000
3000
2000
1000
30

amplitude

6000
5000
4000
3000
2000
30

time (s)

Figure 12: Cross-correlations of the pitch angle (where a positive value represents bow down motion) and the bow spray elevation
for SAIC high-speed video run 77 (top), 78 (middle) and 79 (bottom).

12

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Alan Brandt
Johns Hopkins University, USA
A significant fraction of the bow sheet does
not form droplets. Is it not important to also account
for this in the measurements at model-scale?
AUTHORS REPLY
In the field tests on the R/V Revelle and R/V
Athena and also in the model scale (i.e. bow wedge,
see Karion et al., 2004) tests, there is a portion of the
bow sheet that does not break up into droplets that is
visible in the high-speed video camera field of view.
In each high-speed video image, this portion of the
bow sheet is filtered out so that it does not register as
a large droplet and affect the droplet size
distributions. It would be of interest to quantitatively
characterize this portion of the bow sheet,
particularly the surface roughnesses; however, this
has not been a focus of our work.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
DISCUSSION
Thomas C. Fu
Naval Surface Warfare Center, Carderock Division,
USA
1.

Your measurement of spray droplet velocity is


restricted to the focal plane of the camera, and
you intentionally minimize your depth of field to
reduce error. While the bow sheet is relatively
thin, and moving primarily along the hull, one
can see ligaments and droplets that appear and
move through the imaged plane. How big a
limitation do you think this is in your present
correlation with the ship motion, and do you
think this will be an issue as you construct your
empirical model, both in developing the model
and its application?

AUTHORS REPLY
The correlation with ship motion focused on
the relationship between the maximum elevations of
the spray sheet with pitch angle. In the model, the
initial ejection velocity for the spray droplets will be
provided by the base flow code. If the full threedimensional velocity of the spray droplets could be
measured (perhaps by using multiple synched highspeed video cameras), then these measurements could
be used to validate the velocity prediction by the base
flow code.

Thomas C. Fu
Naval Surface Warfare Center, Carderock Division,
USA
3.

You have recently (October 2003 and May 2004)


taken data on the R/V Athena. Knowing that you
have not had a great deal of time to look at that
data have you noticed any obvious differences
between the R/V Revelle and R/V Athena data? If
there are differences do you think these are due
to the differences in geometry, ambient
conditions, ship motion, or something else?

AUTHORS REPLY
Preliminary image processing of the R/V
Athena data revealed mean droplet sizes that were
smaller than those from the R/V Revelle data. This
difference is likely due to the deployment of a higher
resolution high-speed video camera on the R/V
Athena. General trends such as mean droplet size
decreasing with ship forward speed and mean droplet
velocity increasing with ship forward speed appear to
be consistent, though additional image processing of
the R/V Athena is required.
DISCUSSION

DISCUSSION

Thomas C. Fu
Naval Surface Warfare Center, Carderock Division,
USA

Thomas C. Fu
Naval Surface Warfare Center, Carderock Division,
USA

4.

2.

Presently you are working on an empirical spray


model, do you have plans to work on a more
physics based model, and what kind of
data/effort would be required to do this?

One of your main conclusions is that: Increases


in the distance aft of the forward perpendicular
along the bow wave appear to cause the mean
spray droplet apparent diameter to decrease.
Do you think this is due to a more energetic
breakup of the bow sheet or is there the
possibility that smaller droplets generated further
forward are being carried aft?

AUTHORS REPLY
AUTHORS REPLY
The spray model is actually semiempirical; the droplet size distribution is based on a
theoretical droplet size distribution by a turbulent
breakup mechanism (Novikov and Dommermuth,
1997). It is not yet possible to run a physics-based
calculation like Dommermuths (1999) twodimensional spray sheet breakup due to grid
resolution limitations.

Our previous modeling efforts have shown


that smaller droplets generated further forward are
being carried aft by the air flow. The smaller
droplets are lighter and remain in the air longer;
therefore, they are carried further aft by the air flow.
This causes the mean spray droplet apparent diameter
to decrease with distance aft of the forward
perpendicular.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

04-05-13
25th Symposium on Naval Hydrodynamics
St. Johns, Newfoundland and Labrador, CANADA, 8-13 August 2004

Computational Design of Trans-Cavitating Propellers


and Experimental Evaluation of Their Performance
Yoshitaka UKON1, Tatsuro KUDO1,
Junichi FUJISAWA1, Noriyuki SASAKI 2
(1 National Maritime Research Institute, 2Sumitomo Heavy Industry, Japan)
fluctuations as that of the TCPs, while for the 35.0kt
case, some TCPs and one designed CP working behind
a ship hull showed better performance not only for the
efficiency but also for the pressure fluctuations than
the traditional CPs.
This paper concludes that a TCP is promising for
a highly loaded propeller working behind a high-speed
ship with shallow draft and presents future tasks to
enhance the performance of TCPs and to improve the
evaluation on cavitation performance of propellers
working in non-uniform flows under TC condition.

ABSTRACT
This paper presents the results of a research
project on the theoretical design of trans-cavitating
propellers (TCP) for high-speed and high-powered
ships with shallow draft.
Extensive experimental
evaluation was carried out on these propeller models
working in a uniform flow and behind a complete ship
model of twin-screw large fast ferry. Ten propellers
were designed to generate the necessary thrust under
the operating condition between sub-cavitation and
super-cavitation.
First of all, this paper describes a theoretical
design method for TCPs, developed from the NMRI
super-cavitating propeller design method. The present
method employs two kinds of combinations with
super-cavitating and non-cavitating blade sections to
design a hybrid propeller. In this paper, six propellers
including conventional propellers were designed at
30.6kt and four propellers were designed at 35.0kt
under each normal operating rate condition. In each
case, the design of the propellers was made not only by
current methods but also by a newly developed method
which can theoretically control sheet cavitation over
the propeller blades using a couple of high
performance blade sections, so as to satisfy a given
thrust with maximum efficiency. Extensive evaluation
in cavitation open water tests shows that the designed
TCPs had higher efficiency than the CPs under the
severer cavitating condition.
Secondly, this paper introduces a method to
evaluate the propeller efficiency working behind a ship
model under trans-cavitating (TC) conditions for a
given propeller, thrust load coefficient. Using this
evaluation method, extensive experimental evaluation
of TCPs was made in the NMRI (SRI) Large
Cavitation Tunnel with respect to the propeller thrust,
propeller efficiency, pressure fluctuations and erosion.
For the 30.6kt case, conventional propellers still kept
in the same level of efficiency and pressure

INTRODUCTION
In these days, high speed ships, for example,
Ro/Ro ferries, become faster and bigger. Since high
speed and/or highly loaded propellers cannot evade
cavitation, a new type of propeller is needed, which
can generate the necessary thrust under the operating
condition between sub-cavitation and super-cavitation.
Under this condition, most of the suction side of the
propeller blades is covered with cavitation and
significant thrust breakdown is revealed, if a subcavitating propeller (Non-Cavitating Propeller, CP)
with aerofoil (NC) sections is used under this
condition. Most of the propeller blade (near the tip) is
super-cavitated and the rest of the blade (near the root)
is partially cavitated. This is called the transCavitating (TC) condition. Under this condition the
efficiency of this type of a propeller becomes
tremendously worse. Propeller designers have to give
up their propeller design by a normal design method.
Then, the trans-cavitating propeller (TCP) with a
hybrid concept is expected to be one of the possible
candidates
In order to design a highly efficient propeller
under such a condition, Yim invented the TCP (Yim,
1998, Vorus, 1988). The propeller blade is divided
into two or three domains by the borderline or the
intermediate region (domain C) as shown in Fig. 1.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

04-05-13

geometrical shape was modified by their lifting surface


theories similar to the NMRI SCP design method
(Ukon, 1994 and 1995) under the same design
condition as that performed by the NMRI (Ukon,
1994), that is, CT=0.334 (KT=0.1587, J=1.10) and
V_SC=0.4. To design a TCP, two-term SC sections for
SC domain (B) and NACA 66 sections for NC domain
(A) were employed. The measured thrust coefficient
and efficiency were 0.183 and 0.676 at the designed
advance coefficient, while the previous design for a
hybrid propeller by the NMRI obtained the higher
efficiency of 0.695 in the experiment (Ukon, 1995,
0.786 in Wangs paper is misreading). The reasons
to design a lower efficiency TCP might be suggested
that the comparison was made at higher loaded
condition (CT=0.385) than the design one and less liftdrag ratio of two-term SC sections than the SRJN
section were employed.
Three cases described above are the application
of TCPs to lightly loaded propellers. This paper
discusses the application of TCPs to a more highly
loaded case and experimental and design issues for
TCPs.

Super-Cavitating (SC) blade sections are adopted near


the tip (domain B), while aerofoil sections are
employed near the root (domain A). In the domain C,
blade sections are smoothly interpolated by using the
respective offset data given by two different kinds of
blade sections with the help of suitable polynomials.

Fig. 1 Concept of Trans-Cavitating Propeller


Vorus et al. made the first development of a TCP
and the TCP was designed for a 40kt-25m high-speed
patrol boat (Vorus et al., 1988). The pitch and blade
sections of the four bladed trans-cavitating propeller,
called hybrid propeller, change at the borderline
(0.75R) with kinks on the face blade in the radial
direction but the back blade was continuous and
smooth. The propeller was designed under the
condition that the advance coefficient J, the thrust
coefficient KT and the cavitation number v_SC were
1.410, 0.154 and 0.409 (n=0.813) respectively. The
expected propeller efficiency of 0.731 was obtained
under the very lightly loaded design condition of
CT=(8/)(KT /J2)=0.197. The model test results in
two cavitation tunnels showed that this TCP has the
similar high performance as a Newton-Rader propeller.
The latter propeller is well known as a high
performance propeller but causes severe vibration due
to cavitation self-oscillation under a certain operating
condition (Ukon, 2001).
Yim et al. designed a four bladed TCP for a
34kt-23m air cushion vehicle (Yim, 1998). The
design condition was given by one that the advance
coefficient J was 1.048, the thrust coefficient KT was
0.164, and the cavitation numberV_0.7R based on the
ship speed and the static pressure at 0.7R in the
upright position was 0.65. Cavitation tests on this
propeller were performed in the KRISO cavitation
tunnel. The expected propeller efficiency O was
0.62 but the obtained efficiency was 0.57 under the
design condition of CT=0.380 (KT=0.164, J=1.048)
andV_0.7R = 0.65.
Wang et al. (Wang, 2001) designed a TCP by
their lifting line theory for preliminary design and its

DESIGN METHOD OF TCP


Assumption in Present Design
This paper describes the design method (Kudo,
1999c) on the geometrical blade shape of TCPs under
the assumption that the principal particulars, such as
propeller diameter, blade number, boss ratio and the
parameters prescribing the working condition, such as
advance coefficients, thrust coefficient and cavitation
number are given. Some of the parameters on the
propeller working condition can be determined by a
parametric study and with the assistance of propeller
design charts.
In the vicinity of propeller tip, the optimum SC
blade sections of a TCP are aggressively designed by
the NMRI super-cavitating propeller design method
(Ukon, 1994) as possible, to avoid the cavity collapse
on the blades. On the other hand, a traditional design
method is applied for the NC blade sections near the
root of the TCPs. In the intermediate region C, two
kinds of blade sections generated by different concepts
as shown in Fig. 1 are connected smoothly using
polynomials.
Employed Blade Section
In this paper, two combinations with four kinds
of blade sections are employed for the design of TCPs
as shown in Table 1.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

04-05-13
sections their blade width is determined as large as
possible to get the optimum performance.

Table 1 Combination of Blade Sections


Near Root

Near Tip

NC Section

SC Section

Combination I

NACA 16

SRJN

Combination II

UT-NC

UT-SC

Two-Dimensional Blade Load Distribution


The load distributions on the blade section at
each radial position of TCPs are determined by the
chordwise pressure distribution on optimum twodimensional blade sections. The load distribution of
the SRJN sections can be computed by the linear
vortex panel theory, while that of NACA sections can
be given by one of the NACA camber lines, such as
a=0.8 mod. The load distributions for UT-SC and UTNC are calculated from the pressure distribution for
each blade section.

An attempt was made to apply the combination I


for the design of SCPs and a relatively good result was
obtained (Ukon, 1995). The SRJN SC section used
near the tip was developed for high performance SCPs
using the face blade shape based on Johnson five terms
and the back blade shape given by the linear vortex
panel method with thickness at the trailing edge
(Ukon, 1994, 1995, 2000b and Kudo, 2000).
The back shape of the SC section is determined
by the required lift coefficient and strength at a given
local cavitation number R and thrust coefficient KT.
Near the root, a NACA 16 series blade section is
employed because the leading edge radius is so small
to connect to the SC section smoothly. In the
combination II the UT-SC section was developed for
an optimum TC section at lower cavitation number
and higher lift coefficient with zero thickness at the
trailing edge, to connect this section to the NC
sections easily. The NC section was developed using
Epplers method to design an optimum pressure
distribution with high lift drag ratio and wider
cavitation bucket zone (Yamaguchi, 1999).

Three-Dimensional Blade Section


Since the required three-dimensional load
distributions at each radial position are known, the
final geometrical lifting surface can be determined by
the lifting surface correction using a lifting surface
theory as the design of the NCPs and the SCPs. Based
on the obtained three-dimensional camber lines, threedimensional blade sections can be obtained by
superposing these camber lines and two-dimensional
blade thickness or cavity thickness.
Fairing at Intermediate Zone
In the design of TCPs, the fairing is needed at the
intermediate zone C between the SC domain A and the
NC domain B as shown in Fig. 1. In this paper, one
or two radial points from each of two end regions are
selected for the fairing and the offsets of blade sections
are interpolated by quadratic or cubic polynomials.

Optimum Circulation Distribution


The present design of TCP utilized the Lerbs
optimum circulation distribution as used for the design
at NCPs and SCPs as a target. The circulation
distribution was computed by the Lerbs lifting line
theory. It is difficult to satisfy the target circulation
distribution especially in the intermediate region using
two different kinds of optimum blade sections.

Interactive Employment of Lifting Surface Theory


The present design of TCPs is performed with the
help of the improved lifting surface theory SC-VLM 3
interactively to confirm the design result on thrust and
propeller efficiency and especially the effects of the
interpolation at the intermediate zone. The SC-VLM3
was improved from the SC-VLM used for the design
of SCPs (Kudo, 1994) at several points and developed
for the application for the TCPs.
The improved points are given as follows,
* Increase in the patterns of panel from two (NC and
fully cavitated) to three (partly cavitated added) to
avoid sharp and stepwise change in the loading
between neighboring panels of the TCP
* Definition of the rear end position of cavity in the
last panel to satisfy the cavity closure condition
* Introduction of a nonlinear term on the pressure to
calculate cavity shape more rigorously

Determination of Blade Contour


The present TCPs are assumed to be used for
controllable pitch propellers and not only their
expanded area ratio is restricted to be less than 0.64
but also the blade width and skew line are constrained
not to hit each other, when the pitch of blades is
changing. In the present design of TCPs, the blade
width at each radial position is determined not only to
satisfy the required section modules but also to obtain
the maximum lift drag ratio iteratively. The blade
contour is determined in different ways for the
respective SC blade sections. For the SRJN sections,
their blade width is given as small as possible, to
obtain the highest lift drag ratio, while for UT-SC

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

04-05-13

the conventional propellers (CPs) and the transcavitating propellers (TCPs) are called CP and HP,
respectively. Usually the thrust load coefficient art
35.0 kt should be higher than that at 30.6 kt. In order
to clearly demonstrate the effect of cavitation number
on the design results, however, the thrust load
coefficient was kept to 0.728 throughout the present
propeller design. Except the conventional propeller of
CP-1, other propellers were designed as the propeller
working in the starboard side.
All of the propellers were designed for those
turning outwards, based on the self-propulsion tests
performed at the NMRI 400m towing tank. The
designed advance coefficients for some TCPs (HP-5, 6,
7) were modified with taking account of the tangential
wake in the propeller plane.

* Introduction of an empirical modeling on the base


pressure at the trailing edge surface. The length of
separation is assumed to be one panel length to give
the pressure condition at the trailing edge for the
lifting surface equation of TCPs.
DESIGN OF CP AND TCP AND
EVALUATION IN UNIFORM FLOW

THEIR

Design Condition
The present design method was applied for the
TCPs equipped to twin-screw large high-speed ferries
with twin rudders whose design speeds are 30.6 and
35.0 kt. The principal particulars of the ship model,
NMRI M.S. No.610 are given in Table 2. The
working conditions for the designed propellers at the
normal operating rate (NCR) are given in Table 3.

Design and Evaluation on VS=30.6kt Case


The design results are usually evaluated by the
experiment. In this paper, two kinds of experiment
were conducted in the NMRI Large Cavitation Tunnel.
One of the tests was made for the open water test
under non-cavitating and several cavitating conditions
in the No.1 working section. Another was performed
behind the complete ship model in the No.2 working
section under the NOR and MCR conditions.

Table 2 Principal Particulars of Model


Description
Length between Perpendiculars
Length at Dead Water Line
Breadth
Depth for Cavitation Tunnel
Draft
Block Coefficient
Prismatic Coefficient

key
LPP
LDWL
B
DCAV
d
CB
CP

unit
m
m
m
m
m
-

7.000
7.233
1.089
0.389
0.272
0.500
0.556

CP-1
This propeller as shown in Fig. 2 was designed
using the existing camber lines and pitch distribution
of which controllable pitch propeller offered good
results (Toyama, 1996).
The blade thickness
distribution was determined with satisfying the
requirement of the NK rule. At the design condition
in a uniform flow, this propeller was fully cavitated
with thrust breakdown as shown in Fig. 2 but the
propeller efficiency is in a reasonable level.

Table 3 Propeller Working Condition for Design


VS [knots]
V
J
KT
CT

CP-1~3
HP-4
HP1~3
30.6
1.371
0.916
0.240

HP-5

HP-6

HP-7

0.732
0.153

0.935
0.250

35.0
1.048
0.935
0.250
0.728

The principal particulars of the designed


propeller models are given in Table 4. In this paper,

Table 4 Principal Particulars of Propeller Models


Prop. Name

CP-1

CP-2

M.P.No. (NMRI)

411/
412

413

Diameter [mm]
Boss Ratio
Pitch Ratio
Exp. Area Ratio
Proj. Area Ratio
Rake at Tip [mm]
Skew at Tip
Blade Number
Blade
Section
Rot. Direction
Material

CPBlade 3
431

HP-1

HP-2

HP-3

HP-4

HP-5

HP-6

HP-7

433

432

434

444

445

446

447

1.478
0.625
0.506

1.506
0.630
0.507

1.088
0.584
0.504

1.410
0.779
0.636

SRJN
NACA

SRJN
NACA
Right

SRJN
NACA

UT-SC
UT-NC

1.475
0.731
0.600

1.479
0.731
0.600

1.412
0.731
0.605

1.292
0.637
0.521

1.447
0.580
0.468

NACA

MAU

Mod.
NACA
R/L

UT-SC
UT-NC

SRJN
NACA

194.4
0.30
1.282
0.634
0.518
0
35.54
4
UT-SC
UT-NC

Aluminum (Anodized)

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

04-05-13

Fig. 2 Blade Shape and Cavitation on CP-1


CP-2
This propeller as shown in Fig. 3 was designed
using the MAU chart with MAU blade sections and
the same blade contour, thickness and pitch
distribution as CP-1. The pitch of this propeller was
determined to satisfy the same power and propeller
revolution rate as CP-1. This propeller was also fully
covered with super-cavitation as shown in Fig. 3 and
the serious thrust and efficiency breakdown was
observed in the cavitation open water test.

Fig. 4 Flow Chart of Design Procedure for CP-3

Fig. 5 Example of Designed Blade Section for CP-3

Fig. 3 Blade Shape and Cavitation on CP-2


CP-3
This propeller was designed at the condition of
30.6 knots by a design program based on the QCM
(Sasaki,1994). The design procedure of CP-3 is shown
in Fig.4. As shown in Fig.5, the blade section of CP-3
is optimized so as to suppress the cavitation extent by
increasing the loading in the vicinity of trailing edge.
The CPmin values on the blade surface at each radial
position are almost the same as CP-1 and the camber
distributions near the trailing edge are increased to
obtain wide shock free zone for the angle of attack.
The magnitude of camber increment near the trailing
edge depends on the wake steepness. Near the root,
the blade sections as shown in Figs.5 and 6 were
truncated to prevent root cavitation in an oblique flow
around the propeller shaft. The propeller revolution
rate and the thrust were kept the same as those of CP-1.
Propeller open water characteristics measured in
the cavitation tunnel on CP-3 are shown in Fig.6.
Under the design condition for 30.6kt case, only
bubble cavitation as shown in Fig.6 was observed and
the cavitation occurrence on CP-3 was completely
suppressed, compared to CP-1 and CP-2.

Fig. 6 Blade Shape and Cavitation on CP-3


Propeller open water characteristics on CP-3
measured in the cavitation tunnel are shown in Fig. 7.
Under the design condition for 30.6kt case, only
bubble cavitation was observed as shown in Fig. 6.
HP-1
This propeller as shown in Fig. 8 utilized UT-NC
and UT-SC in the root region and tip region,
respectively. The optimum circulation distribution
was given by the Lerbs lifting line theory. The blade
contour was determined with referring HP-2 except the
tip region. To satisfy the optimum local lift coefficient

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

04-05-13

The measured thrust coefficient at the design


advance coefficient is 0.247 and 2.9% higher than the
target value in the design. In the outside of 0.9R, no
cavitation occurs in spite of prediction of supercavitation as shown in Fig. 8. Tip vortex cavitation
was not observed.

for UT-SC section in the outside of 0.8R, the chord


lengths near the tip were expanded. As the result of
the design, the radial pitch distribution of this
propeller has two peaks as shown in Fig. 9 to get
prescribed circulation distribution in the intermediate
region as smooth as possible. Since the geometrical
fairing was not made for these propellers, wavy
geometrical offsets were generated. With regard to
HP1, by the SC-VLM3 the converged solution could
not be obtained. One of the reasons is the wavy or
zigzag blade offsets in the radial direction.

HP-2
This propeller as shown in Fig. 10 was designed
based on the Lerbs optimum circulation distribution.
The SRJN sections were used from the tip to 0.7R
where the optimum SRJN sections could be adopted,
while the NACA 16 sections were employed from 0.5R
to the root, using the pitch distribution of CP1.
Between 0.5R and 0.7R, the blade sections were
generated by the interpolation. At 0.7R, the chord
length was increased to obtain the smooth blade
contour and the thickness at the trailing edge was
reduced because of the excessive strength and smooth
radial connection at the trailing edge.
From the experiment in a uniform flow of the
cavitation tunnel, the measured thrust was 6.0%
higher than the target one in the design. The
discrepancy in the thrust, that is, over pitch is due to
the simple and direct adoption of the NC sections of
CP1. Longer sheet cavitation occurs outside 0.65R,
while partial cavitation is observed inside of 0.6R as
the expectations as shown in Fig. 10.

Fig. 7 Propeller Open Characteristics Curves


Measured in Uniform Flow at Cavitation Tunnel

Fig. 10 Blade Shape and Cavitation on HP-2

Fig. 8 Blade Shape and Cavitation on HP-1

1.6
1.4
H/DP

HP-3
This propeller as shown in Fig. 11 was designed,
referring to the open water test on HP-1 in the
cavitation tunnel. The offsets of HP-3 were generated
by fairing the geometrical shape of HP-1. The pitch
outside of 0.9R was increased to stimulate sheet
cavitation because of less sheet cavitation in this
region of HP-1. In the intermediate region, the fairing
was made to remove the wavy geometrical shape.
Although the measured thrust is 7.9% higher
than the target, the former is 2.8% higher than the
prediction by SC-VLM3 in the design. The measured

CP-1
CP-2
CP-3
HP-1
HP-2
HP-3
HP-4
HP-5
HP-6
HP-7

1.2
1.0
0.8
0.6
0.2

0.4

0.6
r/RO

0.8

1.0

Fig. 9 Pitch Distribution of Designed Propellers

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

04-05-13

around the shaft line of the dynamometer, K&R J26 as


shown in Fig. 12.

propeller efficiency of this propeller is the highest


among the designed propellers for 30.6kt case. The
super-cavitation area became wider than HP-1, while
unexpected long sheet cavity between 0.5R and 0.6R
occurred due to excessive smoothing.

HP-5
HP-5 propeller as shown in Fig. 13 was designed
by the same way in HP-4 but including the tangential
component of wake distribution at the propeller plane.
It amount to 5% of the incoming uniform flow at 0.7R
of the propeller. The design condition on the advance
and the thrust coefficients was modified and they were
increased 2.1% and 4.2%, respectively.
In order to enhance the propeller efficiency and
to design the cavity thickness at each radial position as
thin as possible, the analytical computation was
employed to determine the optimum pitch distribution
iteratively. Thus, not only the thrust but also torque
can be designed within 3% accuracy with the help of
the analytical tool. The measured thrust agrees with
the target value but sheet cavitation partly occurs at
0.7R and around the tip. The observed patterns are
different from the prediction as shown in Fig. 13
.

Fig. 11 Blade Shape and Cavitation on HP-3


Design and Evaluation on VS=35.0kt Case
The present experiments on CP-1~3 and HP-1~3
indicated that the conventional propellers, that is,
NACA type propellers aiming at suppressing the
occurrence of cavitation have still good performance at
the ship speed of 30.6kt, that is, CT=0.728 and V=
1.371. Then, the target speed of the ship was
increased to 35.0kt that is, V=1.048 but the thrust
load coefficient was not changed to demonstrate the
effect of cavitation number on the TCP design.
HP-4
This propeller as shown in Fig. 12 was designed,
using the SRJN SC sections from 0.7R to 0.95R and
NACA 16 sections from 0.3R to 0.4R.
The
interpolation between 0.4R and 0.7R was made by
cubic polynomials.

Fig. 13 Blade Shape and Cavitation Pattern on HP-5

Fig. 12 Blade Shape and Cavitation on HP-4


Fig. 14 Propeller Open Characteristics Curves Measured in
Uniform Flow at Cavitation Tunnel

The measured thrust is 2.5% higher than the


target. The observed cavitation pattern was not the
expected. The sheet cavity from 0.7R but no sheet
cavity between 0.75R and 0.95R was found. Unstable
sheet cavitation near the root occurred due to the wake

Propeller open water characteristics on HP-5 are


shown in Fig. 14. The propeller efficiency of HP-5

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

04-05-13

increases with the decrease of cavitation number and


becomes higher than that under the non-cavitating
condition as other HPs.
HP-6
HP-6 propeller as shown in Fig. 15 was designed,
increasing the propeller revolution rate by 27.2%, the
optimum circulation distribution was obtained by
Lerbs lifting line theory and 1.5% increase in the
propeller efficiency to HP-5 was predicted by this
theory. This propeller employs the SRJN SC sections
from 0.8R to 0.95R and the NACA sections from 0.3R
to 0.6R. In the intermediate region between 0.6R and
0.8R, the fairing was made by parabolic interpolation.
Both of the measured thrust and torque were 30%
and 40% higher than the target and prediction
respectively. One of the reasons is that bubble
cavitation was observed on the whole of blades as
shown in Fig. 15 in the test, while sheet cavitation was
predicted by SC-VLM3 in the design.

Fig. 16 Blade Shape and Cavitation Pattern on HP-7


CP-3
This propeller was not designed under the
designed condition for 35.0kt case but this propeller
was examined under this condition for the reference
because the design procedure and results should be the
same except the blade strength.
From the
measurements, the thrust and efficiency at the design
condition for 35.0kt case were 0.231 and 0.630, 8.0%
and 9.6% less than at that for 30.6kt case. Cavitation
patterns are shown in Fig. 17. The water repellent
stimulated cavitation drastically and reduced the thrust
and efficiency (Kudo, 1999b). The water repellent
gives the similar effects as the high Reynolds numbers
effects (Kawanami, 2000). The application of a
groove cavitator also gives the similar results (Kudo,
2001).

Fig. 15 Blade Shape and Cavitation on HP-6


HP-7
This propeller as shown in Fig. 16 used UT-SC
sections from 0.5 to 0.95R and UT-NC section only at
0.3R. Not only in the intermediate region from 0.3 to
0.5R but also at 0.7R and 0.8R, the fairing was made
because of the lack of smoothness in radial
geometrical shape.
The measured thrust of HP-7 was 5.6% and
12.3% higher than the target and the prediction by
SC-VLM3 respectively, because the SC-VLM3 in the
design predicted lower thrust than the target and due
to the intentionally increased pitch the propeller blades
were almost fully covered with sheet cavity as shown
in Fig. 16 as the theoretical prediction.

w/o (Left) and with (Right) Water-Repellent


Fig. 17 Cavitation Pattern on CP-3 in a Uniform
Flow at 35.0kt Case
Summary of Evaluation Results in a Uniform Flow
Comparisons of propeller efficiency measured in
the cavitation open water tests among the designed
propellers for 30.6kt and 35.0kt are shown in Figs. 18
and 19.
VS=30.6kt; V=1.371, J=0.916
HP-3 shows the highest propeller efficiency of
68.4% at prescribed thrust load coefficient, CT=0.728
as shown in Fig. 18. The efficiency of CP-1, HP-1,
CP-3 and HP-2 are higher than that of CP-2, that is,
the MAU propeller that is a Japanese standard

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

04-05-13

propeller in this order. Since the efficiency of HP-3


and HP-1 increase with lowering the cavitation
number, they have superior cavitation performance as
HP-2 but show sharper drop in the efficiency than HP2, CP-1 and CP-3 against the increase of thrust load.
Under the design condition at VS=30.6kt, little
cavitation was found on the blades of CP-3, while CP1 and CP-2 were fully cavitated.

Fig. 19 Propeller Efficiency Measured in Propeller


Open Test in the NMRI Cavitation Tunnel
EVALUATION OF DESIGNED PROPELLER
PERFORMANCE BEHIND SHIP HULL
Tested Ship Model and Propeller Models
This paper also evaluates the performance of the
designed propellers working behind a complete ship
model in the NMRI Large Cavitation Tunnel. In the
cavitation test, the ship model was installed in the
No.2 working section as shown in Fig. 20. In order to
stimulate and stabilize cavitation occurrence, two lines
of stainless wires were equipped at the square station,
S. S. 2 1/2 to generate hydrogen bubbles by electrolysis.
Before the cavitation test, the resistancepropulsion tests and the wake measurement were
performed at the NMRI 400m towing tank. The
experimental conditions were determined based on the
self-propulsion. The propeller models as shown in
Table 4 were used for the behind cavitation tests.

Fig. 18 Propeller Efficiency Measured in


Propeller Open Test in the NMRI Cavitation
VS=35.0kt; V=1.048
HP-5 shows the highest propeller efficiency of
64.7% at CT=0.728 as shown in Fig. 19. The
propeller efficiencies of HP-7, HP-4 and HP-6 are
higher than that of CP-3 in this order. The efficiency
of all of HPs increases with the decrease of cavitation
number. HP-4 and HP-5 generate the necessary thrust
but no sheet cavitation occurred on these propellers
unlike that as predicted. The efficiency of HP-7 was
the second highest but sharply decreased with the
increase of the thrust load coefficient. On the other
hand, the efficiency of this propeller was not so high
but HP-6 kept high efficiency even in the off-design
and highly loaded condition.

Experimental Condition and its Setting


The present cavitation test including pressure
fluctuation measurements and erosion tests were
carried out under the maximum continuous rating
(MCR) condition but the same thrust load condition as
the NOR condition given for the propeller design.

Fig. 20 Arrangement of Ship Model in NMRI Large Cavitation Tunnel

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

04-05-13

Then, since the left-hand side of the equation (4), Ni


and T are known, the unknown coefficients A, B and
C can be calculated by a least squares method (Kudo,
1999a). These coefficients can be used for the
propellers with the same diameter and even for the
propeller with thrust breakdown.
Thus, the
experimental condition on n, Ni and T the pressure at
reference point for the test P0 can be set based on the
measured thrust load coefficient CT.

Usually cavitation tests are performed by the


thrust coefficient (KT) identity method because the
inflow velocity to a propeller model in a cavitation
tunnel cannot be identified easily. This method can be
applied for any propellers but without thrust
breakdown. For the TCPs treated in this paper,
multiple advance coefficients correspond with one
thrust coefficient due to the thrust breakdown. In the
extreme case, the thrust coefficient is constant against
the advance coefficient. Then, it is very difficult or
impossible to identify the correct experimental
condition in cavitation tunnels by the thrust coefficient
identity method.
This paper proposes the thrust load coefficient
(CT) identity method to apply for fully cavitated
propellers with thrust breakdown. A basic idea is to
use a propeller model as a current meter even under
trans-cavitating condition and has been used in the
propulsion test in a towing tank similarly (Ukon,
1999). Using a propeller model whose open water
characteristics are known as a current meter, the
propeller inflow velocity can be determined from the
measured thrust and propeller revolution rate by a
propeller dynamometer under NC condition.
From the torque measurements, the propeller
efficiency in the behind ship condition B and the
relative rotative efficiency R can be determined
using the propeller open water characteristics. From
the propeller open water characteristics measured at a
towing tank, the advance coefficient can be expressed
using the thrust coefficient KT as follows,
2
3
4
J = a0 + a1 KT + a2 KT + a3 KT + a4 KT
(1)
The thrust coefficient is defined as follows,
K = T n2 D 4
(2)

Wake Distribution
The wake distribution was measured by a four-rake
five-hole NPL type pitot tube in the No.2 working
section of the cavitation tunnel. The measured results
on the starboard side are shown in Fig. 21. Before the
wake measurement in the cavitation tunnel, the wake
distribution was measured at the towing tank at Froude
number FN=0.370, using the same apparatus and
measuring system as those in the cavitation tunnel.
Comparing two wake measurements, they are almost
the similar to each other but the axial wake in the
cavitation tunnel is relatively steeper than that in the
towing tank totally. It can be said that the present
experiment should offer more strict evaluation on the
cavitation performance.

where T is thrust, is density of water, n is propeller


revolution rate, DP is propeller diameter.
In the assumed propeller working area including
the propeller operating condition to be tested, the
measurements of propeller thrust are conducted by
changing the propeller revolution rate n and impeller
revolution rate Ni under the NC condition. The inflow
velocity to the propeller is given by
V A = JnDP
(3)
In cavitation tunnels, the inflow velocity to the
propeller should be determined by the following
equation,
V A = A Ni + B T + C
(4)
If the propeller revolution rate n and the thrust T are
known, VA can be obtained by using the equations
(1)~(3) together with impeller revolution rate Ni.
T

Fig.21 Wake Distribution Measured behind


Ship Hull in Cavitation Tunnel
According to the procedure mentioned above, the
propeller performance was measured behind the
complete ship model. The comparison of measured
propeller efficiency of designed propellers is shown in
Fig. 20 under the NOR conditions at 30.6kt and 35.0kt.
Evaluation on 30.6kt Case
As shown in Fig. 22, the behind propeller
efficiency of HP-3 was 64.5% and the highest among
CP-1~3 and HP-1~3 and no big difference among
them was observed except that the efficiency of CP-2
was 59.5% and extremely low under the NOR
condition. Under the MCR condition, the efficiency of
HP-3 was 64.3% and the highest among them.

10

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

04-05-13

that of CP-1. On the hub vortex cavitation (HVC),


those of CP-1, CP-3 and HP-3 were rarely visible,
while those of HP-1 and HP-3 were thicker than CP-1.
The HVC of CP-2 was extremely thick due to nonoptimum circulation distribution especially near the
root.

Propeller Efficiency behind Ship Hull


at NOR Condition (CT=0.576)
66

CP-1(30.6kt)
CP-2(30.6kt)
CP-3(30.6kt)

64

HP-1(30.6kt)

B[%]

HP-2(30.6kt)
62

HP-3(30.6kt)
CP-1(35.0kt)

Evaluation on 35.0kt Case


As shown in Fig. 22, the behind propeller
efficiency of CP-3 was 61.4% and the highest among
the designed propellers under the NOR condition for
35.0kt case. The efficiency of HP-3, HP-5, Hp-7, Hp-6
and HP-4 became worse in this order but the no
remarkable difference among them was found. The
efficiency of CP-1 and CP-2 become tremendously
worse. On the discrepancy on the thrust between the
measurements and targets, HP-4 generated the target
thrust but HP-5, CP-3, HP-6, HP-7 produced 3%
higher, 3% lower, 9% higher and 7% higher than the
target one respectively.
On the other hand, under the MCR condition, the
efficiency of CP-3 was 60.4% and also the highest
among the tested propellers. In the order of HP-5, HP6, HP-7, HP-4, the propeller efficiencies of them
became worse. CP-1 and CP-2 could not generate any
significant performance under this condition.
Cavitation patterns on HP-5, HP-7 and CP-3 are
shown in Fig. 24 under the MCR condition (CT=0.728,
V=1.048).
In this case, CP-3 showed the best performance
on the efficiency. One of the reasons is that less sheet
cavitation and large extent of bubble cavitation
occurred on the blades of CP-3 and the thrust breakdown was small under the behind condition in spite of
hydrogen bubble seeding to supply sufficient nuclei.
Although the propeller models were tested at the
Reynolds number ReK based on the definition by
Kempf around 1.3x106, unstable cavitation on CP-3
was observed. The size of tested propeller models
might be small for this propeller (Kawanami, 2000).
This paper cannot conclude whether the performance
of CP-3 was overestimated or not. This is a future task
on the development of more rigorous cavitation test
procedure.
On the TVC on the tested propellers working
behind the ship hull, those of HP-5, HP-7 and CP-1
were thicker, and HP-4 and HP-6 were thinner than
that of CP-3. On the HVC, those of CP-1 and CP-3
were thin and those of HP-6, HP-4, HP-5 and HP-7
were thicker than the CPs. There is a room to improve
the design techniques.
HP-7 and HP-5 fitted with the propeller boss cap
fin (PBCF) were tested in the behind condition, the

CP-3(35.0kt)

60

HP-4(35.0kt)
HP-5(35.0kt)

58

HP-6(35.0kt)
HP-7(35.0kt)

56

Tested Propellers under Trans-Cavitating Condition

Fig. 22 Comparison of Propeller Efficiency


behind Ship Hull under Trans-Cavitating
Condition
VS=30.6kt Case

(a) CP-1

(b) CP-3

(c) HP-3
Fig. 23 Cavitation of CP-1, CP-3 and HP-3 Working
behind Ship Hull under MCR Condition (VS=31.0kt)
Cavitation patterns are shown in Fig. 23, on CP-1,
CP-3 and HP-3 under the MCR condition (CT=0.728,
V=1.336), stable cavitation was observed on CP-2
and HP-2, while unstable streak and/or bubble
cavitation were found on CP-1, CP-3 and HP-3, in
spite of artificial nuclei seeding. Sheet cavitation on
CP-3 was suppressed under the present condition as
expected.
On the tip vortex cavitation (TVC), those of CP-3
and HP-1 were thinner, those of CP-2 and HP-3 were
almost equivalent and that of CP-3 was thicker than

11

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

04-05-13

HVC became thinner and the propeller efficiency


increased by 2.8% without any harmful effects on the
pressure fluctuations.

Relative Rotative Efficiency


at NOR Condition (CT=0.576)
1.10

CP-1(30.6kt)
CP-2(30.6kt)

1.05

CP-3(30.6kt)
HP-1(30.6kt)

Eta_R

1.00

HP-2(30.6kt)
HP-3(30.6kt)

0.95

CP-3(35.0kt)
HP-4(35.0kt)

0.90

HP-5(35.0kt)
HP-6(35.0kt)

0.85

HP-7(35.0kt)

(a) HP-5

0.80

Tested Propellers under Trans-Cavitating Condition

Fig. 25 Comparison of Relative Rotative Efficiency


behind Ship Hull under Trans-Cavitating Condition
EVALUATION OF PRESSURE FLUCTUATIONS
(b) HP-7

If a high performance propeller induces high


pressure fluctuations, this propeller cannot be accepted
as a good propeller. In this paper, the pressure
fluctuations induced by the designed propellers were
measured behind the ship model hull under the MCR
condition for two cases of ship design speed. The
comparison on the pressure fluctuations was made
mainly at the first blade rate because the higher
components above the second blade rate were much
smaller than the first one for the tested propellers.
The comparison of the pressure fluctuation
amplitude converted to the full-scale value among the
designed propellers is shown in Fig. 26 for two cases
based on that of CP-1. The propeller revolution rates
were estimated from the measured advance coefficients
and assuming that the effective wake coefficients are
not affected by the deviation of the operating condition
from the designed one.

(b) CP-3
Fig. 24 Cavitation of CP-1, CP-3 and HP-2 Working
behind Ship Hull under MCR Condition
(VS=35.5kt)
Comparison of Propeller Efficiency between Open
Water and Behind Conditions
The ratio of the propeller efficiency behind the
ship model to that in the cavitating open water is
shown in Fig. 25 for two cases. The relative rotative
propeller efficiencies of the designed propellers except
CP-3 in 35.0kt case became 5~7% lower than those in
the cavitation open water test except CP-2 for two
cases. Since the existing towing tank tests indicate
that this efficiency of twin screw propeller is 0.95~1.0
in NC conditions, almost the same results were
obtained in the behind cavitation tests under the TC
condition.
On the other hand, CP-3 showed the completely
different tendency of the propeller performance from
other propellers due to less and unstable sheet
cavitation in non-uniform flows brought by unsteady
cavitation and Reynolds effects.

Evaluation on 30.6kt Case


The pressure fluctuating amplitude of CP-3 was
the lowest and in the order of CP-2, HP-3, CP-1, HP-2,
HP-1, the amplitudes became smaller. The amplitudes
under the cavitating became three or four times of
those under the non-cavitating condition. The second
blade rate amplitudes of CPs amounted to around 20%
of the first blade rate, while those of HPs were about
8%. One of the reasons is that on the HPs the cavity
volume variation was smaller but the displacement
effects due to cavity was much bigger than those of
CPs.
The pressure amplitudes of TCPs under the
cavitating conditions become 35~39% higher than that
of CP-3. TCP should be designed by taking into
account of the local flow effects on the propeller and

12

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

04-05-13

cavitation in detail to reduce the cavity volume and


cavity thickness.

Evaluation for 35.0kt Case


More intensive cloud cavitation on the blades and
root cavitation were observed on HP-4, HP-5, HP-6
and CP-3, while misty cloud cavitation was found on
the blade of HP-7. After the 30 minute erosion tests,
small removed off paint area on CP-3, HP-4 and HP-5
was found near the root. The paint on one blade of
HP-7 was removed off. Assuming that the intensity of
erosion is proportional to the removed off paint area,
more intensive erosion was predicted in the order of
HP-4, Hp-5 and CP-3. To reduce root cavitation, a
new blade section was proposed (Ukon, 2004).

Pressure Amplitude against CP-1 at 31.0kt

Pressure Fluctuation Amplitudes


at First Blade Rate Induced by TCPs at MCR Condition
1.5
1.4

CP-1(31.0kt)
CP-2(31.0kt)
CP-3(31.0kt)
HP-1(31.0kt)
HP-2(31.0kt)
HP-3(31.0kt)
CP-1(35.5kt)
CP-3(35.5kt)
HP-4(35.5kt)
HP-5(35.5kt)
HP-6(35.5kt)
HP-7(35.5kt)

1.3
1.2
1.1
1.0
0.9
0.8
0.7
0.6

VS=30.6kt Case; CT=0.718, V=1.371


Propeller Efficiency behind
Ship Hull [%]

Tested Propellers under Trans-Cavitating Condition

Fig. 26 Comparison of Pressure Fluctuations


behind Ship Hull under Trans-Cavitating Condition
Evaluation on 35.0kt Case
As shown in Fig. 26, the first blade rate pressure
amplitude of HP-7 was the lowest and those of HP-6,
HP-5, CP-1, HP-4 and CP-3 became lower in this
order. Roughly speaking, the tendency and the
relation on the pressure fluctuation amplitudes are the
similar to that for 30.6kt case. On the reduction of the
pressure fluctuations, the air injection from the ship
hull is very effective (Ukon, 2000a) though the
additional power for the air injection is needed.

70.0
HP-3

CP-1
65.0

HP-2
CP-3

Good
60.0

HP-1

CP-2

Bad
55.0
0.0

50.0

100.0

150.0

Pressure Amplitude against CP-1 [%]

Fig. 27 Determination of an Optimum Propeller


- VS=30.6kt Case -

VS=35.0kt Case; CT=0.718, V=1.048


Propeller Efficiency behind
Ship Hull [%]

EVALUATION OF EROSION ON DESIGNED


PROPELLERS
Highly loaded propellers treated in this paper
have a certain risk on erosion not only on the blade
surface but also near the root because of propeller
operation in non-uniform flow and unsteady cavitation
occurrence. In this paper, the Aotak Paint Method
was employed for the erosion tests because of short
testing time. The tests were performed under the
MCR conditions.

65.0
HP-7
60.0

CP-3

HP-5
HP-6

HP-4

Good

CP-1

55.0

Bad
50.0
0.0

50.0

100.0

150.0

Pressure Amplitude against CP-1 [%]

Fig. 28 Determination of an Optimum Propeller


- VS=35.0kt Case -

Evaluation for 30.6kt Case


In this condition, misty cloud cavitation was
observed on the propeller blade of CP-1, CP-3, HP-1
and HP-3, while no cloud was found on the HP-2 and
CP-2. After the 20 or 30-minute erosion test, no paint
removed off was detected. No risk of erosion for all of
the designed propellers was predicted.

SUMMARY OF EVALUATION UNDER BEHIND


CONDITION
In order to find the best propeller for this ship
from the experimental results on propeller efficiency,
pressure fluctuation and erosion, a reasonable
comparison is need. Although the risk of root erosion
was detected, the intensity was not so serious. If the

13

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

04-05-13

conventional propellers. The present experiment


indicates that the sophisticatedly designed CP
shows high performance as the TCPs do. From
the present experimental results, the cavitation
experiment with higher Reynolds number and
more careful nuclei control is needed for the
propeller designed to suppress the occurrence of
cavitation.

blade sections near the root are modified, root


cavitation could be reduced effectively. Then, the
comparison was made on two objects, the behind
propeller efficiency and the pressure fluctuation
amplitudes at the first blade rate as shown in Figs. 27
and 28.
For the case of 30.0kt, one can say that CP-1,
CP-3, HP-1 can be evaluated as good propellers in this
order from Fig. 27. On the other hand, HP-7 is the
best propeller and CP-3, HP-4, HP-5, HP-6 might be
evaluated as favorable propellers in this order for the
case of 35.0kt as shown in Fig. 28.

ACKNOULEDGEMENT

This paper describes the theoretical design


method on trans-cavitating propellers (TCPs) and
extensive experimental evaluation on the designed
propellers including conventional propellers (CPs) for
two high powered and high-speed ferries with shallow
draft. The following conclusions can be drawn,

The authors express their gratitude to Mr. Yuzo


Kurobe (Tokyo University of Marine Science and
Technology) and Mr. Noboru Matsuda for their
sincere experiments. The authors also thank to Prof.
Hiroharu Kato, Prof. Hajime Yamaguchi and Dr.
Shunji Soejima for their discussion and support.
The part of work described in this paper was
performed as the cooperative work between the
National Maritime Research Institute and the
Shipbuilding Research Association.

1.

BIBLIOGRAPHIC REFERENCES

CLOSING REMARKS

2.

3.

4.

5.

The present theoretical design method can offer


high performance TCPs roughly within 5%
accuracy on propeller thrust under transcavitating open water conditions. For a CPP type
of TCP equipped to a 35.0kt high-powered large
fast ferry, the efficiency of the designed propeller
can achieve around 65% in the trans-cavitating
open water condition by the present method.
The present design method employs two kinds of
combinations with SC blade sections and NC
blade sections.
Both combinations are very
effective for designing high performance TCPs.
The present thrust load coefficient (CT) identify
method makes possible a reasonable experimental
evaluation on the efficiency of propellers working
in non-uniform flow, such as behind a ship hull
under the trans-cavitating condition.
In the case of CT=0.728 andV=1.371 equivalent
to the ship speed of 30.6kt, a conventional
propeller employing NACA sections with
sophisticated camber and pitch distribution can be
designed as the most efficient and less vibratory
propeller. On the other hand, a TCP can be
designed as a high efficiency propeller but it emits
higher pressure fluctuations than the CP does. In
the design of TCPs, more attention should be paid
on the reduction of pressure fluctuations.
In the case of CT=0.728 and V=1.048 equivalent
to the ship speed of 35.0kt, TCPs can be designed
as favorable propellers with higher efficiency and
less vibratory source than the existing

Kawanami, Y., Ukon, Y., Kudo, T. and Matsuda, N.,


Measurement Techniques on the Hydrodynamic
Characteristics of a Fully Cavitating Propeller, Proc.
of 74th General Meeting of SRI, June 2000, pp.
205208 (Written in Japanese)
Kudo, T. and Ukon, Y., Calculation of
Supercavitating Propeller Performance Using Vortex
Lattice Method, Proc. of The Second International
Symposium on Cavitation, Tokyo, April 1994, pp.
403408
Kudo, T., Fujisawa, J. and Ukon, Y., Estimation
Method of Propeller Advance Ratio for a Cavitation
Test in Non-Uniform Flow, Proc. of 73rd General
Meeting of SRI, June 1999a, pp. 188189
Kudo, T. and Ukon, Y., Cavitation Stimulation
Technique Using Water-Repellent Coating in a
Propeller Model Test, Proc. of 3rd ASME/JSME Joint
Fluid Engineering Conference & 1999 ASME Fluids
Engineering Division Summer Meeting (FEDSM99),
San Francisco, July 1999b
Kudo, T., Ukon Y. and Kato, H., Study on
Theoretical Design Method of Trans-Cavitating
Propeller, Journal of the Society of Naval Architects
of Japan, Vol. 186, Dec. 1999, pp. 4149 (Written in
Japanese)
Kudo, T. and Ukon, Y., Design and Evaluation of
Transcavitating Propellers for High-Speed Vessels,
Proc. of 74th General Meeting of SRI, June 2000, pp.
227230 (Written in Japanese)

14

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

04-05-13

Yim, B., Kim, K.-S., Ahn, J.-W. and Lee, J.-T.,


Design
of Trans-cavitating Propellers and
Performance Analyses of the Test Result, Journal of
Ship & Ocean Technology, Vol. 2, No. 1, June 1998,
pp. 1330.

Kudo, T., Ukon, Y. and Sumino, Y., Proposal of a


Groove Cavitator on a Supercavitating Propeller,
Proc. of CAV 2001, The 4th Int. Symposium on
Cavitation, Pasadena, June 2001
Ukon, Y., Kudo, T. and Hoshino, T., Design
and Evaluation of New Supercavitating
Propellers, Proc. of The Second International
Symposium on Cavitation, Tokyo, April 1994,
pp. 395402
Ukon, Y., Kudo, T., Kurobe, Y., Matsuda, N. and
Kato,
H.,
Design
of
High
Performance
Supercavitating Propellers Based on a Vortex Lattice
Method, Proc. of PROPCAV95 Symposium,
Newcastle upon Tyne, May 1995, pp. 151162
Ukon, Y., Kudo, T., Fujisawa, J. and Matsuda, N.,
Experimental Evaluation of Trans-Cavitating
Propellers, Journal of the Society of Naval Architects
of Japan, Vol. 186, Dec. 1999, pp. 5158 (Written in
Japanese)
Ukon, Y., Kudo, T. and Fujisawa, J., Reduction of
Pressure Fluctuations Induced by Cavitating Propellers
due to Air Injection through the Hull at the Stern of a
Ship, Trans. of the West-Japan Society of Naval
Architects, Vol. 99, March 2000a, pp. 3342
(Written in Japanese)
Ukon, Y., Fujisawa, J., Kudo, T. and Kurobe, Y.,
Application of a Theoretical Design Method to Full
Scale Supercavitating Propellers for High-Speed
Boats, Trans. of the West-Japan Society of Naval
Architects, Vol. 100, Sept. 2000b, pp. 133143
(Written in Japanese)
Ukon, Y., Unstable Phenomena due to Cavitation on
Marine Propellers, Proc. of the 11th Symposium on
Cavitation, Sept. 2001, pp. 1316 (Written in
Japanese)
Ukon, Y., Kawanami, Y., Fukasawa, R., Fujisawa, J.,
and Kudo, T., Anti-Root Erosion Blade Section for
Marin Propellers Equipped to an Inclined Shaft, Proc.
of 12th Symposium on Cavitation, Fukuoka, March
204, pp.57~60
Vorus, W. S. and Kress, R. F., The Subcavitating/Super-cavitating Hybrid Propellers, Proc.
of SNAME Spring Meeting/STAR Symposium 1988
Wang, G.-Q. and Yang, C.-J., Design of Cavitating
Propellers by Lifting Surface Theory, Proc. of
PRADS, Sept. 2001
Yamaguchi, H., Kato, H., Maeda, M. and Toyoda, M.,
High Performance Foil Sections with Delayed
Cavitation Inception, Proc. of Int. Symp. Cavitation
Inception, 3rd ASME/JSME Joint Fluids Eng. Conf.,
ASME, San Francisco, FED SM99-7294-1-11 April
1999

15

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Stephane Cordier
Bassin dessais des Carnes, France
In the propeller series, the diameter D is
kept constant when rpm is increased in some cases:
(HP1, 3, 6). These propellers generate higher
pressure fluctuations, which is expected but also
better efficiency. Can you comment on the choice of
D associated with the different rpm?
AUTHORS REPLY
In the propeller design, the designed
propeller diameter was restricted to be less than 5.0
m due to the shallow draft of the present twin screw
high-speed ferry. Then there are no choices of larger
diameter in the present propeller design. The
difference in the propeller revolution rate of HP-1
and HP-3 corresponds to unexpected error due in the
theoretical propeller design, the amount of thrust
breakdown due to the design concept or inclusion of
the tangential wake in the design condition, while
that of HP-6 depends on the design requirement of
the higher propeller revolution rate.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Roger Kinns
RK Acoustics, UK
Variations in cavitation patterns between the
different propeller designs suggest that there may be
large variations in broadband noise. The consequent
vibration can be very annoying and dominate
components at blade rate and its multiples. Has this
been considered in the study?
AUTHORS REPLY
In the present study, cavitation noise emitted
by Trans-Cavitating Propeller (TCPs) was measured,
whose data were not included in this paper.
Rougly speaking, the respective sound
pressure levels generated by TCPs including CP-1
increase 610 dB at frequency corresponding to first
blade rate and 20 dB at the frequency above 10 kHz,
comparing with that under non-cavitating condition.
The sound pressure level of CP-3 was the lowest at
the first blade rate but larger than other TCPs at
multiples, while that of CP-1 is the lowest in the
frequency of 110 kHz.
In general, a cavitation controlled propeller
often offers low pressure amplitude at the first blade
rate but higher one at the higher order of the blade
rates.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Manfred Mehmel
Schiffbau-Versuchsanstelt Potsdam, Germany
Many thanks for your fine presentation.
You mentioned pressure amplitudes and made a
comparison on the base of CP1. My question, can
you give me the total pressure amplitude to get an
idea about the pressure pulses?
AUTHORS REPLY
The present propeller design was made for a
fast twin-screwed ship with very shallow draft. It
was expected that the propeller designed by a
conventional method for this ship, that is, CP-1,
surely causes thrust break-down and high pressure
fluctuations. The press fluctuation amplitudes at the
first blade rate of CP-1, HP-1 and CP-3 are predicted
15, 12 and 11 kPa at the ship sped of 30.6 kt by the
present CT identity method. The pressure amplitudes
of the second and higher blade rates are about 5% of
the first blade rate and negligible, especially for the
hybrid propellers.
At 35 kt, the pressure amplitudes of HP-7,
CP-3 and CP-1 are evaluated 11, 16.5 and 15 kPa.
The pressure amplitude level is unacceptable for the
ship builders and ship owners. The ship hull form to
reduce the effective horse power and to increase tip
clearance should be improved. In the present project,
the reduction of pressure amplitude due to air
injection along the hull surface above the propeller
blade was tested. The reduction rate is proportional
to the air injection volume. The pressure amplitudes
became around 30% of those without air injection
and acceptable level.
REFERENCE
Ukon, Y., et. al., Reduction of Pressure Fluctuations
Induced by Cavitating Propellers due to Air Injection
through the Hull at the Stern of a Ship, Trans. of the
West-Japan Society of Naval Architects, Vol. 99,
(2000.3), pp. 33-42

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics


St. Johns, Newfoundland and Labrador, CANADA, 8-13 August 2004

A Vorticity Based Propulsor Turbulent Inflow Model


Dr. Stephen A. Huyer and Dr. David Beal
(Naval Undersea Warfare Center, Division Newport, RI USA)
ABSTRACT:
A vortex element method used to model a
turbulent propulsor inflow is presented. An undersea
vehicle configuration consisting of vehicle hull,
upstream stators and downstream rotor is used as the
test case. Velocity data previously acquired from
experiments taken for this configuration are used to
initialize, calibrate and validate the model. Vortex
elements are used to represent the horseshoe vortex
structures seen in experimental observations of
turbulent boundary layers. The filaments are initialized
one propeller diameter upstream in the form of single
vortex loops that encircle the vehicle hull. Sinusoidal
perturbations in the vortex loop model the horseshoe
vortex structures. Vortex blobs are used to model the
higher wavenumber component of the turbulence.
Simulations are then compared with mean velocity,
turbulence and velocity spectra from the experiments to
determine the quality of the model.
Excellent
agreement in mean velocity, turbulence and velocity
spectra was seen for the axial component. Agreement
is mostly seen for the radial component except in the
region where the tip vortex is expected. The swirl
component is underpredicted near the surface, with
better agreement outboard. Potential improvements to
the model are then discussed as well as potential
applications for computation of surface boundary
conditions on propuls or rotors.
INTRODUCTION:
Propellers that operate on actual vehicles
ingest a complex unsteady inflow, which greatly
affects the unsteady forces and radiated noise. This
unsteady inflow is generated by the wakes and
boundary layers produced by upstream surfaces. On
undersea weapons and unmanned undersea vehicles,
upstream control surfaces and stators produce wakes,
which combine with the hull boundary layer to produce
a very complex spatially and temporally varying
inflow. As the propeller encounters coherent wakes,
there is an unsteady response resulting in periodic,
narrowband forces resulting in an acoustic signature
that may be used to characterize the vehicle. The

turbulence in the wake can further excite the


narrowband forces due to the inherent unsteadiness.
The hull boundary layer contains turbulent eddies of
various length scales that are ingested into the
propeller. This results in a broadband type of response
function that increases the overall noise. The ability to
properly model the comp lex turbulent inflow is vital to
predict the unsteady hydrodynamic forces and
subsequent radiated noise.
Previous work on turbulent inflow models as
applied to propulsors has treated the problem in the
frequency (wavenumber) domain (e.g. Sevik (1974),
Wojno et al (2002), Lysak and Brungart (2003), Gavin
(2002)). The usual procedure is to define the inflow
velocity flow field.
If homogeneous, isotropic
turbulence is assumed, the correlation function is
known. The Sears gust response function, assuming
the flow can be decomposed into 2-D strips along the
span of the blade, provides an estimate of the blade
response to the ingested turbulence and hence the far
field radiated noise. Difficulties with this method
include the fact that for most propellers, the ingested
boundary layer turbulence is neither homogeneous nor
isotropic. Another difficulty is that the propeller blade
response is an inherently 3-D problem and
consequently the application of 2-D strip theory
problematic.
Gavin (2002) conducted a set of
experiments to measure the velocity correlation in a
flat plate turbulent boundary layer. He then derived a
vortex model of the turbulent boundary layer structures
to model the correlation coefficient.
Lysak and
Brungart (2003) utilize CFD in their methodology.
They assume isotropic turbulence and turbulent theory
to construct the correlation coefficients based on
integral length scale and turbulent kinetic energy.
Time domain methods offer an alternative
approach. The advantage with these methods is that it
is possible to define the structure of the turbulent
boundary layer and evolve it in time, thus
automatically providing the spatial correlation of the
flow field.
Also, three-dimensional solution
methodologies of the propulsor blade response function
may be used in a straightforward manner. Specifically,

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

vortex lattice methods are an approach used to model


the propeller blades and wake (Kerwin, 1984). Here,
the spatially unsteady inflow is used as boundary
conditions to compute the blade circulation and
unsteady loading. This method was extended by Huyer
and Snarski (2002) for examination of fully turbulent
inflows.
Extensive research examining turbulent
boundary layers has described the various vortical
structures present.
Theodorson (1952) originally
proposed turbulent boundary layer structure in the form
of an omega shaped vortex. The feet of the vortex were
parallel to the wall and the main body of the vortex was
inclined to the surface at some angle (thus the omega
shape). Falco (1977) and Head and Bandyopadhay
(1981) both conducted flow visualization and
concurrent hot-wire anemometry in order to more fully
characterize the turbulent boundary layer structure and
subsequent flow evolution. Vortex loops resemble half
a circular vortex ring. As the loop becomes elongated
resulting in moderate aspect ratios, it evolves into what
is described as a horseshoe vortex. As the stretching
becomes severe, high aspect ratios result with each side
of the vortex becoming very close to resemble a hairpin
(thus the name). Vortex loops dominate at the lower
Reynolds numbers whereas horseshoe and hairpin
vortices are seen at higher Reynolds numbers.
Interestingly, these vortical structures are consistently
inclined at a 40 50 deg angle relative to the surface.
In addition, a number of these structures are packaged
together forming a turbulent bulge. This bulge appears
to form an angle of approximately 20 deg with the
surface. These vortices are advected by the local flow
and evolved resulting in the turbulence. The boundary
layer shed from the stator wakes contains turbulent
eddies that produce a mean velocity defect that tends to
further excite the narrowband response. Since the
turbulent boundary layers and stator wakes contain
coherent vortex structures, a vorticity based method
may better model the actual turbulence and provide the
required structure to better predict the turbulent
hydrodynamic forcing and radiated noise.
Recent efforts utilizing vortex filaments to
model turbulent boundary layer flow structure have
been presented by Bernard et al (1999) as well as
Chorin (1993). Particularly, Bernard utilizes a vortex
method that includes surfaces. A layer of vortex sheets
are used to model the layer close to the wall and a
finite difference formula is used to compute the viscous
diffusion dominant near the surface. Vortex filaments
are then constructed from the top layer of vortex sheets
and then are advected by the flow as material elements.
In the outer region, the Euler equations are used to
evolve the vorticity and viscous diffusion is assumed to

be small. A form of this method has been implemented


into a larger vortex method algorithm developed by the
author (Huyer and Grant, 2000).
This paper presents the development of a
turbulent inflow model used to compute velocity
boundary conditions on a propeller. Experimental data
are used to initialize the vortex element calculations
and calibrate the model. Measurement (Muench, 2000)
and subsequent analysis (Huyer and Snarski, 2003) of
the full three-dimensional, unsteady, inflow velocity
field during operation of a SISUP propulsor have
provided a wealth of data to calibrate and validate the
turbulent inflow model.
The inflow is essentially
composed of two separate parts: The hull turbulent
boundary layer and the stator wakes. Vortex filaments
are used to model the hull boundary layer structures.
These essentially provide a low wavenumber
component to the flow. A random distribution of
vortex blobs is then used to model the higher
wavenumber components. Vortex filaments are used
to model the shed vorticity from the stator wakes and
are oriented in such a manner as to provide the
measured velocity defect.
A boundary element
calculation of the hull is used to model the hull near
wall vorticity. Finally, since the stators produce lift
(and hence a swirl component), a vortex lattice is used
to model the mean stator near wall vorticity and wake
vorticity due to this effect. This is treated separately
from the shed stator boundary layer vorticity. Results
of the model are then compared with experimental data
as stated above.
METHODOLOGY:
Turbulent Inflow Modeling:
In order to properly construct the turbulent
inflow, the computational model must reproduce the
mean velocity, turbulent intensity and velocity spectra.
As stated earlier, a turbulent boundary layer can be
represented as a collection of vortical structures. Falco
(1977) and Head and Bandyopadhay (1981) showed
that these structures are oriented in the boundary layer
in a semi-organized manner. This makes the proposed
turbulent inflow model amenable to vorticity-based
models. A qualitative concept of a boundary layer
orientation of the horseshoe vortices is shown in Figure
1.

Figure 1: Qualitative orientation of boundary layer


hairpin vortices.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Recent efforts utilizing vortex filaments to


model turbulent boundary layer flow structure have
been presented by Bernard et al (1999) as well as
Chorin (1993). This method was implemented into the
latest version of a vorticity based solution developed at
NUWC by Huyer (2002) based on work by Huyer and
Grant (2000). In this version, vortex blobs and vortex
filaments are used to model unsteady wake flows.
Expressions for the velocity and vorticity fields due to
vortex blobs are given by (see Marshall and Grant,
1996 for the full derivation):

v v 2
P 3 / 2, x x n / R 2n
v v
v v
v v

u( x, t) = v(x, t ) +
n (x x n )
v v 3
4 x x n
(1)
P(a,z) is the incomplete gamma function with limits P
= 0 at z = 0 and P = 1 as z . For a = 3/2 and z =
x2 , where x is real, P(a,z) is given in terms of the error
function:

P ( 3 / 2, x ) = erf ( x)
2

2xe x

(2)

1/ 2

The expression for the vorticity is:

n
v v 2

2
3 / 2 3 exp( x x n / R n )

Rn

N
v v

( x, t) = P 3 / 2, xv vx 2 / R 2
n
n

v v

n =1
n (x x n )

v v 3
4 x x n


(3)
The vorticity in (3) is divergence free. If the vorticity
is known at the control points, a matrix equation needs
to be developed to solve for the element amplitudes.
The induced velocity due to a set of vortex
filaments with core radius, n , length, l n ,
,
circulation, , and unit circulation unit vector,
n

is given by Bernard (1999) and Chorin (1993) as:

v v
N filaments

v v
n l n (x x n )
n (
u ( x, t) =
r /n )
v
v
3
4

n =1
x xn

(4)

where is a higher order smoothing function given


by:

(r / n ) = (1 (1

3 r 3 ( r / n ) 3
)e
)
2 3n

(5)

v v

r = (x x n )
n
As in the method described by Bernard, only
the vortex stretching and advection term will be
included, so the evolution of vorticity may be
approximated as:

v
v v v
v
v
+ u = ( )u
t

(6)

The second term on the left is due to advection and the


right hand side is due to stretching. Since the blobs
and filaments are advected with the local flow velocity,
the advection term is implicitly included. Also, the
horseshoe vortices may be modeled by a series of
filaments that are connected. Since the endpoints of
the filaments are advected, vorticity is automatically
conserved using the relation:

(l n n ) t+ t = (l n n )t

(7)

This implicitly includes the vortex stretching term.


Since the circulation is constant, lengthening the
filament will reduce the radius and, as can be seen from
the eqn. 8, the vorticity will increase.
This
methodology evolves the volumetric outer boundary
layer vorticity and shed wake vorticity due to lift from
the stators.
Hull Surface Effects:
The near wall vorticity associated with the
hull is modeled using a boundary element method. To
accomplish this, the axisymmetric hull is discretized
using quadrilateral panels. The solution due to Hess
and Smith (1966) is used to solve the integral for the
surface source and vortex panel strengths. This method
can be described as follows:
Each panel on the body surface carries two
velocity generators: a surface vortex distribution lying
in the plane of the panel and a potential source. Figure
2 shows the panel geometry for an arbitrary surface.
The sources ensure the no-flux boundary condition is
met properly. Both distributions are taken to be
uniform over an individual panel and lie in an infinitely
thin sheet on the surface. Thus the vortex strength
parameter characterizing a panel is the velocity jump
across the panel. The velocity due vto a potential
source, , and vortex panel strength, , on a surface
S is:

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

v v v
v v v v
v
1 ( x ) (x x )
1 ( x )(x x )
u ( x) =
dS +
dS
4 S vx xv 3
4 S vx xv 3
=

{ n B n + n B n }
N

U
18.7

n =1

(8)

y, u y

where n and n are the discretized strengths for a


panel n of surface area Sn and:
v
1
(xv xv ) dS
Bn =
(9)
4 S xv vx 3
n

v
The boundary condition at points x m on the surface is:
v
v
v
v
v
u elements( x m ) + u stator( x m ) + U = 0

(10)

v
v
Here, u elements( x m ) is the induced velocity due to all
v
v
vortex elements in the field and u stator( x m ) is the
induced velocity due to the stator.
Vortex Lattice Calculations:
Since the stators produce lift, the surface and
shed wake vorticity will produce an induced velocity.
This is treated as a mean effect. Instead of shedding
the vorticity in the form of filaments every time step
(one possible approach), it was decided to conduct a
separate calculation solving for the flow past a stator
using a vortex lattice method. The modified propulsor
unsteady flow (PUF) code (see Kerwin (1978, 1986),
Huyer and Snarski (2002)) was used for these
computations. The flow field was solved and induced
velocities were comp uted on a fixed grid.
An
interpolation routine was then used to compute the
induced velocity at any point in the field due to the
stators.
The unsteady velocity at a given point in the
field is therefore a summation of the velocities due to
1) Freestream velocity; 2) Hull near wall vorticity; 3)
Stator vortex lattice; 4) Hull boundary layer vorticity
due to filaments and blobs; and 5) Stator boundary
layer shed vorticity using vortex filaments.

n
s

= surface normal

= unit vector streamwise direction

= tangent to normal and streamwise

Figure 2: Undersea vehicle geometry

, u
x, u x

Figure 3: Solid mo del of SISUP stator and propeller


blade. The local and global coordinate system is
displayed. From Muench (2000).
Wind Tunnel Experiments For Model
Calibration/Validation:
Previously conducted experiments were used
to calibrate and validate the turbulence model.
Summaries of these data can be found in Muench
(2001) and Huyer and Snarski (2003).
SISUP Geometry:
The SISUP propeller, developed by the Naval
Ocean Systems Center (see Mautner et al, 1988), is an
open (no-duct) propeller with complex geometry
incorporating rake, twist and skew. The stators are
located upstream of the propeller and are set at an
angle of attack to counterbalance the torque generated
by the propeller. There are a total of eight stators,
equally spaced along the perimeter of the hull in 45 deg
increments.
The propeller consists of six-blades
equally spaced at 60 deg intervals. The SISUP
propulsor configuration is shown in Figure 3. The
global and local coordinate systems are displayed as
well. The X-coordinate is defined positive in the
direction of the freestream. The Y-coordinate is
positive vertical and the Z-coordinate is positive to
port. The local cylindrical coordinate system, x, r, ,
for definition of the inflow measurement plane is also
shown. x is positive in the direction of the freestream
and parallel with the hull; r (y in the figure) is positive
normal to the hull surface; is positive in the counterclockwise direction (looking upstream) and is 0 at an
angle of 83.41 deg relative to top dead center of the
wind tunnel. The local cylindrical coordinate system is
used for all comparisons with experimental data. At
the point of inflow measurement, the hull converges at
an 18.7o angle relative to the tunnel centerline. The
origin of the local coordinate system (x-r) is at the
propeller mid-chord at the root chord location.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

0o

90o

v 0.1
u'

1.0

ux

0.0

0.25

(a)

(a)
MEAN AXIAL VELOCITY
Uinf = 29.3 m/sec, J = 2.343

0.14

r/Rprop=0.39
r/Rprop=0.57
r/Rprop=0.65
r/Rprop=0.82

0.6

r/Rprop=0.98
r/Rprop=1.15

0.4

r/Rprop=1.32

0.2

sqrt(u'x+u' r+u' )/Ui n f

0.12

0.8

ux /Ui n f

TURBULENT INTENSITY
Uinf = 29.3 m/sec, J = 2.343

0.1
r/Rprop=0.39
r/Rprop=0.57

0.08

r/Rprop=0.65
r/Rprop=0.82
r/Rprop=0.98

0.06

r/Rprop=1.15
r/Rprop=1.32

0.04
0.02
0

0
0

20

40

60

80

Theta (deg)

10

20

30

40

50

60

70

80

90

Theta (deg)

(b)

(b)

Figure 4: Axial velocity component of the mean


inflow scaled by the freestream velocity (a) and plot of
the mean axial velocity over two stator wakes (b).
Inflow plane begins 1 cm upstream of propulsor root
leading edge. U = 29.3 m/sec. View is looking
upstream.

Figure 5: Magnitude of the rms turbulence scaled by


the freestream velocity (a) and plot of the turbulent
intensity over two stator wakes (b). Inflow plane
begins 1 cm upstream of propulsor root leading edge.

Inflow Measurements:
Inflow measurements were conducted for
various freestream velocities with non-dimensional
results at 29 m/sec used for model calibration and
validation.
Velocities were chosen so that the
boundary layer thickness in air at 36.6 m/sec was the
same as that for equivalent Reynolds numbers in water
(20.57 m/sec). At the end of the cylindrical UUV
section (before the afterbody), the Reynolds number
was approximately 120 million and the measured
boundary layer thickness was approximately 4.3 cm.
The in-water boundary layer thickness was
approximately 5% less than in-air measurements.

U = 29.3 m/sec. View is looking upstream.


Cylindrical hot-film anemometry was used to
measure the local unsteady velocity field.
TSI
Incorporated constructed customized u xu r (Model
1246AI-20) and uxu (1246C-20) x-wire probes with
probe lengths of 3.05 cm. The cylindrical hot-films
had diameters of 50 m, were 1 mm in length, and had
a frequency response of 50 kHz. Since only two
components could be measured at a time, the
instantaneous three-dimensional velocity field is
uncorrelated. Mean velocity, turbulence and spectra of
the propeller inflow in the axial, radial and tangential
directions were measured and analyzed. Errors in the
axial velocity component were estimated at 5.2% and
errors in the radial and tangential components were

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

estimated at 10.2% based on local mean values. A


rotating hub mechanism was used to collect data at 12
radial and 600 tangential locations.

Mean Velocity Calculations:


The mean velocity field also needed to be
computed to advect the element control points.
Experimental
data
existed
only
immediately
downstream of the stator so that the mean velocity
effects proximal to the stator and upstream of the stator
were unknown. It was therefore determined that the
best way to approximate the mean inflow was to
perform a vortex lattice calculation of the eight stators.
Velocities from the experimental inflow plane were
circumferentially averaged and the swirl (tangential
component) removed.
Axial and radial velocity
profiles were also smoothed so as to be more
representative of a typical turbulent boundary layer.
The axisymmetric velocity field was computed on a
fixed grid with the same resolution (13 radial and 600
circumferential points) at a given axial location as the
experimental measurement grid. These velocities were
extrapolated upstream so that mass was conserved.
The first axial plane was located one propeller radius
upstream. An additional two axial planes of data were
placed over the stator and six additional planes were
downstream of the experimental data plan for a total of
ten axial planes. A vortex lattice calculation of the
flow past the eight stators was then performed (see
Vortex Lattice Calculations section above).
The
induced velocity due to the vortex lattice solution to the
stator flow was then computed on each of the grid
control points. The velocity on the grid was then a
superposition of the original grid velocities and the
induced velocities from the stators and their wakes.
The local velocity of the vortex filaments was then
obtained by linearly interpolating the velocity from the
fixed grid.
Figure 6 shows the results of the
calculations taken midway between the two stators. As
can be seen, the radial profile was smoothed initially
and does not show the defect and overshoot seen at
r/Rprop = 1.0. The tangential profile shows that the
swirl is underestimated by the calculations very near

1.4
1.2

Tangential
1
Axial

r/R prop

Figures 4 and 5 display the axial velocity


component and the turbulent intensity, respectively. It
can be seen that the stator wakes are quite clear and
dominant. They give rise to the tonal in the radiated
noise acoustic signature. In order to model the
turbulence in the stator wakes, vortex filaments
extending along the entire span of the stator will be
used. The filaments may be initialized based on the
velocity defect from the mean velocity profile.
Variation in the filaments will provide the wake
turbulence levels. These data will be used to further
calibrate and validate the model.

MEAN VELOCITY PROFILES


Between Stators, Stator Computations

Radial

0.8
0.6
0.4

-0.3

0.2
-0.1

Experiments
Vortex Lattice Soln.
Vortex Lattice Soln.
0.1
0.3
0.5
0.7
ux , ur or u (r)/U inf

hub

0.9

Figure 6: Mean velocity profiles of the vortex lattice


calculations for flow about the eight stators.
the surface. It does not exhibit the humps in the profile
seen at r/Rprop = 0.73 and 1.06. This may be due to
incorrectly modeling the tip vortex, which the vortex
lattice method is not designed to do. Excellent
agreement is seen in the axial profiles, although the
solution slightly underestimates the velocity close to
the surface. It should be noted, however, that the
solution does not take into account the induced velocity
due to the rotor. Rotor effects will be discussed later.
Vortex filament initialization:
The vortex filaments are initialized with a
circulation, radius and length and placed upstream of
the propulsor rotor immediately downstream of the
stator. Vortex filaments representing the hull boundary
layer and the stator wakes are treated separately. The
boundary layer vortices are modeled using single
vortex loops that encircle the vehicle hull. Figure 7
shows an example of a single boundary layer vortex
loop at successive downstream stations. Each vortex
loop consists of approximately 200 vortex filaments.
Superimposed on the ring shape is a sinusoidal
perturbation. This perturbation is intended to model
the horseshoe vortices. This perturbation results in a
number of effective horseshoe vortices that are inclined
at angles between 40 to 50 deg relative to the hull
surface. The number of perturbations for each loop is
randomly varied between 20 and 60 and an additional
random perturbation of 0.01 is introduced in the axial
and radial direction that results in the differing
inclination angle. At each time step (t = 0.025), 11 of
these vortex loops are evenly stacked normal to the
surface beginning at a normalized distance of 0.05
above the hull surface extending out a distance of 0.25.
As the individual vortex loops are advected
downstream, Figure 7 shows that the mean shear

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

1.1
Axial Velocity
0.0

Hull Surface

Inflow Plane

Computational Outflow
Rake
Plane

Figure 7: Plots of single vortex loops at five different


timesteps.
boundary layer flow stretches the vortex filaments
creating elongated hairpin vortices.
To maintain
resolution, filaments are divided equally after they
double in length based on their initial value. Figure 8
shows a collection of the boundary layer vorticity using
vector plots at each filament control point. Vortex
loops are removed after they have advected two
propeller radii downstream. For a typical calculation,
the boundary layer consists of 1,000 loops and 200,000
filaments. To match the mean velocity profile of the
hull boundary layer a mean circulation of 0.002 is used
for each filament. To better achieve the profile, the
first two layers have circulations 1.5 times the mean
value and the outer four layers have circulations half
the mean value. To provide an additional random
component modeling the turbulence, the circulation
and radius are randomly perturbed by 50% of the mean
value.
The model of the stator wake vorticity is
intended to capture both the shed boundary layer
vorticity and wake vorticity due to lift. To accomplish
this, a simple vortex pair model is used. The wake
thickness is known from the experimental results and
the vortex lattice model provides the circulation of the

Figure 8: Isometric and side views of the filament and


blob velocity vectors at the element control points.
vorticity (due to lift) initially shed into the wake. What
is not provided is the effect of the boundary layer
vorticity shed. Each vortex loop consists of 36
filaments where the filament control points correspond
to the control points at the stator trailing edge from the
vortex lattice code.
The filament centers
corresponding to an upper and lower loop are separated
a distance of 0.5 wake thicknesses (t wake) in the axial
direction and 0.25 twake in the circumferential direction.
Additionally, each filament control point is randomly
perturbed an additional distance of 0.125 twake in the
axial direction and 0.05 twake in the circumferential
direction. To achieve the correct velocity defect, a
filament circulation of 0.005 is used. A circulation
differential corresponding to the vortex lattice output of
the wake circulation si used to model the stator lift
effect. To introduce an additional component of

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

turbulence, each filaments circulation and radius is


perturbed by 10% of its mean value.

Experiments
ur

0.8

Velocity (ux , ur or u /Uinf)

Finally, vortex blobs are used to introduce a


higher wavenumber turbulent component. These blobs
are introduced immediately downstream of the stator at
each time step with 300 blobs in the circumferential
direction and six radial stations for 1800 blobs each
time step. Typically, this results in approximately
100,000 additional blobs in the boundary layer. Each
blob has a radius of 0.01 with the radius perturbed by
50% of its mean value. Blob circulation is randomly
initialized between +0.0005.

CIRCUMFERENTIAL VELOCITY DISTRIBUTION


Simulation
r/Rprop=0.39

Axial
0.6
0.4
0.2
Radial
0
0

20

40

60

80

100

-0.2
Tangential
-0.4

Figure 8 shows an element vector plot of the


flow domain. Velocity vectors are colored based on
the axial velocity and are plotted at the filament and
blob control points. The filament initiation plane, the
plane where experimental data were collected, and the
outflow plane where points are removed are clearly
delineated.

Circumferential Angle

Figure 9: Axial, radial and tangential circumferential


velocity distributions for simulated and experimental
cases taken at r/Rprop = 0.39.
CIRCUMFERENTIAL VELOCITY DISTRIBUTION
r/Rprop =0.65
1
0.8

Velocity (ux, u r or u /Uinf)

The effective turbulent intensity and velocity


spectra data due to the vortex blobs and filaments were
computed and compared with the experimental
database. The size, strength and distribution of the
vortex filaments were adjusted to match the
experimental data with the vortex blobs providing an
additional random component to mimic the higher
wavenumber behavior.
Once qualitative and
quantitative agreement was reached, the data were
deemed suitable for the full flow simulation.

Axial

0.6
0.4
Radial
0.2
0
0

-0.4

40

60

80

100

Tangential

Simulation
Experiments
ur

Circumferential Angle

Figure 10: Same as Figure 9 except r/Rprop = 0.65.

CIRCUMFERENTIAL VELOCITY DISTRIBUTION


r/Rprop =0.9
1

Velocity (ux, u r or u /Uinf)

RESULTS:
Mean Velocity Comparisons
Figures 9, 10 and 11 show circumferential
distributions of axial, radial and tangential velocity at
r/Rprop = 0.39 (closest to the surface), 0.65 (middle of
the boundary layer) and 0.9 (outer boundary layer
under the influence of the tip vortex). For these and all
cases, comparisons are made in the local (x, r, )
coordinate system. The plots are shown from position
angles between 0 and 100 deg encompassing two
stators. For r/Rprop = 0.39, the axial velocity displays a
clear drop in velocity due to the stator wake.
Simulations slightly underpredict by 5% the magnitude
of the velocity decrease and demonstrate overshoots
not seen in the experimental data. This is likely due to
the lack of resolution of the boundary layer vorticity
shed from the stator. Aside from this, the simulated
axial velocity predicts the measured velocity between
the stator wakes to within 2%. The simulated radial
velocity distribution appears to follow the measured
distribution, but is offset by a value of -0.02.
Simulations of the tangential component do a better job

20

-0.2

0.8

Axial

0.6
0.4

Radial

0.2
0
0
-0.2
-0.4

20

40

60

80

100

Tangential

Circumferential Angle

Figure 11: Same as Figure 9 except r/Rprop = 0.9.

Simulation
Experiments
ur

of following the measured values. In all cases, the


effect of the wake is predicted by the simulations to
within 5%.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

1.2
Tangential
1
r/R prop

Axial
Radial

0.8
0.6
0.4

-0.3

0.2
-0.1

0.1

Experiments
Computations
R/Rprop
0.3
0.5
0.7
ux, u r or u (r)/Uinf

hub

0.9

Figure 12: Axial, radial and tangential velocity profiles


taken midway between two stators for simulated and
experimental cases.
At the middle of the boundary layer (r/Rprop =
0.65), the simulated axial velocity closely follows the
measured. The wake defect is slightly underpredicted
as is the mild overshoot prior to entering the wake.
Between the stators, the velocity distribution is
relatively flat with predicted values off less than 1%.
The predicted radial values appear to be greater by a
value of 0.02 compared with the measured velocities.
Predicted tangential velocities are greater by a constant
value of 0.1.
At the edge of the boundary layer, under the
influence of the tip vortex (r/Rprop = 0.9), there
continues to be excellent agreement between the
predicted axial velocities and the measured. The
magnitude of the velocity defect is predicted within
1%. Also, the mild decrease in velocity between the
stator wakes is replicated as well. The computations
slightly overpredict the velocity between the stators by
3%. The trends in the radial velocity are predicted as
well, but the simulated distribution appears shifted
upward by a value of 0.05. The simulated and
measured tangential velocities agree very well except
for the stator wake. Here, the tip vortex can be inferred
in the measured velocity distribution where an
overshoot on the left side of the wake followed by an
undershoot on the right side of the wake is observed.
This suggests that the measurement probe came very
close to traversing the center of the tip vortex. The
predicted values only display an undershoot suggesting
that the center of the tip vortex is still above the
computational rake.
Figure 12 shows radial axial, radial and
tangential velocity profiles taken midway between the

first two stators to highlight the hull turbulent boundary


layer.
Unlike the mean velocity vortex lattice
calculations shown in Figure 6, this figure shows the
time averaged velocity field based on the fully
turbulent inflow. Excellent agreement is achieved with
the axial velocity with all values within the
experimental error of 5.2% (shown by the error bars).
The radial profile agrees in the near wall region.
Outboard of r/Rprop = 0.8, where the tip vortex would
be expected, the agreement is not as good. The current
model of the tip vortex does not capture the radial
velocity defect and overshoot seen in the experiments.
Also at the furthest outboard locations, simulations
over-predict the radial component by a value of 0.04.
As a reference, the freestream radial component is
0.32, so that for both simulations and experiments, the
freestream values have not been reached.
Alternatively, the tangential profiles agree generally
well outboard of r/Rprop = 0.73, but do not agree well
near the wall. In the near wall region (below r/Rprop =
0.65), tangential velocity is underpredicted, except for
the radial location closest to the surface. In the outer
boundary layer, there are two mild overshoots in the
measured profile at r/Rprop = 0.82 and 1.06, but none in
the predicted.
Turbulent Velocity:
Propulsor inflow turbulence was estimated by
taking the standard deviation of the unsteady velocity
in the axial, radial and tangential directions. Figures
13, 14 and 15 show radial profiles taken between the
two stators of the axial, radial and tangential turbulence
components respectively. Simulations are compared
with measured values and flat plate turbulence profiles
are shown to provide another comparison. In all three
cases, the simulated turbulence is significantly greater
0.1

rms TURBULENCE PROFILES


Axial Direction

0.09
rms Turbulence / U inf

1.4

VELOCITY PROFILES
Between Stators

0.08

Flat Plate TBL, zero P (Klebanoff)

0.07

Computations
Experiments

0.06
0.05
0.04
0.03
0.02
hub

0.01
0
0.3

0.5

0.7

0.9
r/R prop

1.1

1.3

Figure 13: Profiles of the axial Turbulence with


comparisons to experimental data and flat plate
turbulent boundary layer data (Hinze, 1975).

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

rms TURBULENCE PROFILES


Radial Direction

0.09

0.08

Flat Plate TBL, zero P (Klebanoff)

0.07

Computations
Experiments

0.06
0.05
0.04

0.08

Turbulence (u'x /Uinf)

0.09
rms Turbulence / U inf

AXIAL TURBULENCE
r/Rprop =0.73

0.1

0.1

0.03

0.07
0.06
0.05
0.04
0.03
0.02
Simulation

0.01

0.02
hub

0.01

Experiments

0
0

20

0
0.3

0.5

0.7

0.9
r/R prop

1.1

1.3

Figure 14: Same as Figure 13 except radial turbulence.


0.1

60

80

100

Figure 16: Circumferential distribution of the axial


turbulence with comparisons to experimental data for
r/Rprop = 0.73.
AXIAL VELOCITY POWER SPECTRA
r/Rprop = 0.47

rms TURBULENCE PROFILES


Circumferential Direction

0.09

1.E-04

0.08

Flat Plate TBL, zero P (Klebanoff)

0.07

Computations
Experiments

0.06

Computations
Experiments

1.E-05
Power (u/Uinf)^2

rms Turbulence / U inf

40

Circumferential Angle

0.05
0.04
0.03

1.E-06

1.E-07

1.E-08

0.02
hub

1.E-09

0.01

20

40

60

80

100

120

Frequency/Shaft Rate

0.3

0.5

0.7

0.9
r/R prop

1.1

1.3

AXIAL VELOCITY POWER SPECTRA


r/Rprop = 0.65

Figure 15: Same as Figure 13 except tangential


turbulence.

1.E-04

Computations
Experiments

than the measured values at the lowest radial location


(r/Rprop = 0.39). By the second measurement location
(r/Rprop = 0.48), the simulated axial turbulence is in
much better agreement with the measured values.
From r/Rprop = 0.65 to the outboard locations, all three
turbulent components are in good agreement with the
measured values. At stations outboard of r/Rprop = 0.9,
the freestream turbulence of 1% is evident in the
measurements. Freestream turbulence was not part of
the simulated turbulence.
Figure 16 shows a circumferential distribution
of the axial turbulence at r/Rprop = 0.73, in the middle
of the hull turbulent boundary layer.
Overall,
agreement with the measured turbulence is seen in the
region between the stators.
Simulations slightly
underpredict the stator wake turbulence and the
simulated wake appears wider by comparison. Not

Power (u/Uinf)^2

1.E-05
1.E-06
1.E-07
1.E-08
1.E-09
1.E-10
0

20

40

60

80

100

120

Frequency/Shaft Rate

Figure 17: Axial velocity spectra with comparisons to


experimental data for r/Rprop = 0.47 and 0.65.
Frequency is non-dimensionalized by the rotor blade
shaft rotation rate of 27.3 Hz.
shown are the radial and tangential distributions. They
demonstrate qualitatively similar behavior with some
important differences to note. The broadband radial

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

appears represented, but the increased turbulence in the


stator is not seen in the radial component. This is due
to the simple vortex loop model of the wake flow.
Since the vorticity is along the span of the stator, there
is no radial velocity component, only axial and
tangential. Simulations indicate that the tangential
component of turbulence exceeds that measured by
approximately 3% in the stator wake. As seen in
Figure 15, the broadband is represented well.
Velocity Spectra:
Velocity spectra were computed based on the
unsteady velocity data and appropriately scaled for
comparisons with the experimental data. Figure 17
shows the axial velocity spectra for r/Rprop = 0.47 and
0.65 at a circumferential position midway between two
stators. 500 data points were used to compute the
spectra at a non-dimensional time step of 0.02. This
yielded a maximum non-dimensional frequency of 120.
As can be seen, there are considerably more
fluctuations in the simulated spectra compared with the
measured. This is likely due to the limited resolution.
Regardless, the slope of the data agrees very well with
the measured spectra with the data characteristically
broadband turbulent.
DISCUSSION AND CONCLUSIONS:
A model of the fully turbulent unsteady
propulsor inflow based on a vortex element method has
been presented.
Vortex filaments were used to
construct the vortices characteristic of experimentally
observed turbulent boundary layers. Initially, a number
of sinusoidal perturbations in a single vortex loop that
completely encircled the vehicle hull modeled the
boundary layer horseshoe vortex structures. Several of
these vortex loops were introduced one propeller radius
upstream for each time step. The vortex filaments
were advected by the mean flow, which was computed
using a vortex lattice method at a number of points
corresponding to a fixed grid. The mean flow included
effects from the hull and the upstream stators and the
stator wakes. The unsteady stator wake component of
the inflow model was initialized using a single vortex
loop consisting of several filaments and fully
connected. This modeled the shed vorticity from the
stator upper and lower surfaces. As the flow evolved,
velocity shear caused the boundary layer vortex
filaments to stretch so that in some cases, high aspect
ratio vortex structures were formed resembling hairpin
vortex structures. The model was constructed to
replicate as many of the physics as possible.
Experimental data were used to initialize the
circulation and radius of the vortex elements. A
random component in circulation, radius and position
was introduced to the vortex filaments to add a

component of turbulence. Random distributions of


vortex blobs with zero mean and various circulations,
positions and radii added another higher wavenumber
component.
Comparisons
with
experimental
data
demonstrated excellent agreement in the mean axial
velocity data. Turbulent boundary layer velocities and
wake velocity defects were both reproduced. In some
cases, the spatial location of the velocity defect was
slightly out of phase with the measured position. This
is likely due to the differences in computed tangential
velocity based on the vortex lattice method and the
measured. These differences were evident in the model
predictions of the tangential velocity component as
well. Mean tangential velocities were consistently
underpredicted near the surface (below r/Rprop = 0.65)
by values on the order of 0.05 to 0.1. Radial velocities
matched well near the surface, but were off in the
region of the tip vortex. Some explanations for the
observed discrepancies include neglecting the effect of
the rotor. Vortex lattice calculations of the flow past
the rotor, based on the experimental inflow, showed
very little contribution upstream in the swirl velocity.
Axial velocities of 0.03, radial velocities of 0.02 and
negligible tangential velocities were computed at the
surface with decreasing values at outboard locations.
By including the axial velocity effect, the subsequent
vorticity close to the surface will need to be increased
to match it possibly altering the swirl velocity
component. Another possible cause to note is the lack
of a stator root vortex model. As was seen from the
vortex lattice calculations to compute the mean flow,
the swirl close to the surface was underpredicted. This
may be expected for fully potential flow, but it was
hoped that the vortex filaments would better model the
swirl near the surface. By introducing a root vortex
model, it may be possible to increase the swirl near the
surface to better match the experimental observations.
It also appears that an improved tip vortex model is
needed to better predict the changes in radial velocity
seen in the radial profiles. It appears that this effect is
felt between the two stators. The induced velocity
effect from the model is much more localized by
comparison. These hypotheses will be tested in the
upcoming research in order to improve the model.
Substantial agreement was obtained with the
axial turbulence component. Again the hull boundary
layer and stator wake were matched very well. Points
closest to the surface were not matched as well. This is
likely due to the lack of resolution and use of the
vortex sheets to model the near wall boundary layer.
At outboard locations, the radial and tangential
components remained approximately 1%-2% high.
Again, this could be due to a lack of spatial resolution

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

of the vortex structures in the radial and tangential


directions. The fact that the overall agreement matched
as well as it did, however, is highly encouraging for
this model.
Velocity spectra of the computed turbulent
inflow model demonstrated that the wavenumber
characteristics of the inflow could be matched.
Increased noise in the inflow model spectra may be
explained by a lack of temporal resolution. Additional
runs are planned to compute the spectra for a large
number of points to address this issue. The overall
agreement with experiment suggests that the length
scales of the vortex structures are being modeled
properly.
The turbulent inflow model is now at the stage
where it can be used to compute turbulent inflow
boundary conditions on a propulsor rotor. These
computations are currently being conducted, but are
not the focus of this paper. After the unsteady forces
are computed, the direct radiated nose will be
computed based on the unsteady blade forces. The
current set of experimental data includes radiated noise
data. Comparisons will be made with the predicted and
measured radiated noise to fully evaluate the potential
of this model.

Gedney, C.J., Abbot, P.A., and Corriveau, P.J.,


Inferring Blade Rate Forces from Wind Tunnel Sound
Power Measurements, ASME Winter Annual
Meeting, Symposium on Flow Noise Modeling,
Measurement and Control, November 1998.
Gavin, J.R., Unsteady Forces and Sound Caused by
Boundary Layer Turbulence Entering A
Turbomachinery Rotor, Ph. D. Thesis, Graduate
Program in Acoustics, The Pennsylvania State
University, August 2002.
Hahn, N.J., Renick, D.H., Taylor, T.E., PUF-14.4: An
Unsteady Analysis Code for Wake-Adapted, Multistage Ducted Propulsors, Massachusetts Institute of
Technology Department of Ocean Engineering,
December, 2000.
Head, M.R., Bandyopadhyay, P., New Aspects of
Turbulent Boundary Layer Structure, Journal of Fluid
Mechanics, Vol. 107, pp. 297-338, 1981.
Hess, J.L., Smith, A.M.O., Calculation of Potential
Flow About Arbitrary Bodies, Progress in Aerospace
Science, Vol. 8, pp. 1 38, 1966.
Hinze, J.O. Turbulence, 2nd Edition, McGraw-Hill,
1975.

ACKNOWLEDGEMENTS:
This work was funded by the NUWC Internal
Research Program, Mr. Richard Philips, program
manager and the Office of Naval Research under
Contract N0001402WX20474, Dr. Kam Ng, program
manager. The experimental data used for comparisons
was collected during SISUP experiments, supervised
by Dr. John Muench for his thesis work and funded by
the Office of Naval Research under Dr. Patrick Purtell.

Huyer, S.A., Snarski, S.R., Analysis of Turbulent


Propeller Inflow, ASME Journal of Fluids
Engineering, May 2003.

REFERENCES:
Bernard, P.S., Dimas, A.A., Collins, J.P., Turbulent
Flow Modeling Using a Fast, Parallel, Vortex Tube and
Sheet Method, European Series in Applied and
Industrial Mathematics, Vortex Flows and elated
Numerical Methods III, ed. Gagnon, Y., Cottet, G.-H,
Dritschel, D.G., Ghoniem, A.F., Meiburg, E., Vol. 7,
pp 46-55, 1999 (http://www.emath.fr/proc/Vol.7/)

Huyer, S.A., Grant, J.R., "Simulation of UUV


Recovery Hydrodynamics," Proceedings from the 23rd
Symposium on Naval Hydrodynamics, Val de Reuil,
France, Septemb er, 2000.

Chorin, A.J., Hairpin removal in vortex interactions


II, Journal of Computational Physics, Vol. 107, 1993,
pp 1-9.

Keenan, D.P., Marine Propellers in Unsteady Flow,


Ph.D. Thesis, Department of Ocean Engineering,
Massachusetts Institute of Technology, May 1989.

Falco, R.E., Coherent Motions in the Outer Region of


Turbulent Boundary Layers, Physics of Fluids, Vol.
20 (10), Part II, October 1977.

Kerwin, J.E., Lee, C.S., Prediction of Steady and


Unsteady Marine Propeller Performance by Numerical
Lifting Surface Theory, Transactions of the Society of
Naval Architects and Marine Engineers, Vol. 86, 1978.

Huyer, S.A., Snarski, S.R., Unsteady Propulsor Force


Prediction for Spatially and Temporally Varying
Inflow, ASME Paper No. FEDSM2002-31346, ASME
Fluids Engineering Division Summer Meeting,
Montreal, Quebec, Canada, July 2002.

Huyer, S.A., Grant, J.R., "Computation of Unsteady


Naval Hydrodynamics Using a Lagrangian Vorticity
Method," AIAA Paper No. 2000-2532, June 2000.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Kerwin, J.E., Marine Propellers, Annual Review of


Fluid Mechanics, Vol. 18, pp 367-403, 1986.
Lysak, P.D., Brungart, T.A., Velocity Spectrum
Model for Turbulence Ingestion Noise from
Computational-Fluid-Dynamics Calculations, AIAA
Journal, Vol. 41, No. 9, pp 1827-1829, September
2003.
Marshall, J.S., Grant, J.R., Penetration of a Blade into
a Vortex Core: Vorticity Response and Unsteady Blade
Forces, Journal of Fluid Mechanics, Vol. 306, pp. 83109, 1996.
Mautner, T.S., Nelson, D.M., Gillcrist, M.C.,
Investigation of the SISUP (Swirl Inducing Stator
Upstream of Propeller) Concept for Marine
Propulsion, Naval Ocean Systems Center,
Proceedings: Propellers 88. September 1988.
Muench, J.D., Periodic Acoustic Radiation from a Low
Aspect Ratio Propeller, Ph.D. Thesis, Department of
Mechanical Engineering and Applied Mechanics,
University of Rhode Island, 2001.
Sevik, M.M., The Response of Propulsors to
Turbulence, Proceedings of the 7th Symposium on
Naal Hydrodynamics, Rome, Italy, August 25-30,
1968.
Theodorsen, T., Mechanism of Turbulence,
Proceedings of the 2nd Midwestern Conference on
Fluid Mechanics, Ohio State University, 1952.
Uhlman, J.S., An Examination of the Frequencies of
the Unsteady Harmonic Forces Genereated by
Propulsors, NUWC-NPT Technical Report 10,470,
Naval Undersea Warfare Center, May 1995.
Wojno, J.P., Mueller, T.J., Blake, W.K., Turbulence
Ingestion Noise, Part 2:Rotor Aeroacoustic Response
to Grid-Generated Turbulence, AIAA Journal, Vol.
40, No. 1, pp 26-32, January 2002.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Stephane Cordier
Bassin dessais des Carnes
In the paper, the vortex filament intensities
are tuned to one geometry. How would this
method be applied when experimental data is not
available for a new geometry? Do the authors have
examples of this method applied based on a
computed (RANSE or LES) flow?
AUTHORS REPLY
It is correct that the results presented in this
paper were calibrated for a particular SISUP
geometry. Experimental mean 3-D velocity and
turbulence data were used in this process at a given
upstream plane. It would actually be easier to utilize
CFD data to calibrate the model for a generic
geometry, but, at present, we have no examples to
present. As long as mean and rms velocities are
computed, this information can be used to calibrate
the model.
Right now, this method is being
transitioned to a 6.2 program where turbulent inflow
into a propulsor will be computed. Unlike the SISUP
cases presented where experimental data were used,
the inflow velocity data is computational using a
RANS formulation. Mean velocities and turbulent
intensity will be used to calibrate a model to predict
the broadband unsteady hydrodynamic blade forces
and pressures. We will present these results in the
future as the calculations progress.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Joseph R. Gavin
General Dynamics Electric Boat Division, USA

need of computational approaches, and the present


research might make a real contribution in trending
the physics.
AUTHORS REPLY

The authors are to be congratulated on an


important and well executed piece of work. The
inflow modeling is closely related to observable
quantities that give wonderful insight to the
turbulence physics, and certainly this improves our
collective abilities in modeling, testing and
mitigation. I am anxious to see the next step, which
will compare the measured and predicted
aeroacoustic response.
I am still frankly unclear if this approach
will serve mainly to increase our knowledge, or if it
will ultimately become a robust design tool. The
authors vision for their research would be most
welcome, along with some insight to the
computational load for a typical problem, and the
perceived robustness of the calibration constants.
How might this method compliment the existing
infra-structure of RANS, LES and experiments?
Thus far the model has achieved good
correlation with single point turbulence statistics.
The boundary layer mean flow, blade wakes, and
RMS turbulence statistics are certainly within
encouraging agreement of measurements.
The
spectral content of the velocity fluctuations might be
improved with finer discretization and longer
averaging as suggested, but our independent work
also emphasizes the need for great care in avoiding
artificial periodicity within the flow domain. The
near-wall behaviors also require some attention, but
these are generally less important to the turbulence
ingestion problem. The authors have probably
considered the use of vortical image sources below
the wall to better enforce the no-slip boundary
condition, but its unclear if this is already included or
a future enhancement.
The authors are certainly well aware that the
upcoming aeroacoustic predictions will rely not just
on simulating the single-point statistics, but also the
spatial distribution (or wavenumber content) of the
turbulent velocity fluctuations. The turbulent blobs
are a key feature for modeling the small scale
fluctuations. The key question remains if the
underlying structure is sufficiently random so that the
velocity fluctuations become completely decorrelated
at reasonable axial and circumferential separation
distances.
The authors should also consider the
extension of this approach to the more challenging
problem of predicting the wavenumber-frequency
content of the wall pressure fluctuations beneath
turbulent boundary layers. This is a problem in sore

Currently, this approach is intended mainly


as a robust design tool to mimic the physics
associated with turbulent inflows. It is not entirely
physics-based
as
either
empirical
and/or
computational data are required to initialize and
calibrate the filament and blob circulations.
Evolution of the vorticity associated with the
simulated turbulence is physics-based. This does not
mean, however, that problems cannot be designed to
increase our knowledge. By using this methodology
to establish a structurally realistic turbulent inflow,
specific response of some type of surface to the
resultant flow can be examined. In the present
problem, we examine the propeller response to the
turbulence to predict the radiated noise. Other
applications include examination of sonar or wide
aperture array surface response to better understand
the self-noise problem. Also, problems such as
modification of the turbulence due to ingestion into a
propeller may be examined. Eventually, a physicsbased approach is desired but is beyond the scope of
the currently funded research.
Indeed, there is very good agreement with the single
point turbulence statistics and agree that finer
discretization in space and time will improve the
agreement. We are in the process of conducting
longer runs (in time) with these goals in mind.
Introducing a false periodicity into the flow is
certainly a potential problem if care is not taken to
introduce sufficient randomness. For that reason,
Monte-Carlo methods are used to both spatially and
temporally introduce the vorticity upstream to
minimize this effect. As may not have been made
clear in the paper, a boundary element solution
solving for the no-slip and no-flux boundary
conditions using vortex and source surface panels is
used. The thin vortex sheets obtained from this
solution ensure that these boundary conditions are
met. The right hand side of the matrix equation
includes the induced velocities from all velocity
generators including the freestream and volume blob
and filament vorticity. This effectively addresses the
vortical image sources that Dr. Gavin speaks of.
We are currently conducting aeroacoustic
predictions with some initial results presented at the
Symposium. The radiated noise computations agree
well with experiments. The broadband components
are noisier compared with experimental data due to
the lack of sufficient averaging of the computational
results. Again, longer time runs should address this

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

problem. Also, initial investigation of two-point


statistics has shown that the turbulence is in fact
decorrelated at reasonable axial and circumferential
distances.
We greatly appreciate the suggestion that the
current method may be used to examine the
wavenumber-frequency content of wall pressure
fluctuations for turbulent boundary layers. This is an
example of a research topic that can be examined to
increase our knowledge of the underlying physics.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Toward High-Fidelity Prediction of Tip-Vortex Around


Lifting Surfaces What Does It Take?
Sung-Eun Kim and Shin Hyung Rhee
Fluent Incorporated, Lebanon, N.H., 03766, U.S.A.

The presence of multiple features with


widely varying length scale translates into challenge
in meshing. The question comes down to how to
effectively allocate computational elements in the
solution domain so that one can obtain the best
possible solution accuracy, maximizing the return of
investment in terms of the usage of the
computational resources. Global mesh refinement refining the mesh everywhere - is obviously a very
expensive proposition for three-dimensional flows.
One very promising avenue can be found in local
mesh refinement and coarsening based on the
solution itself, which is made possible by the
numerics capable of handling unstructured meshes.
Another difficulty, arguably the most
challenging of all, comes from turbulence modeling.
Typical lifting surface flows, even at a moderate
angle of attack, involve strongly non-equilibrium
boundary layer, undergoing a rapid change from the
moment it impinges on the wall until it leaves or
separates from the surface in the form of a free
vortex-sheet. Furthermore, the flows off the wall are
dominated by strong rotation in the vicinity of the
vortices generated around the wing tip. As well
known, isotropic eddy-viscosity based turbulence
models are ill-equipped to handle this kind of flows.
All this begs the same, much-asked question of what
level of turbulence modeling would be needed for an
accurate prediction of lifting surface flows, or better
put, what consequence one should expect of using
different turbulence models. In this regard, it is
useful to mention the authors previous study of
turbulent flow past a 6:1 spheroid at incidence (Kim
et al., 2002), another example of vortex-dominated
flows, which revealed that the second-moment
closure based on the solution of Reynolds-stress
transport equations yields a remarkably accurate
prediction of the flow, outperforming all the eddyviscosity models
In the present study, we conducted a
computational study on the flow around a simple yet

ABSTRACT
An assessment is made of the fidelity of
computational fluid dynamics (CFD) prediction of a
turbulent wingtip vortex. Efficacy of a featureadaptive local mesh refinement is showcased to
resolve steep gradients in the flow-field along the tipvortex. The impact of turbulence modeling is
evaluated using several popular eddy-viscosity
models and a Reynolds-stress transport model. The
results indicate that, with a judicious combination of
computational mesh with adequate resolution, highorder spatial discretization and advanced turbulence
models, one can predict the tip-vortex flow with a
commendable accuracy.
INTRODUCTION
Turbulent flow around lifting surfaces such
as propeller blades, rudders, and hydrofoils has long
been a topic of both fundamental and practical
interest in naval applications for their ubiquity around
naval vessels and significant impact on various
aspects of hydrodynamic performance. Despite the
long history of deployment and designers vested
interest, the task of numerically predicting turbulent
flows around lifting surface flows remains a difficult
one.
There are several challenges to be dealt with
before one can predict the subject flow accurately.
First, one has to overcome the sheer numerical
difficulty of resolving large gradients of the flowfields in the boundary layer on lifting surfaces and
the tightly braided vortices emanating from the tip.
One difficulty constantly encountered in real
applications is that there are many such regions
needing high-resolution simultaneously. When there
is a need to trace the tip-vortices over a long distance,
as is often the case when tracking trailing vortices in
the wakes of aircrafts, the difficulty is even more
acute.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

generic finite-wing configuration. The flow was


studied in the wind tunnel at NASA Ames (Chow et
al., 1993; Chow et al., 1997) with the focus on the
tip-vortex generated around the wing placed at an
angle of attack of 10 degree, a moderate one that
rules out the possibility of a massive, unsteady
separation.
Finely resolved mean flow and
turbulence measurement shed much light on the
formation and evolution of the tip-vortex, providing a
rare case for CFD validations. There have been very
few numerical studies on this flow (Dacles-Mariani et
al., 1993; Dacles-Mariani et al., 1995; DaclesMariani et al., 1996; Chen, 2000). In the most recent
study (Dacles-Mariani et al. 1996), the one-equation,
eddy-viscosity transport models were used in
conjunction with a fifth-order upwind finite
difference scheme. The authors showed, via a
systematic grid refinement, that the prediction with
the modified Spalart and Allmaras turbulence model
(Dacles-Mariani et al., 1995) on a fine mesh (2.5million cells) closely matched the experimental data,
although the minimum static pressure along the tipvortex core was under-predicted appreciably.
This study is aimed at addressing the two
issues brought up earlier. First, we evaluate the
potential of a feature-adapted local mesh refinement
to efficiently resolve the formation and evolution of
the tip-vortex. A systematic, global mesh refinement
study is conducted in parallel to provide a basis for
comparisons. Secondly, the impact of turbulence
modeling is investigated with three contemporary
eddy-viscosity models and a differential Reynoldsstress model. The results obtained with the popular
two-equation models and the second-moment closure
will complement the earlier numerical studies,
providing a complete picture as to the role of
turbulence modeling.

axial velocity magnitude reaching nearly up to a


maximum of 1.8 times the free-stream velocity,
highly pronounced low (negative) static pressure
peak, and high levels of turbulent stresses, all along
the vortex-core. A kink in the vortex-axis was
observed downstream of the trailing edge. The
measurements also showed that the contours of
turbulent shear stresses have a characteristic pattern
resembling a four-leaf clover tilted by 30 ~ 40
degrees from the vertical and horizontal axes.
NUMERICAL METHOD
All the computations were carried out using
the segregated solver in the FLUENT code. The
flow was assumed incompressible. FLUENT adopts
a cell-centered finite-volume method in conjunction
with a linear reconstruction scheme capable of
handling both structured and unstructured meshes.
Gradients of the solution variables are computed
using Green-Gauss theorem. Diffusion terms are
discretized using second-order central differencing
scheme. For convective terms, there are several
choices offered in the code including first-order
upwind, second-order upwind (SOU), QUICK and
third-order MUSCL schemes. In the present study,
the QUICK scheme was used.
The discretized equations are solved using
pointwise Gauss-Seidel iteration in conjunction with
an algebraic multi-grid method to accelerate the
solution convergence. The details of the numerical
method can be found in the references (Mathur and
Murthy, 1997; Kim et al. 1998; Kim et al., 2003).
TURBULENCE MODELING
Among the numerous eddy-viscosity
turbulence models such as the k- and k- families
and Spalart-Allmaras one-equation model (Spalart
and Allmaras, 1994) most widely used these days, we
had to take a few to keep the scope of the study
tractable. And for the reasons alluded to in the
introduction, we included a second-moment closure
for the study. In the following, we briefly describe
the turbulence models chosen for the present study.

WING MODEL AND EXPERIMENT


The model used in the experiment is a finitewing with 4 ft (chord) x 3 ft (span) rectangular planform, being mounted on the tunnel bottom at a 10o
incidence with the free-stream.
The Reynolds
number based on the chord-length and the freestream
velocity is 4.6 x 106. The free-stream velocity used
in the experiment is about 170 ft/sec, sufficiently low
to be considered incompressible. The boundary layer
is tripped along the leading edge. The wind tunnel
has a 48 in x 32 in cross-section, which makes the
blockage effect non-negligible.
The measurements focused on the near-field
region close to the wing tip and the wake. Velocity,
pressure (static and total), and turbulent stresses were
measured extensively on several cross-flow planes.
The measurements revealed an exceptionally large

Eddy-viscosity models
We selected three most popular eddyviscosity models for this study. In view of its
popularity in the aerospace and the ship
hydrodynamics communities, the one-equation model
of Spalart and Allmaras (1994) (S-A hereafter) was
selected. We adopted the modification proposed by
Dacles-Mariani et al. (1995, 1996) to suppress the
unduly large build-up of eddy-viscosity in the vortex

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

core. The modification employs a modified strainration defined as,


S = + C prod min [0, S ]

in this study. The solution domain and a partial view


of the meshes are shown in Figure 1. The four
different meshes used in this study are summarized in
Table 1. The first three meshes (Mesh I, Mesh II,
and Mesh III) were obtained by globally refining the
coarsest one (Mesh I). The value of y+ in the fine
mesh is around 1.0 at the wall-adjacent cells.

(1)

which is used to compute the production term. In the


above equation, Cprod = 2.0, S and are the moduli
of strain-rate and rotation-rate tensors, respectively.
Among the k- family, the realizable k-
model (RKE hereafter) of Shih et al.(1995) was
chosen for its good track-record for non-equilibrium
boundary layer flows. For the present computations
with fine near-wall mesh resolution, the model originally formulated as a high-Reynolds number
model - was modified to account for the near-wall
effects by employing a zonal hybrid model. Lastly,
the shear-stress transport (SST) k- model (Menter,
1994) was selected to represent the k- family.

Table 1. Meshes used for the computations


Mesh I
Mesh II
Mesh III

# of cells
385K cells
1.04M cells
2.3M cells

Remarks
coarsest mesh
globally refined
globally refined

Mesh IV

1.38M cells

locally refined

In addition, we also employed a locally


adapted mesh (Mesh IV). Our intention was to
evaluate what may be called a feature-adaptive
refinement by which cells are marked for refinement
when the value of a selected variable at the cells falls
in the specified range. The current implementation of
this approach allows one to choose any quantity from
the solution variables (e.g., velocity components,
pressure, k, , etc.) or any derived and user-defined
quantities based on the solutions (e.g., derivatives,
curvatures, vorticity, helicity, total pressure, etc.).
For the present flow, any flow-variable that
can demarcate the tip-vortex region would be a good
candidate. To that end, the second-invariant of
deformation tensor, defined as

Reynolds-stress transport model (RSTM)


The RSTM model used in this study largely
replicates the model used by Gibson and Launder
(1978). The near-wall, low-Reynolds number effects
are modeled using the approach of Launder and
Shima (1989). The details of the model are described
in Kim (2002). The RSTM has been validated for a
number of complex three-dimensional internal and
external flows (Kim, 2001; Kim, 2002).
DOMAIN AND MESH

Q=

(2)

is a conceivable choice, inasmuch as the vortex-core,


where rotation-rate dominates over strain-rate, can be
identified as the region of a positive Q,. The idea of
local refinement based on the second-invariant is
illustrated in Figure 2, which shows the cells marked
for refinement. For this example, the medium mesh
(Mesh II) along with the solution on it was used,
which resulted in a 1.8-million-cell mesh. The figure
shows that the cells marked up for refinement largely
occupies the region around the tip-vortex.
Static pressure could also be an equally
good choice, considering that the tip-vortex carries a
very low static pressure along its core. Figure 3
shows the cells marked up for refinement based on
static pressure coefficient (CP) again using Mesh II.
It can be seen that the cells marked for refinement all
nicely overlap with the region close to the tip-vortex.
The resulting mesh has approximately 1.4 million
cells, much less than what a global refinement would

48"

137.55"

1 2
S2
2

37.3"

55.2"
32"

Figure 1. Solution domain and partial view of the


mesh - Only the bottom surface mesh is shown
here.
In view of the relatively simple geometry,
hexahedral meshes with an O-H topology were used

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

result. In the present study, we used this mesh (Mesh


IV).
Admittedly, one difficulty in this featureadaptive mesh refinement lies with the fact that the
choice of the quantity used to detect the feature is
arbitrary and assumes a considerable knowledge of
the flow in question on the part of the users.
Imaginably, this usability problem becomes more
serious for complex industrial applications in which
the predictions are most likely to be sensitive to the
choice of the variable.
Most ideally, local
refinement or coarsening based on a rigorously
derived error-indicator as adaptation criteria would
offer better alternatives.

BOUNDARY CONDITIONS
Some of the earlier numerical studies
(Dacles-Mariani et al., 1995) used the experimental
data to specify the boundary conditions (Dirichlet)
not only on the inlet but also on the outlet boundary.
However, the practice of using a Dirichlet-type
boundary condition on the outlet - generally not
known a priori - is questionable both mathematically
and practically.
In the present study, the free-stream velocity
was specified at the upstream inlet boundary for the
sake of simplicity, which was deemed justifiable as
our focus is on the region away from the tunnel wall.
The exit boundary is modeled as a pressure outlet on
which an area-averaged static pressure is specified.
The velocity on the pressure outlet is linearly
extrapolated from the adjacent interior cells in such a
way that the overall continuity (mass conservation) is
satisfied. Other solution variables are extrapolated in
a similar manner.
On the tunnel wall and the wing surface, we
adopted a generalized wall-function approach (Fluent
Inc., 2003) that invokes proper wall-laws depending
on the y+ value to provide the wall boundary
conditions for the momentum and the turbulence
equations, thus being applicable to the entire inner
layer including the viscous sub-layer, buffer zone,
and logarithmic layer.
RESULTS

Figure 2. Cells marked for local refinement


based on the second-invariant of deformation
tensor

Mesh-Dependency of the RSTM Results


In this section, we present and discuss the
results obtained using the RSTM on the four meshes.

Figure 4. Mesh-dependency of the RSTM results


for the minimum static pressure along the vortex
core

Cells marked for local refinement


Figure 3.
based on static pressure

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

RSTM. The results exhibit a considerable meshdependency. When viewing these results, however, it
should be kept in mind that the local extrema of the
static pressure and the axial velocity magnitude
would appear by far more sensitive to the mesh
resolution than overall error norms. The coarsest
mesh captures the major qualitative features of the
flow. However, the static pressure and the maximum
axial velocity that are the measures of the strength of
the tip-vortex are severely under-predicted. The
results from the finer meshes show a typical meshconvergence of the solutions, although the
convergence rate is rather slow. The finest mesh,
Mesh III, gives fairly good predictions of all the
aspects of the flow. The static pressure and the
velocity excess along the core of the tip-vortex are
predicted well, although the predictions gradually
deviate from the measurements in the wake. As
shown in Figure 6 and Figure 7, the location of the
vortex-core is closely captured by the predictions.
Interestingly, the sudden change of the vertical
location (z-coordinate) of the vortex axis observed in
the experiment (Chow at al., 1993) - referred to as
kink in the paper - is also seen in the predictions.
The same figures show the results obtained
with the locally adapted mesh (Mesh IV). Despite
the much less number of cells (1.4 million cells in
total) than in the finest mesh (2.3 million cells), the
level of the accuracy is comparable to that of the
finest mesh.

Figure 5. Mesh-dependency of the RSTM results


for the maximum axial velocity along the vortex
core

Figure 6. Mesh-dependency of the RSTM results


for the Y-coordinates of the vortex core

x/c=0.345

x/c=0.985
x/c=0.542 x/c=0.729
x/c=1.191
x/c=0.864

-3.8 -3.6 -3.4 -3.2 -3 -2.8 -2.6 -2.4 -2.2 -2 -1.8 -1.6 -1.4 -1.2 -1 -0.8 -0.6 -0.4 -0.2

(a)

x/c=0.345

x/c=0.985
x/c=0.542 x/c=0.729
x/c=1.191
x/c=0.864

0 0.15 0.3 0.45 0.6 0.75 0.9 1.05 1.2 1.35 1.5 1.65 1.8

Figure 7. Mesh-dependency of the RSTM results


for the Z-coordinates of the vortex core

(b)
Figure 8. Contours of static pressure coefficient
and maximum axial velocity at cross-flow planes
predicted using the RSTM (a) CP (b) Umax/U

Figures 4, 5, 6 and 7 show the negative peak


static pressure coefficient (CP), the maximum axial
velocity magnitude, y-coordinate, and z-coordinate,
respectively, along the vortex core predicted by the

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Figures 8(a) and 8(b) depict the contours of


the static pressure coefficient and the maximum axial
velocity at a selected number of cross-flow planes
predicted by the RSTM with the finest mesh. Figure
9 depicts an overall impression of the corresponding
static pressure distribution on the wing surface. The
characteristic pattern of the surface pressure
distribution such as the low-pressure region on the
suction side near the wing tip - the footprint of the
tip-vortex - is well captured by the prediction.

to that of the primary vortex, as show in red) that


merges with the primary vortex a little behind the
trailing-edge where the kink is observed.

Figure 10. Pathlines in the tip region colored by


the axial component of the vorticity vector
The RSTM offers a unique opportunity to
make direct comparisons of the predicted Reynoldsstresses and the measured ones. It was found that
the predictions largely under-estimate the normal and
shear stresses near the vortex-core at all cross-flow
planes where the measurements were made.
Nonetheless, the overall trends observed in the
experiment are closely reproduced by the predictions.
Figure 11 depicts the contour of vw predicted

(a)

using the RSTM, which shows the characteristic


pattern resembling a four-leaf clover tilted about 45
degrees off the vertical and horizontal axes with
alternating sign, as was found in the experiment.
Evidently, the linear eddy-viscosity models employed
in the study are unable to predict this feature.
(b)
Figure 9. Surface static pressure distribution on
the wing surface predicted by the RSTM on the
finest mesh (Mesh III) (a) pressure side (b) suction
side
The aforementioned kink in the axis of the
vortex-core observed slightly downstream of the
trailing edge has been attributed to the merging of the
primary and the secondary vortices. The pathlines
traced using the RSTM prediction shown in Figure
10 appear to support this hypothesis. In the figure,
the pathlines are colored by the axial component of
vorticity vector. The figure shows the presence of a
secondary vortex (whose sense of rotation is opposite

Figure 11. Contours of the Reynolds shear stress


component, <vw>, distribution at x/c = 1.420

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

It is noteworthy that the four turbulence models give


different rates of the recovery of the static pressure
and the axial velocity in the wake. Overall, the
RSTM gives the slowest recovery of all, which is in
better agreement with the experimental data.
It comes as a little surprise that the
predictions with the RKE and the SST models are
much poorer than the S-A model. All indications
show that the two-equation models, more
sophisticated than the one-equation model, give a
substantially weaker tip-vortex, giving far less
pronounced negative pressure peak, much smaller
velocity excess than the S-A model.
As alluded to earlier, the remarkable success
of the S-A model for this flow is due to the reduced
level of eddy-viscosity in the core made possible by
the modification in Equation (1). It can be easily
seen from the equation that production of eddyviscosity ( P t S ) becomes negative when S/ < 0.5.

Impact of turbulence modeling


The comparisons among the results obtained
using the three eddy-viscosity turbulence models (SA, RKE, SST) and the RSTM are presented in
Figures 12 and Figure 13.
As shown, the RSTM gives the best result.
What is intriguing is the remarkable performance of
the SA model, which is largely consistent with the
finding reported by Dacles-Mariani et al. (1996).
Compared to their S-A result, the present S-A result
compares more favorably with the measurement in
terms of the minimum static pressure at the vortex
core. Their result underpredicted the negative peak
pressure along the vortex core by more than 11 % on
the average, whereas the present S-A result slightly
overpredicts the negative static pressure peak near the
trailing-edge by a few percent, albeit it gives a faster
recovery in the wake.

As a result, turbulent viscosity is destructed along the


core of the tip-vortex, where rotation-rate is much
larger than strain-rate. Figure 14 illustrates this with
an iso-surface of S/ = 0.3, colored by the production
of turbulent viscosity, essentially showing that the
region where S/ is smaller than 0.5 precisely
coincides with the vortex core.

Figure 12. Minimum static pressure along the tipvortex core - predicted by four different
turbulence models

Figure 14. Contour for iso-contour of S/ =0.3,


colored by the production of turbulent viscosity
The RKE model has a similar provision that
contributes to a reduction of the production of
turbulent kinetic energy that is effected by the
variable-C sensitized to rotation. Perhaps, thats
why the RKE model performs better than the SST
model that has no such modification. It should be
noted, however, that all these modifications adopted
to mimic the stabilizing effects of rotation (or
curvature) in the framework of isotropic eddyviscosity are ad hoc. The real physics associated

Figure 13. Maximum axial velocity along the core


of the tip-vortex predicted by four different
turbulence models

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

with rotating flows like the present tip-vortex can be


rigorously represented only by the second-moment
closure, which explains the success of the RSTM
model for the present flow and other similar vortexdominated flows.

refinement seems to have sufficient merits to warrant


further investigation.
As regards turbulence modeling, the secondmoment closure (RSTM) gave the best result, which
was anticipated in view of the salient flow physics
involved in the subject flow that can be rigorously
described by second-moment closures only. Yet, the
modified Spalart-Allmaras (S-A) model also
performed admirably, despite the ad hoc nature of the
modification to mimic the stabilizing effects of
rotation near the vortex-core. It was found that
strong rotation near the vortex core makes the
production of turbulent viscosity negative locally,
ultimately reducing the diffusion and preventing the
vortex from being smeared. Mentors k- model
(SST) and the realizable k- (RKE) model performed
poorly, essentially under-predicting the strength of
the tip-vortex.
It is concluded based on the present study
that tip-vortices, which are one of the most prominent
aspects of lifting surface flows yet have been
perceived difficult to tackle numerically, are tractable
indeed with modern CFD, and can be predicted with
a good accuracy.

SUMMARY AND CONCLUSION


Turbulent flow past a finite-wing has been
studied numerically to assess the capability of
modern computational fluid dynamics to capture the
salient features of lifting surface flows.
The
emphasis was laid upon the level of accuracy with
which the tip-vortex can be captured. One challenge,
which we hoped to address in this paper, lies with the
fact that the flow considered in this paper and lifting
surface flows in general have multiple features with
widely varying length scale. This poses a sheer
numerical challenge in terms of numerically
resolving them. To adequately resolve large solution
gradients near the tip-vortex or any other significant
features requires extremely dense clustering of the
computational cells at different parts of the solution
domain. Another challenge comes from the complex
physics involved in the vortex-dominated flows,
which makes turbulence modeling important and yet
difficult.
We presented in this paper the
computational results obtained using four different
turbulence models including a one-equation, eddyviscosity model (S-A), a couple of two-equation
turbulence models (RKE, SST), and a secondmoment closure (RSTM). The computational mesh
was progressively refined to establish the gridconvergence of the numerical solutions, which
resulted in three globally refined meshes.
In
addition, taking advantage of the capability of our
CFD code to handle unstructured meshes including
hanging-nodes, we evaluated the efficacy of the socalled feature-adaptive local refinement in which
cells are refined when a refinement criterion based on
a user-specified variable is satisfied.
To summarize, with a judicious choice of
computational mesh (either the finest mesh or the
locally refined mesh) and turbulence models (RSTM,
S-A), we were able to predict all aspects of the tip
vortex with a commendable accuracy, such as the
minimum static pressure and the axial velocity excess
at the vortex-core. Particularly interesting was the
finding that the computation with the locally refined
mesh yielded excellent result (in view of its low cell
counts) that closely matches the accuracy with the
finest (globally refined) mesh. Far from being
perfected yet, mainly due to its usability for practical
engineering calculations, the feature-adaptive local

ACKNOWLEDGMENT
The authors acknowledge that FLUENT was
used for all the computations presented here.
REFERNCES
Chen, B., RANS Simulations of Tip Vortex Flows
for a Finite-Span Hydrofoil and a Marine
Propeller, Ph. D. Thesis, Dept of Mechanical
Engineering, University of Iowa, May 2000.
Chow, J. S., Zillac, G. G., and Bradshaw, P.,
Measurements in the Near-field of a Turbulent
Wingtip Vortex, AIAA Paper 93-0551, 1993.
Chow, J., Zillac, G., and Bradshaw, P., Turbulence
Measurements in the Near-field of a Wingtip
Vortex, NASA Technical Memorandum
110418, Febuary 1997.
Dacles-Mariani, Rogers, S., Kwak, D., Zillac, G. G.,
and Chow, J. S., A Computational Study of
Wingtip Vortex Flowfield, AIAA Paper 933010, 1993.
Dacles-Mariani, Zillac, G. G., Chow, J. S, and
Bradshaw, P., A Numerical/Experimental Study
of a Wingtip Vortex in the Near-field, AIAA
Journal, Vol. 33, No. 9., pp. 1561 1568, 1995.
Dacles-Mariani, J., Kwak, D. and Zillac, G.,
Accuracy Assessment of a Wingtip Vortex
Flowfield in the Near-Field Region, AIAA
Paper 96-0208, 1996.
FLUENT Users Guide V6.2, Fluent Inc., 2003.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Gibson, M. M. and Launder, B. E., Ground Eeffects


on Pressure Fluctuations in the Atmospheric
Boundary Layer, Journal of Fluid Mechanics,
Vol. 86, pp. 491 511, 1978.
Kim, S.-E., Mathur, S.R., Murthy, J.Y., and
Choudhury, D., A Reynolds-Averaged NavierStokes Solver Using Unstructured Mesh-Based
Finite-Volume Scheme, AIAA Paper 98-0231,
1998.
Kim, S.-E., Unstructured Mesh Based Reynolds
Stress Transport Modeling of Complex
Turbulent Shear Flows, AIAA Paper 20010728, 2001.
Kim, S.-E., Rhee, S.H., and Cokljat, D., HighIncidence and Dynamic Pitch Maneuvering
Characteristics of a Prolate Spheroid CFD
Validation, 24th Symposium on Naval
Hydrodynamics, Fukuoka, Japan, July 8 13,
2002.
Kim, S.-E., and Rhee, S.H., Assessment of Eight
Turbulence Models for a Three-Dimensional
Boundary Layer Involving Crossflow and
Streamwise Vortices, AIAA Paper 2002-0852,
2002.
Launder, B. E. and Shima, N., Second-Moment
Closure
for
the
Near-Wall
Sublayer:
Development and Application, AIAA journal,
Vol. 27, No. 10, pp, 1319 1325, 1989.
Mathur, S. R. and Murthy, J. Y., A Pressure-Based
Method for Unstructured Meshes, Numerical
Heat Transfer, Vol. 31, pp. 195 - 215
Menter, F. R., Two-Euqation Eddy-Viscosity
Turblence
Models
for
Engineering
Applications, AIAA J., Vol.32, No.8, 1994,
pp.1598-1605.
Spalart, P. R. and Allmaras, S. R., An One-Equation
Model for Aerodynamic Flows, La Recherche
Aerospatiale, No. 1, pp. 5 21., 1994.
Shih, T.-H., Liou, W. W., Shabbir, A., and Zhu, J.,
A New k-e Eddy-Viscosity Model for High
Reynolds Number Turbulent Flows Model
Development and Validation, Computers and
Fluids, Vol. 24, No. 3, pp. 227 238, 1995.
Zillac, G. G, Chow, J. S., Dacles-Mariani, J., and
Bradshaw, P., Turbulent Structure of a Wingtip
Vortex in the Near-field, AIAA Paper 93-3011,
1993.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Stephen Turnock
University of Southampton, UK
Would the authors like to comment on the
accuracy with which the tip vortex strength and by
implication the side force and induced drag of the
control surface are found using the four different
meshes? Also, for the mesh adapted solution what are
the typical number of cells across the vortex core and
did this degrade downstream of the trailing edge?
AUTHORS REPLY
We have not looked at the side force
predictions, since there were no force data to
compare against.
Yet, we believe that the strength of the tipvortex would significantly affect the resultant force,
since the pressure field (e.g., negative peak pressure
at the vortex-core) would impress low static pressure
on the surface of the wing.
Regarding the mesh resolution near the
vortex-core, the adapted mesh (Mesh IV) has roughly
20 cells across the vortexcore just downstream of
the trailing-edge. The mesh resolution near the
vortex-core in the mid- to far-wake does not degrade,
since the vortex core becomes larger and the mesh
retains pretty much the same resolution throughout
the wake.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Michel Visonneau
Ecole Centrale de Nantes, France
The local grid refinement shown in this
paper is controlled by explicit criteria. Dont you
think it would be more rigorous in the future to use a
criteria based on a numerical error estimate? If yes,
what kind of methodology do you want to use to
estimate discritization error on a given grid?
AUTHORS REPLY
The authors fully agree with the comments.
In fact, it was noted in the paper that local
refinement or coarsening of meshes based on
rigorously derived error-indicators as adaptation
criteria would offer better .
Having said that, Id like to add that error
indicator-based refinements for complex industrial
flows may not as efficacious as one might think,
inasmuch as it will be very difficult to come up with
good error-indicators for all equations being solved
and reconcile them.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics


St. Johns, Newfoundland and Labrador, CANADA, 8-13 August 2004

Advanced Design, Analysis, and Testing of


Waterjet Pumps
Thad J. Michael and Christopher J. Chesnakas
(Naval Surface Warfare Center, Carderock Division, USA)
INTRODUCTION
This paper describes the details of waterjet pump
flow as they relate to head rise and efficiency.
Experimental and computational methods and results
are presented.
Waterjet
pumps
have
different
flow
characteristics than conventional and ducted propellers.
An extensive program of testing and computation was
undertaken at the Naval Surface Warfare Center,
Carderock Division (NSWCCD) to learn about these
flow characteristics.
This paper presents a
comprehensive set of measurements and computations
for a mixed flow waterjet pump tested with two
different rotors.
The first part of the paper describes the water jet
pump, including the differences between the two
rotors. The second part of the paper describes the
water tunnel test and compares the flow fields and
performance. The third part of the paper discusses the
difficulties in designing and analyzing arbitrary
waterjet pumps and how calculations were used to
support this test.
OVERVIEW OF THE PUMP
The pump used for these tests was designed by
Rolls-Royce Naval Marine (RRNM), formerly Bird-

Johnson, and is commonly referred to as the AWJ-21.


The AWJ-21 is a mixed-flow pump with a downstream
stator, similar to many commercial pump designs. The
arrangement of this pump is shown in Figure 1. A
smaller model was tested on a combatant hull by
Chesnakas (2001).

Figure 1: AWJ-21 waterjet pump.

A preliminary rotor design was tested using the


same casing and stator at the Massachusetts Institute of
Technology (MIT) by Kimball (2001). Some of this

NOMENCLATURE
CP
D
F
H
KQ
Lref
n
p
Q
r
R

pressure coefficient
inlet diameter
camber
head
torque coefficient
reference length, in this case inlet radius
rotor speed, revolutions per second
pressure
volumetric flow rate
local radius
inlet radius

U
V

velocity, absolute coordinate system


velocity, relative coordinate system
efficiency
flow coefficient = Q/(nD3)
head coefficient = gH/(nD)2
density

Subscripts
r
radial
s
streamwise
t
tangential
x
axial

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

0.1

0.0

Change in C/D

data are presented for comparison.


A new research rotor was designed to produce the
same head rise as the original rotor, but with different
geometry. The research rotor incorporates a different
loading distribution resulting in drastically different
pitch and camber distributions compared to the original
rotor. Measurements with this rotor can be compared
to measurements with the original rotor to identify flow
features that are dependent on rotor geometry.
Figure 2 shows comparative views of the original
and research rotors. Figure 3 shows the difference in
pitch and camber between the two rotors. The pitch
and camber are defined on stream surfaces. Figure 4
shows the substantial reduction in chord at the tip. The
water tunnel measurements will show this drastic
change in rotor geometry effected the performance of
the pump.

-0.1

-0.2

-0.3

-0.4
0.0

0.2

0.4

Span

0.6

0.8

1.0

Figure 4: Difference in chord distribution between rotor


designs.

TEST DESCRIPTION
EXPERIMENTAL APPARATUS
Pump Loop

Figure 2: Comparison of original (left) and research (right)


rotors.
0.04

Pitch

F/D

10

0.02

0.00

-0.02

-5

-0.04

-10

-0.06

-15
0.0

0.2

0.4

Span

0.6

0.8

F/D Difference

Pitch Angle Difference [degrees]

15

-0.08
1.0

Figure 3: Difference in pitch and camber between rotor


designs.

The original rotor was machined from aluminum.


Due to time and cost constraints, the research rotor was
manufactured from plastic using stereo lithography
apparatus (SLA). The lower strength of the research
rotor limited the speed at which it could be tested.
Both high and low speed measurements were made
with the aluminum rotor.

A pump loop was set up to measure the flow


characteristics of the Rolls Royce Naval Marine AWJ21 waterjet in the 36-Inch Water Tunnel at the David
Taylor Model Basin. The tested pump had an inlet
diameter of 191 mm (7.5 inches), with a 6-bladed rotor
in front of an 11-bladed stator as shown in Figure 1.
The rotor was shaft driven from upstream.
The inlet flow area was significantly smaller than
the cross-section of the 914 mm (36 inch) diameter
water tunnel. Due to the large difference in the areas,
the normal pump loop procedure of sending all of the
facility flow through the pump was not feasible.
To test this pump, the unit was placed inside a
pod, mounted on a strut, and placed in the open jet of
the 36-Inch Water Tunnel as shown in Figure 5. With
this arrangement, little control over the pump operating
point could be obtained by varying the speed of the
water in the test section. To test the pump over its full
projected operating range, the exit area of the pod was
adjusted using four interchangeable nozzles. The
nozzles had areas 60, 80, 100, and 120 percent of the
design nozzle area.
The bellmouth inlet and the four nozzles were
manufactured using SLA. This rapid prototyping
process allows complex parts to be constructed
quickly. The strength of the resin is sufficient for
moderate loads. The bellmouth and nozzles were
designed with the aid of numerical computations
described later in this paper.
Static pressure taps within the pod, as well as
total pressure taps at the inlet and exit, allowed the
pump operating curve to be determined. Windows in
the pod allowed optical access for LDV measurements
to be obtained at four planes, shown in Figure 6: in the
parallel inlet section (Station 3), in between the rotor
blades (Station R), in between the rotor and stator

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

(Station 4), and at the nozzle exit (Station 6). The


windows were curved to match the inside profile of the
pump, thus minimizing the flow disturbance. The thin,
0.76mm, windows ensured that the laser beams would
pass through the curved windows with minimal optical
distortion. In order to gain optical access between the
closely spaced rotor and stator at Station 4, it was
necessary to place a spacer in between these elements.
This spacer was placed where the hub and casing walls
were parallel and in the axial direction.

green (514.5 nm) and violet (476 nm) beams of an


argon-ion laser to measure the vertical and axial
components of velocity, respectively. The vertical
probe utilized the blue (488 nm) beam of the argon-ion
laser to measure cross-stream horizontal component of
velocity. Both probe volumes were 0.07 mm by 1.3
mm.
Interchangeable
Nozzles

SLA Inlet

Fairing
Mid-Stage Spacer
with LDV Window
LDV Probes

Figure 7: Pump in pod, mounted on strut, with LDV probes.

Figure 5: Pump test arrangement in the 36-Inch Water


Tunnel at NSWCCD.

The position of the shaft was encoded with an


8192-counts/revolution signal, which was recorded
with each velocity measurement. Doppler signals were
analyzed with a TSI Model IFA 655 Digital Burst
Correlator. The processor performs a 256-sample,
double-clipped, autocorrelation on each Doppler burst,
allowing the measurement of velocity even when the
signal-to-noise ratio is low. The processors were
operated in the random mode.
The flow was seeded with 1500 grit siliconcarbide powder with particles 1 to 2 m in size. Since
the water in the tunnel recirculates, seed particles
needed to be added only infrequently.
MEASUREMENTS
LDV Measurements

Figure 6: Measurement stations.

Figure 7 is a detail of the duct and pump


arrangement, including the location of LDV probes
used for velocity measurements. Pressure taps were
manufactured into the SLA components for static
pressure measurements. A transparent casing around
the rotor allowed cavitation visualization.
LDV System
The LDV system consisted of two TSI model
9832 fiber optic probes rigidly mounted together on a
traverse which could translate in the three spatial
directions. The focal distance of the probes (470 mm
in water) was sufficient to place the probe bodies
outside of the jet. The horizontal probe utilized the

LDV measurements were made at four planes in


the pump, shown in Figure 6. These measurements
provided for a more accurate flow rate measurement
than could be determined with the pressure
measurements, and revealed many details of the pump
flow. Due to the nature of the window access and the
flow conditions, the measurements sets are
fundamentally different at each of the four planes.
At the inlet to the rotor, Station 3, the window in
the pump allowed only two velocity components to be
measured: The axial and tangential. Since the rotor
has only slight influence on the upstream flow, these
measurements are not encoded with the shaft position.
The measurements were along a single radial line and
are presented only as circumferentially averaged
values.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Two component measurements were also made in


the rotor passage, Station R. These measurements
were encoded with the shaft position so that spatial
analysis of the flow in the rotating frame of the rotor
could be performed. The measurements were grouped
into 1024 circumferential positions, each 8 encoder
counts wide. The measurements were also taken only
on a single radial line, but because of the encoding of
the rotor position, a plane of measurements in the
rotating frame of the rotor was obtained.
In between the rotor and the stator, Station 4,
three component measurements were made which were
encoded with the shaft position.
Again the
measurements were grouped into 1024 circumferential
positions, and a plane of measurements in the rotating
frame of the rotor were obtained.
The final plane of velocity measurements was just
downstream of the nozzle exit, Station 6. At this
location, all three velocity components were measured,
but since the measurements were primarily influenced
by the stator, no shaft encoding was employed.
Instead, a two dimensional grid of measurements
across the nozzle exit was measured, so that the
velocity field in the stationary frame was obtained.
Additional details of the LDV system used are
described by Jessup, et al. (2004).

velocities with a superscript asterisk, which are


normalized by the relative inflow velocity
2

V = U x + ( 2rn ) 2

Measurements of the circumferentially averaged


velocities are presented in the stationary coordinate
system, while the velocity planes are presented in the
rotating coordinate system of the blades.
In order to better illustrate the features of the
flow, the velocities in the rotating frame measurement
planes are rotated into the primary-secondary
coordinate system shown in Figure 8.
In this
coordinate system, a primary, or streamwise, direction
s is defined as being in the axial-tangential plane, at the
circumferentially averaged swirl angle, . Since the
swirl angle is different at each radius, the coordinate
system is different at each radius as well. The
secondary velocities are then the orthogonal velocity
component in the x-t plane, Vc, and the radial velocity,
Vr. These velocity components can be calculated from:
Vs =

Flow Visualization
Two types of flow visualization were employed.
Photos of fully developed cavitation were obtained in
order to reveal regions of low pressure in the flow and
the gross structure of the vortical flow. Surface oil
flow visualization was also employed to reveal the
presence of vortices along the rotor and stator surfaces
and the location of flow separations.

sin V x + cos Vt

Vc = cos V x + sin Vt

(2)

Vr = Vr

Pressure Measurements
Rings of pressure taps were located at Station 3,
Station 4, and just ahead of Station 6, to measure the
circumferentially averaged wall static pressure.
Additionally, measurements were made of the test
section static pressure, the tunnel inflow total pressure,
and the pump nozzle total pressure. The nozzle total
pressure was obtained with a single Kiel probe
measuring at approximately 75% of the nozzle radius.
These pressure measurements, when combined with the
LDV velocity measurements, allowed the pump
operating curve to be determined.

(1)

t
Vx

Vt

90-

Figure 8: Primary-secondary flow coordinate system for


LDV measurements.

The flow vector in the stationary frame, U, and


the flow vector in the rotating frame, V, are related by
Vx = Ux ,

Vt = Ut - 2rn ,

Vr = Ur

(3)

TEST RESULTS

STATOR

COORDINATE SYSTEM
All distances presented here are normalized by
the inlet radius, R = 95.25 mm (3.750 inches). The
velocity measurements presented here are generally
normalized by nD, where n is the rotor rotation rate,
and D is the inlet diameter. The exceptions to this are

Figure 9 shows an axial velocity survey at the


nozzle exit with the original rotor. A hub vortex is
clearly visible. Stator wakes can also be seen.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Ux

0.2

2.40
2.30
2.20
2.00
1.80
1.60
1.40
1.20
1.00
0.80
0.60
0.40
0.20
0.00

0.1

-0.1

-0.2

-0.3

-0.4

-0.5

0.2

-0.2

-0.4

Figure 9: Axial velocity contours and secondary flow


vectors at the nozzle exit, original rotor, looking upstream,
=0.92.

The hub vortex is also visible in the oil-paint flow


visualization shown in Figure 10. This figure shows
how flow separates on the suction side of the stator
blades and forms the hub vortex.
ROTOR
Figure 11 shows the flow between the rotor and
stator for both rotor designs. This view shows
velocities in the rotating coordinate system in the
streamwise direction, looking upstream. Blade wakes
are clearly visible.
Note the vortex close to the tip on the pressure
side of the blade in Figure 11. The leakage vortex is
formed by flow on the pressure side of the blade
crossing through the tip gap to the suction side of the
blade. With typical ducted propellers, the leakage
vortex remains near the suction side (Chesnakas and
Jessup, 2003). Figure 12 shows how the leakage
vortex in the original rotor forms near the leading edge
and crosses the blade passage. This phenomenon was
also photographed at MIT (Kimball, 2001). Cavitation
visualization was not possible with the research rotor
because of the strength of the material.
Figure 13 shows the oil paint flow visualization
on the pressure side of the original rotor. The line near
90 percent span is hypothesized to be caused by the
leakage vortex. Figure 14 shows the same view of the
research rotor, also with a line near 90 percent span. In
both figures, the low energy flow near the hub can be
seen moving outward as it flows over the chord.
PUMP CURVES
The energy flux at the inlet, Station 3, was
determined by assuming that the static pressure was
constant across Station 3, and integrating the product of
velocity and total pressure across the flow area. The
mass flux was determined by integrating the axial
velocity at Station 3.
The static pressure at the nozzle discharge,
Station 6, was measured on the inside wall of the
nozzle. Since the flow at the nozzle exit has only a
small radial component, the pressure field across the
nozzle area was calculated by integrating
U2
dp
= t
dr
r

Figure 10: Flow visualization on stator, looking upstream.

(4)

from the edge of the nozzle flow to the center, r =


0. The circumferentially-averaged tangential velocity
was used, and the pressure was assumed to vary only
with r. The total pressure and axial velocity were then
integrated to determine the energy flux at the
discharge.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Vs
0.80
0.75
0.70
0.65
0.60
0.55
0.50

Figure 12: Water jet pump rotor tip gap vortex cavitation.

V*s
0.80
0.75
0.70
0.65
0.60
0.55
0.50

Vs

Figure 13: Flow visualization on pressure side of original


rotor.

0.80
0.75
0.70
0.65
0.60
0.55
0.50

Figure 11: Stream-wise velocity contours and secondary


flow streamlines at Station 4, =0.92, looking upstream.
Top: original rotor, high speed. Middle: original rotor, low
speed. Bottom: research rotor, low speed.

Figure 14: Flow visualization on pressure side of research


rotor.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

1.2
1.0

/ max, / max, KQ

Low aspect ratio blades


Blade tips interacting with thick boundary layers
1.2
1.0

/ max, / max, KQ

The mass flux and energy flux were used to


compute pump curves, shown in Figure 15. The head
rise coefficient, , represents the increase in energy
between the inlet and the nozzle. Low Reynolds
number data is shown due to speed limitations with the
plastic research rotor.

0.8
0.6
0.4
0.2
0.0
0.7

Original Rotor /max


Original Rotor /max
Original Rotor KQ
Research Rotor /max
Research Rotor /max
Research Rotor KQ

0.8

0.8
0.6
0.4
0.2

NSWCCD /max
NSWCCD /max
NSWCCD KQ
MIT /max
MIT /max
MIT KQ

0.0
0.7

0.8

0.9

1.0

1.1

0.9

1.0

1.1

Figure 16: Comparison of pump performance at high


Reynolds number with the original rotor at NSWCCD and
with a preliminary rotor design at MIT (Kimball, 2001).
Both and are normalized by the highest value measured
with the original rotor at high Reynolds number.

Figure 15: Pump performance with both rotors at low


Reynolds number. Both and are normalized by the
highest value measured with the original rotor at high
Reynolds number.

COMPUTATIONAL METHODS

Computations were made using a lifting surface


blade row code, PBD-14, coupled with an
axisymmetric Euler solver, MTFLOW (Drela and
Giles, 1983, Kerwin, et. al., 1994, Renick, 1999).
These codes are suitable for both design and analysis.
PBD-14 can determine the blade pitch and camber
required for a specified loading distribution or can
determine the loading from a specified geometry. A
solution can be computed in a matter of minutes on a
typical personal computer.
These codes are able to accurately predict the
performance of an open propeller (Renick, 1999).
They have also been used successfully for a compound
ducted unit (Michael, Jessup, and Scherer, 2002).
However, waterjet pumps have features that make them
substantially different:
The long casing, which expands and contracts for
mixed flow waterjets

1.0
0.8

r/R

Figure 16 compares the NSWCCD high Reynolds


number data for the original rotor with MITs
measurements with a preliminary rotor design. A
different loading distribution on the rotor may have
lead to higher torque and head rise. Some differences
may be due to the difference in the inflow, shown in
Figure 17.

1.2

0.6
0.4
36-Inch Water Tunnel
MIT Water Tunnel

0.2
0.0
0.0

0.2

0.4

0.6

U_X

0.8

1.0

1.2

Figure 17: Pump inflow axial velocities NSWCCD and MIT


(Kimball, 2001) water tunnels. The mass flow rate is the
same.

For this project, the water jet pump was modeled


in two configurations. Prior to the testing, the flow
inside and outside of the duct was modeled, as shown
in Figure 18. Once model test data was obtained, a
smaller domain, limited to the inside of the duct, was
used with the inflow velocity distribution measured in
the test. Figure 19 shows the measured inflow and the
inflow used in the computation, both at Station 3.
LOSSES IN WATERJET PUMPS

The three primary sources of head loss in waterjet


pumps are:
1. Blade section drag;
2. Tip losses;
3. Casing losses.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

r/L ref

4
3
2
1
0

-10 -9

-8

-7

-6

-5

-4

-3

-2

-1

x/Lref
CP: -1.0 -0.9 -0.8 -0.7 -0.6 -0.5 -0.4 -0.3 -0.2 -0.1 -0.0 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

Figure 18: Computational domain including duct. Pressure


contours shown.

MTFLOW includes a boundary layer solver


which could be used on the hub and casing, but the
displacement thickness can create problems for the
coupling between the codes and therefore the
MTFLOW boundary layer solver was not used.
As the hub and casing boundary layers enter the
rotor, the bending of the flow through the blade rows
causes a passage vortex to form. This type of flow is
most visible in Figure 11 for the research rotor. The
amount of energy in this vortex is generally small, and
does not contribute significantly to the head loss.
DESIGN OF DUCT AND NOZZLES FOR
MODEL TEST

1.2
1.0

r/R

0.8
0.6
0.4
Computation
Measurement

0.2
0.0
0.0

0.2

0.4

0.6

U_X

0.8

1.0

1.2

Figure 19: Measured inflow at Station 3 and matched


computation.

Of these three, only the blade section drag has


been directly accounted for in the past. (Duct losses
were accounted for with a duct drag coefficient.)
Without these losses, head rise, and therefore
efficiency, are over predicted. A project is currently
investigating adding the other two types of losses.
PBD-14 utilizes drag coefficients for modeling
section drag. The drag coefficients used in this work
were computed based on Reynolds number and include
corrections for section thickness, blockage, lift
coefficient, and cascade diffusion. Kimball (2001)
calculated rotor drag coefficients from LDV data.
Using these drag coefficients he obtained good
correlation with the measurements in the MIT water
tunnel. Kimballs measurements showed very large
drag coefficients at the tip, due to the inclusion of the
tip leakage vortex. For a general method applicable to
arbitrary pump geometries, a separate treatment of each
type of loss is desirable.
Tip gap losses can be broken into two
components, a potential effect and a viscous effect.
The potential effect of leakage flow through a tip gap
induces velocities in the blade rows. These velocities
create an induced drag on the rotor and reduce the
loading on the rotor. The viscous effect is a mixing
loss, or pressure drop, due to the leakage flow from the
pressure side mixing with the faster flow on the suction
side. This type of flow is shown in Figure 11.

Coupled PBD-14 and MTFLOW calculations


were used in designing the inlet bellmouth and
selecting nozzle sizes for the model test. Although the
codes could not accurately predict the pump head rise
without a more complete loss model, it was possible to
estimate the range of nozzle sizes needed to cover the
desired range of flow coefficients and compare
bellmouth designs.
Figure 20 shows the evolution of bellmouth
designs. In the first figure, the stagnation point is well
inside of the duct. There is also a stagnant area on the
outside of the duct, beyond the bellmouth. In the
second figure, the stagnation point has been moved by
reshaping the bellmouth. In the third figure, the
stagnant area on the outside of the duct has also been
eliminated by a small adjustment to the geometry.
Calculations were also used to evaluate the effect
of nozzle size changes on the flow through the duct and
the stagnation point. Figure 21 shows the predicted
change in stagnation point location between the 60%
nozzle and the 120% nozzle flow conditions.
The stagnation point was kept on the inside of the
duct in all conditions to avoid possible separation on
the inside of the duct due to the low pressures that
would occur if the stagnation point were on the outer
part of the lip.
DESIGN OF THE RESEARCH ROTOR

The design of the second rotor was based on the


computed performance of the first rotor. Although the
computer codes in use were unable to accurately model
the losses in the waterjet pump, it was believed that the
relative performance of the pumps could be predicted
accurately.
The objective for the design of the research rotor
was to produce the same head rise as the original rotor
using substantially different geometry and a different
loading distribution. The extent to which the loading
distribution could be altered was limited by the short

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

span of the blades. Figure 22 compares the computed


loading distributions for the two rotors.
2.5

CP
0.9
0.7
0.5
0.3
0.1
-0.0
-0.2
-0.4
-0.6
-0.8
-1.0

1.5
1.0
0.5
0.0-6

-5.5

-5

-4.5

-4

-3.5

-3

x/Lref

-2.5

-2

-1.5

-1

-0.5

0.150
CP
0.9
0.7
0.5
0.3
0.1
-0.0
-0.2
-0.4
-0.6
-0.8
-1.0

2.0

r/Lref

Original Rotor
Research Rotor

0.155

2.5

1.5
1.0
0.5

x/Lref

2.5

CP
0.9
0.7
0.5
0.3
0.1
-0.0
-0.2
-0.4
-0.6
-0.8
-1.0

2.0
1.5
1.0
0.5

0.145
0.140
0.135
0.130
0.125

0.0-6.0 -5.5 -5.0 -4.5 -4.0 -3.5 -3.0 -2.5 -2.0 -1.5 -1.0 -0.5 -0.0

r/Lref

0.160

Circulation (G)

r/Lref

2.0

shows that the PBD-14/MTFLOW method is practical


for rapid evaluation of the relative performance of
pumps with arbitrary geometry in cases where the
losses that are not included in the calculation are
similar. Clearly the accurate prediction of these losses
will make the method useful for many more cases.

0.120
0.0

Figure 22:
distributions.

0.2

0.4

Span

Comparison

of

0.6

0.8

calculated

1.0

circulation

COMPARISON OF CALCULATIONS AND TEST


DATA

0.0-6.0 -5.5 -5.0 -4.5 -4.0 -3.5 -3.0 -2.5 -2.0 -1.5 -1.0 -0.5 -0.0

x/Lref

Figure 20: Evolution of the bellmouth design.


2.5

CP
0.9
0.7
0.5
0.3
0.1
-0.0
-0.2
-0.4
-0.6
-0.8
-1.0

r/Lref

2.0
1.5
1.0
0.5
0.0-6.0 -5.5 -5.0 -4.5 -4.0 -3.5 -3.0 -2.5 -2.0 -1.5 -1.0 -0.5 -0.0

x/Lref

2.5

CP
0.9
0.7
0.5
0.3
0.1
-0.0
-0.2
-0.4
-0.6
-0.8
-1.0

r/Lref

2.0
1.5
1.0
0.5
0.0-6.0 -5.5 -5.0 -4.5 -4.0 -3.5 -3.0 -2.5 -2.0 -1.5 -1.0 -0.5 -0.0

x/Lref

Figure 21: Effect of small nozzle (top) and large nozzle


(bottom) on leading edge stagnation point.

The design was begun by running PBD-14 in


design mode to establish chordwise camber and initial
spanwise pitch and camber distributions. The output
geometry was faired and adjusted until matching
analysis calculations of both rotors predicted the same
head rise.
As the model test data in Figure 15 shows, the
performance of the rotors was well matched. In order
to meet test schedule requirements, the research rotor
design was completed in only three days, with the
hardware delivered another three days later. This work

With an incomplete loss model, the computed


pump performance can be expected to be better than
the actual performance. Including tip losses and casing
losses can be expected to slightly increase the predicted
torque and substantially reduce the predicted head rise,
thereby reducing the predicted efficiency substantially.
Figure 23 compares the computed and measured
head rise, rotor torque, and efficiency for a range of
flow coefficients. The computed torque is relatively
close, departing from the measurements at higher flow
coefficients. The computed head rise is larger than the
measured head rise, as expected. This results in an
efficiency prediction that is unreasonably high.
Figure 24 shows the measured and calculated
axial velocity distribution at Station 4, between the
rotor and stator, for a flow coefficient of 0.92. The
lack of a tip loss model and appropriate casing losses is
evident. The computed velocities near the casing are
significantly higher than the measured velocities. The
addition of a tip loss model and a casing loss model
will cause a redistribution of the flow, with lower
velocities at the casing resulting in increased velocities
near the hub.
Figure 25 shows the measured and calculated
tangential velocity distribution at station 4, between the
rotor and stator, for a flow coefficient of 0.92.
Modeling the tip leakage flow will induce tangential
velocity on the blade surface in the direction of
rotation, reducing the load on the propeller and the
mean downstream tangential velocity.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

it should be possible to better match the discharge


velocities.
Figure 27 shows the measured and calculated
tangential velocity distributions at the discharge. Both
the measurements and calculations show a hub vortex
in the opposite direction of the rotor rotation. That is,
the stator over-turns the incoming swirl at the root.
The measurements show better swirl elimination over
most of the span than the calculations indicate. In
either case, the tangential velocities are generally less
than 10% of the jet velocity.

1.2

/ max, / max, KQ

1.0
0.8
0.6
0.4

/max Measured
/max Measured
KQ Measured
/max Computed
/max Computed
KQ Computed

0.2
0.0
0.7

0.8

1.0

0.9

1.0

Computed
Measured

1.1

0.8

0.6

r/R

Figure 23: Measured and calculated pump performance at


high Reynolds number. Both and are normalized by the
highest value measured with the original rotor.

0.4
1.4
0.2

Computed
Measured

1.2

r/R

0.0
0.0

0.5

1.0

1.5

2.0

UX

2.5

3.0

Figure 26:
Measured and calculated axial velocity
distribution at nozzle exit, Station 6, =0.92.

1.0

1.0

0.8

Computed
Measured

0.8
0.2

0.4

0.6

UX

0.8

1.0

1.2

1.4

Figure 24:
Measured and calculated axial velocity
distribution between rotor and stator, Station 4, =0.92.

0.6

r/R

0.6
0.0

0.4
1.4
0.2

Computed
Measured

1.2

r/R

0.0
-1.0

-0.8

-0.6

-0.4

-0.2

0.0

UT

0.2

0.4

0.6

0.8

1.0

1.0

Figure 27: Measured and calculated tangential velocity


distribution at nozzle exit, Station 6, =0.92.

0.8

CONCLUSIONS

0.6
0.0

0.2

0.4

0.6

UT

0.8

1.0

1.2

1.4

Figure 25: Measured and calculated tangential velocity


distribution between rotor and stator, Station 4, =0.92.

Figure 26 shows the measured and calculated


axial velocity distributions at Station 6, the nozzle
discharge. The calculations have too much flow at the
casing, and too little flow near mid-span. These
problems are the result of the lack of tip loss and casing
loss modeling. With the inclusion of appropriate losses

The design, analysis, and testing of water jet


pumps is a complicated problem. It is significantly
more difficult than for open propellers.
Determining the efficiency and operating points
of an open propeller is a simple matter of varying the
rpm while measuring the thrust and torque. For
waterjet pumps, a more complicated arrangement may
be needed to control the operating point and a velocity
survey, or at least a Kiel probe survey, is necessary to
accurately calculate the head rise. A velocity survey
has the advantage of measuring swirl.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

The leakage vortex trajectory was photographed


and the structure was accurately measured using 3component LDV.
These test data show that the geometry of the
rotor does not have as large an effect on the overall
pump performance as one might expect. Water jet
pumps differ from open and ducted propellers because
of the substantial losses related to the casing and low
aspect ratio blades.
It is hoped that this high quality data set can be
used to develop appropriate loss models so that pumps
of arbitrary geometry can be designed and analyzed
using rapid methods, such as PBD-14 coupled with
MTFLOW. Until that time, the current method has
been found to be excellent for substantial modifications
to configurations with existing test data.
The inclusion of tip and casing loss modeling is
currently being investigated, and it is the authors hope
that the result of that work can be presented in the
future.

Ducted Propulsors Using Generalized Geometry,


SNAME Transactions, 1994.
Kimball, R. W., Experimental Investigations and
Numerical Modelling of a Mixed Flow Marine
Waterjet, PhD. Thesis, Ocean Engineering
Department, Massachusetts Institute of Technology,
June 2001.
Michael, T.J., Jessup, S. D., and Scherer, J. O.,
AHFID
Propulsor
Performance
Prediction,
NSWCCD
Report
NSWCCD-50-TR-2002/003,
January 2002.
Renick, D. H., An Analysis Procedure for Advanced
Propulsor Design, Masters Thesis, Ocean Engineering
Department, Massachusetts Institute of Technology,
May 1999.

ACKNOWLEDGEMENTS

The authors would like to thank Dr. John


Barkyoumb for supporting this work through the
NSWCCD IAR program, sponsored by the US Navys
Office of Naval Research.
The authors thank Dr. Stuart Jessup for the
impetus to write this paper and thank J. Otto Scherer
for his assistance in understanding waterjet
terminology.
REFERENCES

Chesnakas, C. J., 3-D LDV Mapping of the Flow


About a Waterjet-Powered Hull in a Tow Tank, 26th
American Towing Tank Conference, Webb Institute,
Glen Cove, NY, 23-24 July 2001.
Chesnakas, C. J. and Jessup, S. D., Tip-Vortex
Induced Cavitation On a Ducted Propulsor,
Proceedings of ASME FEDSM03, Honolulu, Hawaii,
6-10 July 2003.
Drela, M. and Giles, M., Conservative Streamtube
Solution of Steady-State Euler Equations, Technical
Report CFDL-TR-83-6, Department of Aeronautics
and Astronautics, Massachusetts Institute of
Technology, November 1983.
Jessup, S. D., et al., Propeller Performance at Extreme
Off Design Conditions, 25th Symposium on Naval
Hydrodynamics, St. Johns, Newfoundland and
Labrador, Canada, 8-13 August 2004.
Kerwin, J. E., et al., A Coupled Viscous/Potential
Flow Design Method for Wake-Adapted Multi-Stage,

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Justin E. Kerwin
Massachusetts
Institute
of
Technology,
USA/Consultant, Anteon Corporation, USA
The authors have provided an extremely
valuable set of experimental data and comparisons with
calculations for a mixed-flow waterjet pump. Having
been involved with both calculations and experiments
for the AWJ-21 and earlier designs, I am all the more
impressed with the quality of the authors work.
Testing the waterjet pump in a strut mounted
pod was, no doubt, complicated to arrange, but results
in a much cleaner setup to compare with analytical
predictions. The setup in the MIT water tunnel, of
necessity, included a large upstream drive and a
bellmouth extending into the main contraction section.
As shown in Figure 17, the authors setup produces a
much more uniform inflow profile. Of course, real
waterjets may have even more non-uniform inflow, but
for research purposes, it is best to deal with one
problem at a time.
It is interesting to note that the experimentally
determined pump performance of the original rotor and
the research rotor is almost identical, in spite of the
significant differences in blade planform shape (and
resulting differences in pitch and camber). The authors
state that the difference in design loading distribution
results in large differences in pitch and camber.
However, the spanwise distributions of circulation
shown in Figure 22 look almost identical. The
distributions are both practically linear, and differ from
each other at the hub and tip by no more that three
percent. This would suggest that the differences in pitch
and camber are due principally to differences in
planform shape.
It would also indicate that overall waterjet
performance (for a given passage geometry and flow
coefficient) is principally a function of the radial
distribution of circulation and not detailed blade shape.
This is, of course, the justification for lifting-line
theory, which is well established for traditional open
propellers. It would be nice if this were true for
waterjets, too.
Anticipating this finding, I have been
developing a lifting-line equivalent of the PBD-14/
MTFLOW method used by the authors with the hope
that it will be an even simpler and faster method for
carrying out preliminary designs (Kerwin, 2003).
Both lifting-surface and lifting-line methods
will over-predict efficiency unless the types of losses
enumerated by the authors are incorporated in some
way. The experimental results presented here will be
most useful in developing the needed loss models.

In particular, Figure 11 nicely captures the


leakage vortex. Could the authors extract a numerical
estimate of the vortex strength from the raw data and
compare it with the calculated circulation in Figure 22,
which is based on no tip leakage? This number,
together with the size of the tip gap and the section
thickness at the tip would be very useful.
Figure 27 shows that tangential velocities exist
near the axis at the nozzle exit, and the presence of a
hub vortex is confirmed by the flow visualization
results shown in Figure 10. But, this is present (and
nearly equal) in both in the calculated and experimental
results. Am I right in assuming that the designed
circulation distributions on the rotor and stator were
initially set to cancel the tangential velocity? If this is
true, then the presence of calculated tangential velocity
must be the result of fairing out the (design) kinks in
either the rotor or stator blade geometry close to the
hub. The authors comments would be welcome.
Finally, the flow visualization photograph of
the original rotor shown in Figure 13 shows very well
defined traces, while the corresponding photograph of
the research rotor shown in Figure 14 has a much more
bubbly appearance. Is this simply due to a difference
in lighting or does it mean that the flow is different in
some way?
REFERENCES
Kerwin, J.E., The Preliminary Design of Advanced
Propulsors, Propellers/Shafting 03 Symposium,
Virginia Beach, VA, September 16-17, 2003.
DISCUSSION
Dr. Richard Kimball
Massachusetts Institute of Technology, USA
Congratulations to the authors on their work
and their advancement of knowledge in the area of
waterjet pump design. This work adds insight into the
loss mechanisms encountered in waterjet rotor design.
Specifically, the large vortex structures generated near
the rotor tip and duct casing have been shown to be
responsible for much of the pump energy loss. As
pointed out in their paper, little in known on how to
model these losses in the current waterjet design
procedures.
The measured velocity contours of Figure 11
are especially useful, showing the structure and
magnitude of the rotor tip vortex structures. In the work
of Kimball(2001) such detail was not possible since
only axial and tangential velocity profiles were
collected. In the present work, since all three velocity
components were measured simultaneously, giving an

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

accurate phase-averaged depiction of the flow field


exiting the rotor. Such data can give better quantitative
understanding of the vortex structures including their
strengths, core sizes, trajectories, etc.
The inflow velocity profiles of Figure 17,
shown comparing the NSWCCD and MIT tests, show
markedly different inflow profiles. In the MIT test the
inflow duct was very long which gives rise to the thick
duct boundary layer. The NSWCCD duct was relatively
short, hence thinner developed boundary layers. I
suspect that a conventional waterjet inlet duct would
fall somewhere between these data.
In this paper, the tip clearance from the rotor
to the duct was not noted. However, from Kimball
(2001) the effect of the tip clearance dramatically
affected the pump performance. This data is reproduced
below in Exhibit 1 and shows for a change in tip
clearance to diameter ratio of less than one percent can
reduce the pump headrise by more than 10%. As noted
by the authors the tip leakage loss can be broken into a
potential effect and a viscous effect. Using the tip gap
orifice model by Kerwin (1989), much of the potential
effects of the tip gap can be modeled in the potential
flow design procedure. However, the viscous losses of
the tip gap are still underpredicted using the orifice
model.

1.5

Tested at constant RPM


6
Reinlet=0.88x10

authors is of high quality and is clearly an advancement


toward improved waterjet design capability.
REFERENCES
Kerwin, J.E., A method of treating ducted
propeller tip gap flows in a potential flow panel code.
Technical Report (Internal Report), Massachusetts
Institute of Technology, Cambridge, Ma. October 1989
Kimball, R.W., Experimental Investigations and
Numerical Modelling of a Mixed Flow Marine
Waterjet. Ph.D. Thesis, Ocean Engineering Department,
Massachusetts Institute of Technology, June 2001.

AUTHORS RESPONSE
The authors would like to thank the
discussers for their valuable comments.
The relatively uniform inflow in the water
tunnel was quite different from the inflow in a real
waterjet installation. Figure R-1 shows a measured
inflow at model scale for a typical waterjet inlet.

1.2
1.1

P tot/P tot(design)

0.9
0.8

0.5

0.7

P tot at tipgap t/D i=0.0013


P tot at tipgap t/D i=0.0025
P tot at tipgap t/D i=0.004
P tot at tipgap =t/D i=0.0065
Torq. at tipgap t/D i=0.0025
Torq. at tipgap t/D i=0.004
Torq. at tipgap t/D i=0.0065

0.6

Torq/TorqDesign

0.5
0.4
0.3

0
0.9

0.95

1.05

Flow/Flowdesign

1.1
9/22/00 rwk

Exhibit 1: Effect of Tip gap on Waterjet pump


performance (reproduced from Kimball(2001))

In closing, I believe measurements such as


those presented in this work are necessary to develop
efficient and robust models of waterjet losses. Also,
through better understanding of these losses, it is
possible that the efficiency of waterjet pumps could be
improved significantly. The work presented by the

Figure R-1: Experimentally measured axial velocity


contours at the pump inlet flange.

Professor Kerwin is probably right in


suggesting that the differences in pitch and camber
are mostly due to planform. In addition to changing
the chord distribution, the location of the blades has
been changed. Figure R-2 shows a side view of the
pump, with the original and research rotors shaded in
different colors. Clearly, the research rotor is located
differently on the hub, which can be expected to
cause differences in the pitch and camber at the root.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Original Rotor
Research Rotor

Figure R-2: Comparison of rotor blade locations..

For the original rotor, the tip gap was set at


0.533 mm (0.021 in) at rest and the tip section is 5.6
mm (0.22 in) thick. Unfortunately, the strength of
the leakage vortex cannot be accurately estimated
from these measurements. The vortex crosses the
measurement plane at a low angle and seems to have
diffused
substantially
before
reaching
the
measurement
The authors were not involved in the design
of the stator, and therefore cannot comment on the
intent of the design. However, fairing the research
rotor was difficult. The non-cylindrical geometry,
particularly at the root, results in S shaped
chordwise camber. Where this transitions to a more
conventional chordwise camber, it tends to form a
bump, seen on the research rotor. With more time
available, it may have been possible to fair this out
better. Ultimately, the design concludes with a series
of analysis calculations and small adjustments to the
geometry to achieve the desired operating point.
The difference in the fineness of the flow
visualization traces between Figure 13 and Figure 14
is due to the rotor speed. Because of the plastic
construction, the research rotor was limited to 900
RPM, while the original rotor was tested at 2000
RPM. The lower speed resulted in lower shear on the
oil and thus a thicker film is visible in the photos
following the lower speed test.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics


St. Johns, Newfoundland and Labrador, CANADA, 8-13 August 2004

Propeller Performance at Extreme Off Design Conditions


By

Stuart Jessup, Chris Chesnakas, David Fry, Martin Donnelly, Scott Black and Joel
Park
(Naval Surface Warfare Center, Carderock Division)
observations show very erratic, unsteady leading edge
cavitation, which resembles classic leading edge stall
(Figure 1). Using LDV and PIV measurements in the
X-R plane, the recirculation ring vortex is documented.
Instantaneous PIV images show the erratic movement
of the ring vortex in and out of the propeller disk.
When the vortex moves inside of the propeller tip, it is
hypothesized that the blade loading increases to peak
levels due to extreme blade section angle of attack and
results in leading edge stall. Effort is made to predict
the peak blade loading from PIV derived propeller
inflow. An ad hoc approach is proposed using
measured blade section angles of attack, and lifting
surface corrections to compute peak blade loading for
the assessment of blade structural integrity.

ABSTRACT
Propeller performance at extreme off design
conditions was investigated using experimental
techniques. Experiments were conducted in the
NSWCCD 36WT for a simple open propeller at near
bollard and crashback conditions. Tests included LDV,
PIV, load measurement, cavitation observations and
laser sheet flow visualization. Numerical calculations
were performed using panel methods to predict the
performance.
At near bollard conditions the blade flow was
attached, showing an expected thick suction side
boundary layer. Performance was predicted with
potential flow methods, with empirical corrections for
boundary layer effects.

NOMENCLATURE
C
Cp
Cd
d
D
F
G
J
KT
KQ
n
P
p0
q
r
R
T
U
Ut

Figure 1 Propeller 4381 cavitating during


crashback conditions
At crashback conditions the blade flow produced
unsteady leading edge stall conditions. Cavitation

Blade section chord


Blade surface pressure coefficient, based on U
Blade section Drag coefficient
Blade wake width, normal to streamwise
direction, normalized on R
Propeller Diameter
Blade section camber
Spanwise circulation
Advance coefficient, U/nD
Thrust coefficient, T/n2D4
Torque coefficient, Q/n2D5
Propeller rotation rate, rps
Blade pitch
Tunnel static pressure
Turbulent kinetic energy
Local radius
Propeller tip radius
Blade section Thickness
Tunnel velocity
Tangential velocity, inertial frame

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Vs
Vc
V r, U r
Vt
Ux
X
Y
Z

Streamwise velocity, rotating frame


Orthogonal velocity, rotating frame
Radial velocity
Tangential velocity , rotating frame
axial velocity, inertial frame
Axial distance from Propeller centerline,
positive downstream
Vertical distance from propeller centerline
Horizontal distance from propeller centerline
blade section flow angle, tan-1(Ux / (r +Ut))
Standard deviation of measured Thrust
Cavitation number, based on U, and p0
Propeller angular velocity
Propeller angular position

The backing condition is one of the most complex


and challenging propeller operating conditions to
analyze. It also results in the most extreme loading
conditions, often determining the strength of the
propeller.
The crashback condition is dominated by the
interaction of the free stream flow field with strong
recirculation driven by the local propeller induced
velocity pushing fluid forward against the incoming
free stream flow. Extreme flow unsteadiness directly
coupled to varying degrees of blade surface flow
separation make computation of forces extremely
difficult.
Jiang et al. (1996) documented the unsteady ring
vortex structure using PIV and related the unsteadiness
to the measured unsteady shaft forces. Jiang identified
an oscillation of the ring vortex at a frequency much
lower than the propeller rotation rate.
Application of CFD to the crashback problem was
first reported by Chen (1996). RANS computations for
Propeller 4381 showed large over predictions when
compared with open water data, which were attributed
to effects of cavitation. Viscous effects and extremely
large inductions resulted in the tip region of the blade
producing attached flow, while the root region
exhibited separated flow. Figure 2 from Chen (1996)
shows the mean flow field in the axial plane, which
exhibits a ring vortex outboard of the tip. Chen
modeled only a single blade passage, thus assuming
blade periodic flow. This assumption precluded
simulating the time varying, spatially nonuniform flow
field, and can only simulate unsteadiness within a
blade passage.

INTRODUCTION
Propeller performance prediction at near design
operating conditions has become well developed using
a variety of analysis tools. The state of the art is well
represented by the 22nd ITTC propulsion committee
Propeller RANS/Panel Method Workshop, with results
summarized in the 22nd ITTC propulsion committee
Report (ITTC, 1999). Lifting surface, Panel method
and RANS computational approaches have all
demonstrated reasonable prediction capability near
design operation for convention single, open
propellers. Data at design condition has been provided
by Jessup (1989, 1994) in way of LDV measurements
on simple open propellers and blade pressure
measurements controllable pitch type propellers in
uniform and inclined flow (1982).
The ahead, bollard condition represents some
additional complexity, where the propeller is heavily
loaded, with very high propeller inductions. The blade
sections are at high angle of attack, with large radial
flow, with large section drag coefficients and very
strong tip and hub vortices.
Steady backing conditions represent a further
complication in propeller blade flow. In this case the
ship is moving backwards, the propeller is turning in
reverse. The sharp trailing edge becomes the leading
edge and the blunt leading edge becomes the trailing
edge. Also the sign of the section camber becomes
reversed, ie, negative camber. This results in extreme
blade section pressure distributions and can lead to
leading edge separation, which can be compared to
airfoil stall conditions occurring at angles of attack
above around 25 degrees. Jiang (1991) presented Panel
method calculations for Propeller 4381 operating in
steady backing. Jiang also outlined a procedure to
estimate crashback performance using panel method
with lift corrections due to leading edge separation.

Figure 2 Chen (1996) RANS solution, open water, J=0.8


Unsteady
simulation of
reported by
simulation of

RANS was applied to the motion


a submarine in a crashback maneuver,
Davoudzadeh et al. (1997). A full
the entire submarine body and rotating

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Propeller 4381 Geometric Characteristics

propeller 4381 was performed during which the


propeller rotation rate was reversed. All the propeller
blades were modeled separately to capture the unsteady
flow field within the entire propeller disk. The flow
field was analyzed in the vicinity of the propeller
showing the unstable character of the vortex ring
structure (Figure 2). Unfortunately, the computation
was not validated with any experimental data.

Number of Blades: 5
Diameter: 12.0 inches (305 mm)
Expanded Area Ratio: 0.726
Right Hand Rotation
Thickness Section: NACA 66 (DTMB Modified)
Camber Section: a=0.8 meanline
Material: Aluminum
Skew, Rake = 0.0
r/R

C/D

T/C

P/D

F/C

0.20

0.174

0.250

1.26

0.0312

0.30

0.228

0.156

1.35

0.0369

0.40

0.275

0.107

1.36

0.0348

0.50

0.313

0.077

1.34

0.0307

0.60

0.338

0.057

1.28

0.0244

0.70

0.348

0.042

1.21

0.0189

0.80

0.334

0.031

1.14

0.0147

Figure 3 RANS simulation of crashback flow


(Davoudzadeh et al., 1997)

0.90

0.281

0.024

1.07

0.0122

0.95

0.219

0.026

1.03

0.0133

For the prediction of propeller strength, the


determination of the unsteady flow and unsteady blade
loading is required. There has been some advancement
in the application of unsteady RANS but without
validation of computed blade loads. A fully validated
analysis method is not expected in the near future.
Therefore, presently, prediction of mean and unsteady
blade loads requires reliance on empirical procedures.
This paper will describe additional measurements
conducted with propeller 4381 to further understand
both bollard and crashback conditions. Combined
performance measures using LDV, PIV, cavitation
observation, flow visualization and mean shaft load
measurements were obtained. Traditional analysis
methods were applied to investigate their usefulness.

0.98

0.153

0.037

1.01

0.0164

0.99

0.115

0.050

1.01

0.0211

1.00

0.000

0.070

1.00

0.0280

X=0.0

Figure 4 Propeller 4381

TEST SET-UP
All measurements were made in NSWCCDs 36
water tunnel. Propeller 4381 was used with geometry
listed in the following Table. The propeller was run
with an upstream shaft arrangement, shown in Figure
4.

Measurements were performed when the propeller


was operated at near-bollard ahead and crashback
conditions. A true bollard condition could not be
obtained due to water tunnel recirculation. The Table
below shows conditions run. To evaluate baseline
ahead loading performance, conditions were also run
over a range of positive J.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Laser Doppler Velocimetry (LDV) Measurements


The LDV system consisted of two TSI model 9832
fiber optic probes rigidly mounted together on a
traverse, which could translate in the axial and radial
directions as shown in Figure 5. The focal distance of
the probes (470 mm in water) was sufficient to place
the probe bodies outside of the jet. The horizontal
probe utilized the green (514.5 nm) and violet (476
nm) beams of an argon-ion laser to measure the
tangential and axial components of velocity,
respectively. The vertical probe utilized the blue (488
nm) beam of the argon-ion laser to measure the radial
component of velocity. Both probe volumes were 0.07
by 1.3 mm.

Test Conditions
J

Near-bollard
Crashback
Crashback

+0.3
-0.5
-0.7

n
(rpm)
660
-700
-600

U
(m/s)
1.006
1.778
2.134

U
(ft/s)
3.30
5.83
7.00

MEASUREMENTS
Propeller Dynamometry
Propeller shaft thrust and torque loads were
measured with the 36-inch Water Tunnel standard
strain gage Dynamometry. Tunnel velocity was
calculated via a differential pressure measurement
through the contraction with a correction factor based
upon an LDV survey in an empty test section. Data
acquisition and analysis system was based on National
Instruments hardware; a PXI-1000B rack, PXI-8170
embedded computer, and a PXI-6031E multi-function
I/O device with 16 bit A/D resolution. Software and
data reduction routines were written in LabVIEW v6.1.

Open Jet

Motion

Motion
Fiber-Optic
Probes

Laser Sheet Flow Visualization


Laser sheet flow visualization was performed
using an argon-ion laser producing 5 watts of power
on all lines. The light was coupled into a multimode
optical fiber to direct the light to the test section. The
light sheet was generated using a cylindrical lens
assembly at the fiber output end. The light sheet had a
divergence angle of approximately 7.5 in water, and a
sheet thickness of approximately 10 mm (0.4 inches).
The light was projected through the upper diagonal
window of the test section with the plane of the light
sheet containing the propeller axis. The camera viewed
the sheet through the opposite upper diagonal window,
so that the image direction was perpendicular to the
light sheet. The camera recorded images of cavitation
bubbles passing through the light sheet.

Two-axis Traverse

Figure 5 36WT LDV system


The measurement volume was positioned at a
point in the horizontal plane containing the propeller
axis. The probe volume was translated in the axial and
radial directions in order to get two directions of
movement in the flow field, while the rotation of the
propeller relative to the measurement point provided
the third direction of movement. The position of the
shaft was encoded with an 8192-counts/revolution
signal, which was recorded with each velocity
measurement. The measurements are grouped into
1024 circumferential positions, each 8 encoder counts
wide. Analysis involving time average results merely
averaged data at each traverse position.
Doppler signals were analyzed with a TSI Model
IFA 655 Digital Burst Correlator. The processor
performs
a
256-sample,
double-clipped,
autocorrelation on each Doppler burst, allowing the
measurement of velocity even when the signal-to-noise
ratio is low. The processors were operated in the
random mode.
The flow was seeded with 1500 grit siliconcarbide powder with particles 1 to 2 m in size. Since
the water in the tunnel recirculates, seed particles
needed to be added only infrequently.

Cavitation Observations
Cavitation observations were performed with
strobe-light illumination. Recording was performed
with a Sony Digital Betacam broadcast quality video
camera/recorder. Still images of the cavitation were
downloaded from the video to JPG files using an
AVID editing system. In crashback, the cavitation and
flow visualization experiments were conducted at
tunnel pressures of 83, 110, and 138 kPa (12, 16, and
20 psia).

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

t
Vx

Coordinate systems used for LDV data


All distances presented are normalized by the
propeller radius, R. The velocity measurements
presented are generally normalized by the tunnel
velocity, U. The exception to this are velocities with a
superscript asterisk, which are normalized by the
relative inflow velocity.

V = U 2 + ( 2rn ) 2

Vt

90-

(1)
x

Figure 6 Primary-secondary flow coordinate


system for LDV measurements

Measurements of the circumferentially averaged


velocities shown in the X-R plane are presented in the
stationary coordinate system, while the propeller wake
planes are presented in the rotating coordinate system
of the blades. In order to better illustrate the features
of the flow, the velocities in the downstream
measurement planes are rotated into the primarysecondary coordinate system shown in Figure 6. In this
coordinate system, a primary, or streamwise direction s
is defined as being in the axial-tangential plane, at the
propeller pitch angle, . The secondary velocities are
then the orthogonal velocity component in the x-t
plane, Vc, and the radial velocity, Vr. These velocity
components can be calculated from:
Vs =

LDV Uncertainty Analysis


The uncertainties for the fundamental quantities
measured in this experiment are listed in the table
below.
Uncertainties of fundamental quantities

sin V x + cos Vt

Vc = cos V x + sin Vt

(2)

Vr = U r

Type B

Type A

0.005

0.003

0.00008

0.00014

df

0.003

0.0005

t,r

0.0010
0.05

Those uncertainties, which are the same for all


measurements, are listed as Type B uncertainties, and
those uncertainties, which vary for each measurement,
are listed as Type A uncertainties per the ISO
Uncertainty Guide (1995). Uncertainties are listed as a
fraction of the nominal value, unless otherwise noted.
The uncertainties in x, t, and r are the uncertainties
in positioning the probe volume with respect to the
propeller in the axial, tangential, and radial directions.
Uncertainty in the measurement of the frequency is
assumed to be small relative to the uncertainty due to
finite sample size, and so is ignored. The uncertainty in
the perpendicularity of the three measured components
is assumed to be small compared to the uncertainty in
fringe spacing and probe volume coincidence, and is
ignored as well.
The calculated uncertainties for the quantities
found by combining other measurements are presented
in the table below.

Since the pitch angle is different at each radius, the


coordinate system is different at each radius as well.
For a given radial position, the pitch angle of the
propeller at that radius is used. If the radius is inboard
of the propeller root or outboard of the propeller tip,
is defined by the root or tip P/D, respectively.
The flow vector in the stationary frame, U, and the
flow vector in the rotating frame, V, are related by
Vt = Ut - 2rn ,

Item

Vr = V r

Vx = Ux ,

(3)

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Combined Uncertainties
Case 1 (Inviscid
Flow)
Item

Type
B
0.5

Type
A
0.3

Total

Type
A
0.3

Total

Ref

0.6

Type
B
0.5

Ux, Vx

0.3

0.2

0.6

0.4

0.3

3.0

3.0

Ut

0.1

0.1

0.2

0.2

1.5

1.5

Ur, Vr

0.03

0.1

0.1

0.1

3.0

2.3

Vt

0.1

0.1

0.2

0.2

1.5

1.5

Vs

0.2

0.2

0.3

0.2

2.0

1.9

Vc , Vp

0.3

0.2

0.4

0.3

3.0

2.8

0.0

10.

10.

0.0

10.

10.

Particle Imaging Velocimetry (PIV)


The PIV measurements were made using two
Spectra Physics Quanta-Ray PRO-Series Pulsed
Nd:YAG lasers rated at 800-mJ/pulse output at 532
nm. The laser output was formed into a light sheet of
5-mm thickness. The light sheet passed through
windows in the water tunnel to illuminate the flow.
The sheet was oriented horizontally, parallel to the
propeller-shaft axis shown in Figure 7. The axial
position of the camera and sheet were moved to place
the blade center span in the center of the image. Silicon
Carbide particles of 1-2 m diameter were added to the
flow.
A LaVision Flowmaster 3S PIV/PTV system was
used to control the firing of the lasers and synchronize
image capture with a digital imager, a 1280 x 1024
pixel cross-correlating camera with 12-bit resolution.
The camera axis was perpendicular to the laser sheet
and recorded the in-plane motion of the tracer particles.
The light sheet was sufficiently thick to reduce the
number of particles that entered or exited the light
sheet due to strong cross-plane flow velocities. The
camera was contained in a waterproof housing within
the test section of the water tunnel. The image field-ofview was 125 x 100 mm for the r and x coordinates,
respectively, and the lenses on the camera were chosen
to have a depth-of-field much larger than the thickness
of the light sheet. The PIV images were spatially
calibrated by taking images of a registration target in
the image plane. Images of the target were taken in the
filled test section. The cross-correlation PIV images
were analyzed using the La Vision software DaVis
version 5.4.4. An adaptive multi-pass image process
was employed starting with interrogation windows of
64 x 64 pixels and repeating with a final interrogation
window of 32 x 32 pixels. The spatial resolution of the
velocity vector field is 3 mm. Typically, the bad
vectors were fewer than 5% before post processing.
With the advantage of the high spatial resolution of the
PIV vector fields, assessment of the instantaneous axial
and radial velocity in the plane of the laser sheet was
possible. The PIV test configuration is shown in Figure
7.

Case 2 (Propeller
Wake)

The calculation of the uncertainty in J from the


uncertainties in the fundamental quantities is
straightforward. However, the rest of the items in the
table above can only be calculated with information on
the local flow conditions. This is because all of these
items depend on the values of the velocity, velocity
gradients, or turbulence intensity. The uncertainties are
therefore listed for two representative flow conditions.
The first, case 1, is a point in the inviscid flow
between the blade wakes. In this region, the turbulence
intensity is low and the flow gradients are small. Case
1 is representative of the majority of the flow. Case 2
is a point in the blade wake. At this location, the
turbulence intensity and the velocity gradients are at
their highest values, and so the uncertainties are a
maximum. Case 2 is representative of only a very small
fraction of the flow, but the fraction of most interest.
For case 1, the velocity uncertainties are all below
0.5% of the inflow velocity. In this region of the flow,
the velocity uncertainty is dominated by the
uncertainty in the fringe spacing of 0.3. For case 2, the
uncertainties are higher due to the uncertainty in
finding the mean in a high turbulence region with a
relatively small sample size (~250). Type A
uncertainty in the measurement of the turbulent
velocity fluctuations, q, is dominated by the
uncertainty in finding the variance of a distribution
with a finite sample size. For a sample size of 250, that
uncertainty is approximately 10%. The LDV has a
lower noise floor, below which it cannot measure the
turbulence. For this setup, the noise floor in q was
approximately 1.5% of the measured velocity.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

data and predictions in shown in Figure 8. Design J is


0.889. Discrepancies between the tow tank and water
tunnel results may be due to tunnel effects or
differences in hub boundary conditions. Interestingly,
the efficiency is well matched for all data and
predictions, especially at low J.
Water Tunnel KT
Water Tunnel 10KQ
Water Tunnel O
Tow Tank KT
Tow Tank 10KQ
Tow Tank O
PSF3 KT
PSF3 10KQ
PSF3 O

1
10KQ

0.9
0.8

KT, 10KQ, O

0.7

Figure 7 PIV set up in NSWCCD 36 WT

0.6
KT

0.5
0.4
0.3

PIV was used to take instantaneous snap-shots


of the flow going through the propeller. Data was
acquired at two crashback conditions (J = -0.5 and
0.7). The PIV data consisted of roughly 500 images of
the flow field at each of eight different propeller
angular positions covering a single blade passage. This
PIV data can be used to quantify the magnitude of the
large flow variations through the propeller. The highest
frame rate of the lasers was 8 Hz. When triggering the
images from a once/rev pulse, images were
sequentially obtained every other revolution.
Data was also acquired at three ahead conditions (J
= 0.3, 0.6, and 0.889). Fewer PIV images (200) were
required to define flow field variations that were much
smaller than under crashback conditions.
PIV was also used to look at the larger mean flow
field about the propeller. This larger flow field was
obtained by combining data from six different sheet
and camera locations. The PIV data at each location
consisted of 120 images taken at a single propeller
angular position. The spatial resolution remained 3
mm, but the field of view was increased to 250 x 250
mm for the r and x coordinates. Figure 22 shows one of
these vector mosaics (J = -0.5). Only 50% of the
measured vectors appear in this figure.

0.2
0.1
0

0.2

0.4

JA

0.6

0.8

Figure 8 Comparison of open water data and


predictions
Near Bollard Condition

The bollard condition represented in the water


tunnel was obtained at a J=0.3. This was the lowest J
obtainable, due to flow recirculation in the water
tunnel. Figure 9 shows the propeller in operation with
visible tip and hub vortices.
The spanwise loading can be represented from the
measured circumferential average tangential velocity
using LDV measured at X/R=0.23. Comparison is
shown in Figure 10. The circulation , G(r) is defined
as,
G(r) = UT (r)ave r / Z,

Z=blade number.

(4)

Wake contraction is not accounted for, but the


very large effect of the tip vortex can be seen. From the
hub to the 0.8 radius, the measured result is consistent
with the overprediction of thrust and torque. The
potential influence of the local tip loading on the
overall thrust and torque prediction is clear from the
circulation.
Figure 11 shows the blade wake of propeller 4381
at J=0.3. The blade flow appears attached, but shows a
significantly thicker blade wake that typically observed
for propellers operating at design J. Also seen is the
intense tip vortex. The hub flow has not sufficiently
developed at X/R=0.23 to show the occurrence of a
large hub vortex. The blade wake at 0.7R is shown in

RESULTS
Ahead Condition
Baseline ahead measurements were performed to
verify analysis tools for this relatively simple blade
geometry. In tunnel Open water measurements were
compared to earlier tow tank open water performance
(Hecker, 1968) and predictions with the potential based
panel code, PSF10 (1993), and lifting surface code,
PSF3 (1982). Comparison of water tunnel, tow tank

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

detail in Figure 12. The high gradient side of the wake


is convected from the pressure side of the blade, while
the lesser gradient side is from the suction side of the
blade where the pressure distribution is adverse due to
the high blade angles of attack. The turbulent kinetic
energy is consistent with foil wakes at high angles of
attack, and does not indicate separated flow.
Section drag coefficients were tabulated as twice
the momentum thickness in the wake and compared to
those of Jessup (1989) for Propeller 4119 operating at
design J, shown in Figure 13. Significant increase in
blade drag is seen as compared to sections at ideal
angle of attack. The increase in section drag was used
to recompute the open water performance at J=0.3 as
shown in the table below. The computed KT moves
toward both the tow tank and water tunnel results, but
KQ increased well beyond the test results. The
efficiency prediction is improved using the measured
section drag. Differences in the predicted and
measured spanwise circulation distribution in the tip
region appear to be as significant as the refinement of
the blade section drag.
Figure 14 shows the time average PIV result for
the axial velocity and vector field at J=0.3 in the
measured X-R plane. To better visualize the vortex, the
freestream velocity, U is removed. Results show the
very pronounced tip vortex from an alternative
perspective of that typically seen in the Figure 11, Y-Z
(or R-) plane. Figure 15 shows the progression of the
tip vortex with blade angular position. Each image is
18 degrees progressed from the previous, or blade
passage.

0.3

0.2

KT
0.507
0.495
0.432
0.469

10KQ
0.759
0.857
0.747
.789

0.15
0.1
0.05
0
0.2

0.4

0.6

0.8

r/R

Figure 10 Spanwise circulation, J=0.3


1
V^*_s
1.00
0.96
0.92
0.88
0.84
0.80
0.76
0.72
0.68
0.64
0.60

0.9

0.8

0.7

0.6

0.5

0.4

0.3

Thrust And Torque Predictions At J=0.3


J=0.3
PSF10, Cd=0.007
PSF10, Cd,meas
Water Tunnel
Tow tank

Measured J=0.30
PSF10 J=0.30
PSF3 J=0.30

0.25

0.2
-0.3

0.319
0.275
0.274
0.284

-0.2

-0.1

r 2004

0.1

0.2

0.3

C:\Public-old\p4381-01\Prop 4381 x23 s660 j03 Vs port

Figure 11 Propeller 4381 blade wake from LDV,


streamwise velocity, J=0.3, X/R=0.23
1.0
0.28

0.24
0.9

0.16

V *s
q*

0.8

q*

Vs

0.20

0.12

0.08

0.7

0.04

0.8

0.9

1.0

1.1

Figure 12 LDV measured blade wake, 0.7R, J=0.3

Figure 9 Propeller 4281 operated at J=0.3

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

0.040

1.1

0.030
0.025
0.020

R / R PROP

Drag Coefficient, Cd

0.035

Cd, P4381

0.015

Cd, P4119

0.010
0.9

0.005
0.000
0.3

0.5

0.7

0.8
0.8

r/R

-0.2

-0.1

0.1

Figure 13 Computed drag coefficients, P3381 at J=0.3,


P4119, JDES=0.83
p

0.2

0.3

X / R PROP

J = +0.3, 660 rpm

1.1

R / R PROP

Reference Vector
Ur = +0.65
Ux = -1.64

1.1

1
U

0.9

0.9
Ux
2.50
2.00
1.50
1.00
0.50
0.00
-0.50
-1.00
-1.50
-2.00
-2.50

0.8

0.7

0.6

0.8
-0.2

-0.1

0.1

0.2

0.3

X / R PROP

0.5
1.1

-0.4

-0.3

-0.2

-0.1

0.1

0.2

0.3

0.4

0.5

R / R PROP

0.4
0.6

Figure 14 Propeller 4381 blade wake from PIV,


J=0.3, axial velocity, vectors have U removed

0.9

0.8
-0.2

-0.1

R / R PROP

0.1

0.2

0.3

X / R PROP

1.1

Figure 15 Tip flow from PIV, J=0.3, time average images


18 apart, U removed
1

Crashback Conditions, Time average flow

0.9

The crashback results are initially shown in the


time average sense, with unsteadiness averaged out.
Figure 16 shows the time average thrust and torque.
The minimum seen in the thrust and torque at J=-0.5 is
typical of other propellers. The discrepancy between
the towing tank open water data, and the water tunnel
is hypothesized to be due to tunnel effects or
differences in hub boundary conditions. Unfortunately
this introduces uncertainty when comparing
measurements to predictions, where tunnel effects are

0.8
-0.2

-0.1

0.1

0.2

0.3

X / R PROP

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

not included. Figure 17 shows the unsteady thrust and


torque as a standard deviation from the mean of the
time history data. The peak unsteadiness occurs also at
J=-0.5.

0.0

KT & 10KQ

-0.2

0.0
600 rpm

-0.2

700 rpm

800 rpm

KT & 10KQ

-0.4

J = -0.5

-0.6

KT1
KT2

-0.8

KQ1

-0.6

KQ2

KT

-1.0

-0.8

-1.0
-1.2

KQ

-1.0

-0.8

-0.6

-0.4

-0.2

0.0

0.30
Crashback T1
CB T1, SR 2 kHz
Crashback T2
CB T2, SR 2 kHz
Ahead T1
Ahead T1, repeat
Ahead T2
Ahead T2, repeat

0.15
0.10
0.05
0.00
0.2

0.4

0.6

0.8

10

The LDV data obtained in the X-R plane was time


averaged to show the ring vortex and flow though the
propeller disk for J=-0.7 and J=-0.5. Figures 19 and 20
show the vector field from the measured axial and
radial velocity, and the contours show the axial
velocity. At J=-0.5 the lesser free stream velocity
positions the center of the vortex ring closer to the
propeller disk, The higher free stream velocity, at J=0.7 moves the ring aft. Also noted is the amount of
radial flow exiting the forward edge of the propeller
disk. At J=-0.7 significant radial flow exists over the
blades and forward of the propeller, which moves the
flow outward into the recirculation region. At J=-0.5,
little radial flow exists. The flow is accelerated forward
out of the LDV measurement domain. This could be
contributing to higher flow instability.
The position of the vortex ring can be compared to
the RANS computation by Chen(1996), which was
shown in Figure 2. At J=-0.8, Chen shows the vortex
ring positioned almost directly over the tip of the
propeller. From the measured result, as J is decreased,
with increasing external velocity, the vortex is
expected to move aft. Chens computation may also
have been influenced by different hub boundary
conditions. The water tunnel case incorporated an
upstream shaft, while Chen incorporated the tow tank,
down stream shaft configuration. The absence of the
upstream shaft could expect to move the ring vortex
forward.

Figure 16 Thrust and torque coefficients for


crashback

0.20

Re/10
Figure 18 Effect of loading on nondimensional thrust
and torque, J=-0.5

0.25

Symbols:
Water Tunnel Data
Solid Lines: Tow Tank Data

-1.4

T/T

-0.4

1.0

Abs(J)

Figure 17 Unsteady thrust at ahead and crashback


conditions
Blade bending under load has been observed under
crashback conditions. Highly skewed propellers will
show a tendency to increase pitch under backing and
crashback
conditions,
which
will
cause
nondimensional thrust and torque to increase with
increasing propeller Reynolds number. This was
observed under backing conditions by Jiang (1991) for
Propeller 4383, which was a 72 degrees of skew
version of Propeller 4381. Figure 18 shows very little
variation in KT, and KQ with Reynolds number. This is
attributed to an even chordwise bending under load of
the unskewed propeller, which results in little blade
pitch change.

10

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

the unsteady nature of the blade flow at the crashback


condition.

2.5

U_x

1.5

U INF

1.1

1.08
0.83
0.59
0.35
0.10
-0.14
-0.38
-0.63
-0.87
-1.11
-1.36
-1.60

U X / U INF

1.2
0.85

0.9

0.5
0.8

0.15

0.5

-0.2

0.7
0

-0.55

-0.9

0.6

Figure 19 Time average axial velocity and streamlines,


from LDV, J=-0.7

-1.25
0.5

2.5

-1.6

0.4
-0.4

-0.2

-0.1

0.1

0.2

0.3

0.4

Figure 21 Time average axial velocity and streamlines,


from PIV, J=-0.5

1.20
0.80
0.40
0.00
-0.40
-0.80
-1.20
-1.60

1.5

-0.3

Ux

1.5
0.5

Figure 20 Time average axial velocity and streamlines,


from LDV, J=-0.5

The time average flow is also shown from the PIV


data in Figure 21. The multiple individual images taken
at one of the blade angular positions were averaged
and appears qualitatively consistent with the LDV data
in Figure 20. The PIV mosaic in Figure 22 shows the
stagnation flow forward of the propeller at J=-0.7.
Figure 23 shows the phase averaged LDV data
upstream of the propeller at X/R=-0.24. This plot is
analogous to the downstream wake shown in Figure 11
showing the blade wake. For the crashback case the
flow is accelerated forward. A diffuse blade wake can
be seen in the plot, showing a sharper wake definition
on the back side of the blade, and a thicker, more
diffuse wake on the face of the blade. At crashback, the
blade face becomes the low pressure side of the blade,
which will produce a thicker blade boundary layer and
wake. The diffuse nature of the blade wake is due to

PROPELLER
# 4381

0.5

-1

-0.5

Vector Length Key [V/U] :

0.0 0.1 0.2 0.3 0.3 0.4 0.5 0.6 0.7 0.7 0.8 0.9 1.0 1.1 1.1 1.2 1.3

Figure 22 Time average streamline velocity and


streamlines, from mosaic of six PIV images, J=-0.7

11

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

-0.60
-0.65
-0.70
-0.75
-0.80
-0.85
-0.90
-0.95
-1.00

0.8

0.7

Convergence Of Panel Computations In Crashback

V*s

0.9

0.6

0.4

0.3

-0.2

-0.1

0.1

0.2

KT

KQ

1
2
3
4
Tunnel
Data

0.17770
0.20910
0.21984
0.22073
0.25

0.03785
0.04277
0.04460
0.04450
0.050

KT
Error
-29%
-16%
-12%
-12%

KQ Error
-24%
-14%
-11%
-11%

The favourable prediction appears fortuitous upon


investigating the computed pressure distributions with
a consideration of cavitation. Figure 27 shows the
predicted blade pressure distributions at 0.7R using
PSF10 for three cases: crashback at J=-0.5, steady
backing at J=0.5, and ahead at J=0.5. There is a
similarity for the various cases. The effect of reversed
camber and nose tail line can be seen in the mid chord
region. All cases have a very large suction peak at the
leading edge. For the backing cases, this is more
pronounced due to the sharp trailing edge becoming
the leading edge to the inflow. The table below shows
the reduction in load if the suction peaks are clipped at
the CP Min values shown. Although clipping is not
necessarily realistic to the actual viscous flow
behaviour, it shows a potentially large effect of
cavitation. The load measurements were obtained at
=150. LDV and PIV were conducted at =210. These
conditions were run to eliminate cavitation, but no
careful inception measurements were performed.
Unfortunately, the occurrence of cavitation was very
intermittent due to unsteady effects.

0.5

0.2
-0.3

Iteration

0.3

z
Figure 23 Crashback blade wake flow, X/R=-0.24,
(upstream), J=-0.5, streamwise velocity

Prediction of mean loads for Crashback Condition


using Panel Method
From the time average LDV velocity data
measured downstream of the propeller, panel method
calculations were performed to predict performance.
Jiang (1991) provided predictions of steady backing
and crashback for propellers 4381 using PSF3
(Greeley, 1982) and showed reasonable agreement
with tow tank results. For the present study, the panel
code, PSF10, was used along with the measured inflow
entering the aft end of the propeller, shown in Figure
24.
Figure 25 shows the velocity components
extracted from the LDV data in Figure 24. An
interesting observation is the occurrence of not only
significant forward axial velocity, but also tangential,
swirled inflow. Large swirl exits the forward side of
the disk, due to its generation of thrust. This induces
large rotation in the flow, which may be a driving
component of the ring vortex behaviour. The flow that
recirculates back into the propeller disk retains swirl.
A field point velocity computation with PSF10
was used to match the total velocities shown in Figure
25. Convergence was obtained in four iterations as
shown in the table below. The total and effective
velocity distributions are shown in Figure 26. The
computed load compared reasonably well with the
water tunnel measurements at J=-0.5.

Effect Of Suction Peak Clipping On KT And Kq


CP Min
-9000
-200
-100
-25

KT
0.221
0.156
0.148
0.131

KQ
0.0445
0.0346
0.0351
0.0341

An additional computational uncertainty is the


sharpness of the leading edge. The propeller
incorporates a NACA 66 DTMB modified thickness
distribution. PSF10 assumes a sharp trailing edge, but
the actual propeller model has some rounding of the
trailing edge, which is difficult to define.

12

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

50
Crashback (Iter4) J=-0.5
Steady Backing J=0.5
Ahead J=0.5

40

30

-CP

20

10

-10

-20

Figure 24 Extracted propeller inflow for


prediction using panel method, PSF10, J=-0.5

r/R
0.60

0.40

-1.50

-1.00

-0.50

0.00

U/U inf

0.50

1.00

Figure 25 Total velocity components for prediction


using panel method, PSF10, J=-0.5
1.00
0.90
0.80

r/R

0.70
Vx Total Velocity
Vr Total Velocity
Vt Total Velocity
Vx Effective Inflow after 4 iterations
Vr Effective Inflow after 4 iterations
Vt Effective Inflow after 4 iterations

0.60
0.50

x/c

0.6

0.8

Tow Tank Backing And Crashback Open Water Data

0.40

0.30
0.20
-0.40 -0.20

0.4

Crashback conditions have often been viewed


from a simple outer flow argument resulting in very
large section angles of attack. The measured propeller
induction in crashback, J=-0.5, reduces the section
angle of attack from a geometric angle of 43 degrees to
an effective angle of 13 degrees, which is very similar
to the backing condition at J=0.5 of 16 degrees. The
diagram in Figure 28 shows the velocity triangles for
J=0.5 in the ahead condition, backing condition, and
the crashback condition.
From the tow tank open water data, the backing
and crashback load data are compared in the table
below. At J=0.5 there is similarity in loading at
crashback and backing conditions. At other values of J
shown, there appears no similarity. The axial and
tangential inflow velocity at J=-0.7 is dimensionally
similar to the J=-0.5 case, which would lead to similar
vector diagram as seen below for J=-0.5. With
substantially larger thrust and torque at J=-0.7, it is
hypothesized that outboard radial flow may be
augmenting the blade thrust as J becomes more
negative. The present panel code wake models cannot
likely model the very large outward radial flows as J is
reduced.

0.80

0.20
-2.00

0.2

Figure 27 Computed blade pressure distributions at


0.7R using PSF10

U X/U inf
U T/U inf
U R /U inf

1.00

0.00

0.20

0.40

0.60

0.80

Vx, Vr, Vt

1.00

1.20

1.40

0.3
0.5
0.7
0.9

1.60

Figure 26 Effective velocity components for


prediction using panel method, PSF10, J=-0.5,
(Vt, total = Vt, effective)

KT
backing
-0.353
-0.272
-0.174
-0.067

10KQ
backing
-0.731
-0.579
-0.400
-0.203

13

Copyright National Academy of Sciences. All rights reserved.

KT
crashback
-0.443
-0.331
-0.503
-0.732

10KQ
crashback
-0.915
-0.671
-0.936
-1.285

Twenty-Fifth Symposium on Naval Hydrodynamics

Forward

then broke up into two vortices, then reformed at the


tip, with a period of 2.5 seconds. This observation
correlated with unsteady shaft force periodicity of 2.1
seconds. Discrepancies at J=-0.5 in the present tests,
and the earlier 24WT tests are not fully understood,
but may be due to differences in test section size. Jiang
observed steadier flow at J=-0.732, similar to the
present tests. A simple rotational frequency can be
associated with predominate swirl of the ring vortex,
which is shown in Figure 32. At J=-0.5 a swirl of 0.150.2U corresponds to a rotational frequency of 2.7-3.6
Hz at the propeller tip radius, with a period of 0.3-0.4
seconds, much faster that the period identified by
Jiang. At J=-0.7, the predominate swirl is around
0.08U, significantly less that at J=-0.5.

Crashback, geo= 42
r

= 13

-V
Backing, geo= 16

Crashback, effective= 13

(Crashback, and backing condition, J=-0.5, 0.5)


Forward
geo= 16
= 29
= 13

Rotation period = (UT J n)/ r

(5)

J = -0.7, 600 rpm

2.5
U

(Ahead condition, J=0.5)

q
1.60
1.50
1.40
1.30
1.20
1.10
1.00
0.90
0.80
0.70
0.60
0.50
0.40

Figure 28 Velocity triangles for Prop 4381 at J=0.5 for


ahead, backing and crashback conditions at 0.7R

1.5

Crashback Conditions, Instantaneous Flow

Large unsteadiness in the crashback flow was


expected, based on Jiangs (1996) measurements, and
computations by Davoudzadeh, F., et. al., (1997).
Figure 17 showed large fluctuations in shaft thrust
and torque that were most pronounced at J=-0.5. The
LDV results further identified unsteadiness from the
velocity fluctuations represented by the turbulent
kinetic energy, q. Figure 29 and 30 show levels of
turbulence in the X-R plane. Levels are significantly
larger than any other propeller flow condition, with
levels as high as 1.6 in the region of the vortex ring.
Levels are significantly higher at J=-0.5 compared to
J=-0.7.
Flow unsteadiness was also observed in video
recording of the laser sheet images. A typical image is
shown in Figure 31. The videos did not show any
obvious cyclic behavior of the vortex ring. At J=-0.5,
there was some indication that the ring moved inboard
and then outboard, but at no obvious frequency. Jiang
(1996), using the same propeller, reported a periodic
behavior operating at J=-0.472 in the NSWCCD 24
Water tunnel (Propeller rpms were not stated, but
presumed to be in the 600-700 range). Jiang described
a periodic movement of the ring from the forward
location seen in Figure 30 to an aft location, which

0.5

Figure 29 Turbulent kinetic energy from LDV, J=-0.7


p
J = -0.5, 700 rpm

2.5

q
1.60
1.50
1.40
1.30
1.20
1.10
1.00
0.90
0.80
0.70
0.60
0.50
0.40

1.5

0.5

0.5

1.5

2.5

Figure 30 Turbulent kinetic energy from LDV, J=-0.5

14

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

A major concentration of this effort was the local


blade flows as related to blade strength. The larger
global flow incorporating the vortex ring and its
behavior was believed to be driven by the local blade
flow. Therefore, the PIV data was obtained within a
much smaller spatial domain, encompassing the blade
and only nearby upstream and downstream flow. The
typical measurement domain was shown in Figure 21
for the time average flow and Figure 33 for the time
varying flow.
An hypothesis was developed from observation of
the individual PIV images. The ring vortex appeared to
descend from its time average location around 1.7
fraction of tip radius to locations inside of the tip
radius. When this occurred, the axial velocity in the tip
region reversed direction from its time average sense.
This would produce a large increase in the angle of
attack of the blade sections near the tip, and cause a
large increase in tip loading. The tests of Jiang (1996)
described the convection of the ring vortex
downstream, and its periodic breakup. This global
behavior could have resulted in transient movement of
the vortex inboard of the blade tip resulting in the
extreme reversed tip flow, that is believed to cause
excessive blade tip loading.
The multiple PIV records were ordered by the
occurrence of the reversed flow at the tip. From 1000
PIV records, the images were ordered from most
extreme where reversed tip flow occurred to best
where the tip flow was similar to the time average flow
(Figures 21 and 22). Three samples of the most
extreme, 8th extreme, and best propeller flow is
shown in Figure 33. Figure 34 shows the most extreme
axial velocity profiles and a computed average. Figure
35 compared the extreme profile to the time average.
Observations of blade cavitation provided a
powerful visualization of the blade flow. Figure 34
shows a series of cavitation images taken in a
continuous time sequence using strobe illumination
triggered at shaft rate. Each repetitive revolution
showed dramatically different cavitation patterns on
the blades. Leading edge sheet cavitation occurred
intermittently in a random fashion. The cavities were
not typical of leading edge cavitation seen in the ahead
condition. The cavities were extremely irregular, cloud
like, and very thick. The cavities are hypothesized to
be associated with extreme transient stall behavior
resulting from the instantaneous, very large section
angle of attack occurring due to the inboard migration
of the ring vortex.

Figure 31 Laser light sheet illumination of ring


vortex,, J=-0.5
p
2.5

Ut
0.00
-0.03
-0.06
-0.09
-0.12
-0.15
-0.18
-0.21
-0.24
-0.27
-0.30

1.5

0.5

Figure 32 Mean swirl of ring vortex, J=-0.5

15

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

U INF

1.1

U X / U INF

2.4
1.8

0.9

1.2
0.8

0.6
0

0.7

-0.6
-1.2

0.6

-1.8
-2.4

0.5
0.4
-0.4

-0.3

-0.2

-0.1

0.1

0.2

0.3

0.4

Most extreme of 1000


J = -0.5, 700 RPM
U INF

1.1

U X / U INF

2.4
1.8

0.9

1.2
0.8

0.6
0

0.7

-0.6
-1.2

0.6

-1.8
-2.4

0.5
0.4
-0.4

-0.3

-0.2

-0.1

0.1

0.2

0.3

0.4

th

8 extreme of 1000
J = -0.5, 700 RPM
U INF

1.1
U X / U INF

2.4
1.8

0.9

1.2
0.6

0.8

0
0.7

-0.6
-1.2

0.6

-1.8
-2.4

0.5
0.4
-0.4

-0.3

-0.2

-0.1

0.1

0.2

0.3

0.4

Best of 1000
Figure 33 Instantaneous Axial velocity and
streamlines, from PIV J=-0.5

16

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Vi = 1.77 m/s
Tip Radius = 152.4 mm

R (mm)

150

100

Image 418
Image #021
Image #304
Image #003
Image #461
Image #035
Image #350
Average

50
Hub Radius

-5

Vx (m/s)

Figure 35 Eight Most Extreme Propeller Inflow Axial


Velocity Distributions From PIV Data
1.20
1.10
1.00

r/R

0.90
0.80
0.70
0.60
0.50
PIV Average Inflow
PIV Extreme Load Condition

0.40
0.30
-1.5

-1

-0.5

V x / V inf

0.5

1.5

Figure 36 Average and Extreme Propeller 4381 inflow


from PIV data for crashback condition, J=-0.5

Prediction of Unsteady loads for Crashback Condition


To assess blade strength under crashback
conditions, an approximate approach was developed
using the panel code solution for the mean flow and the
extreme inflow derived from the PIV measurements.
The extreme inflow condition produces velocities,
which almost match the outer free stream velocity.
This would result in geometric angles of attack close to
the 42 degrees shown in Figure 28. It also produces
negative axial inflow, which the code cannot accept.
Therefore a blade element approach was used
considering the instantaneous geometric angles of
attack, and 2D/3D corrections. The procedure only
considered extremes in axial velocity, since tangential
velocity was not measured with PIV. Figure 37 shows
the computed angles, from the axial velocities from

Figure 34 Cavitation of Propeller 4381 in crashback,


J=-0.5, 700 rpm, P=16 psia, =69

17

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25

Figure 36 at J=-0.5. The difference in from mean to


extreme, , was added to the computed values from
PSF10 panel calculations for the mean crashback case,
shown in Figure 38 (top). A lifting surface correction,
d(CL)/d(), was determined from steady backing
calculation at J=0.5 and 0.6. The lifting surface
correction, or the change in CL with was 0.21 and
was almost constant over the span. This was applied to
the mean lift to arrive at the extreme section lift, shown
in Figure 38 (bottom).
The pressure distribution developed from the panel
method analysis for the mean-crashback condition
was modified by the addition of a classic twodimensional flat plate pressure distribution that
corresponded to the predicted change in threedimensional lift coefficient.
When the resulting
pressure distribution was integrated, the blade thrust
increased by 215% and torque increased by 188%.
The bending moment at the root increased by 280%
between the average and the extreme crashback
loading due to the extra loading at the tip. The large
assumed angle of attack loading could also contribute
to large spindle torque, and possible increased pitch
deformation.

Angle of Attack ( )

20

15

10

0
0.20

0.40

0.60

0.80

1.00

r/R

1.00
0.90
0.80

AVG
EXTREME

0.70

CL

0.60
0.50
0.40
0.30
0.20
0.10
0.00
0.20

0.40

0.60

0.80

1.00

r/R

Figure 38 Estimate of extreme blade section loading


for J=-0.5, crashback
CONCLUSIONS
Propeller performance at extreme off design
conditions was investigated using experimental
techniques with some computational predictions
provided. The primary motivation was to predict
extreme blade loading for structural assessment. The
fundamental physics of the crashback condition was
also investigated using flow diagnostic tools. The
following conclusions have been reached.

1.20
1.10
1.00
0.90

r/R

AVG
EXTREME

0.80

1.

0.70
0.60
0.50
PIV Average Inflow

0.40
0.30

PIV Extreme Load Condition

-20

-15

-10

-5

Figure 37 Computed angles at J=-0.5 crashback


conditions

2.

3.

The time average crashback condition at J=0.5 is typified by flow accelerated forward
against the on coming flow which recirculates
and forms a ring vortex outside the propeller
disk, approximately centered at the propeller
plane. The induced velocity from the thrusting
propeller is large enough to keep the blade
angles of attack similar to a steady backing
condition at the same J. Outgoing swirl from
the propeller creates swirl in the ring vortex
structure, which further complicates the flow.
At J=-0.7 the ring vortex moves aft from the
propeller disk, and the flow through the
propeller develops significant radial outward
flow. This is attributed to the greater external
velocity, and leads to greater propeller thrust.
The Crashback condition creates very
significant unsteady flow. The unsteadiness is
higher at J=-0.5 that at J=-0.7, attributed to the
closer proximity of the ring vortex to the blade
tips. PIV images near the blades, in extreme
cases, show the ring vortex migrating inboard
of the propeller tip, resulting in very high

18

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

4.

5.

blade angles of attack. Cavitation observations


show erratic leading edge cavitation that
resemble transient stall. Transient peak blade
loads are estimated to be at least a factor of
two increase over that time average loads.
Depending on the operational capability and
requirements when backing, crashback can be
the critical propeller blade loading condition.
This data obtained will be useful to validate
more rigorous computational approaches to
this problem. RANS, LES, and DES
modelling may be appropriate to capture the
transient stall on the blades, and the global
behaviour of the ring vortex. Sufficient
documentation of the boundary conditions and
detailed flow data is available for validation
purposes (Chesnakas, 2004).
The near bollard ahead condition is dominated
by a very strong tip vortex. Potential based
predictions at this condition are reasonable,
but more accuracy will require significant
improvements in the modelling the tip vortex.

ACKNOWLEDGEMENTS
This work was funded by the Office Of Naval
Research, monitored by Dr. Ki-Han Kim. Supporting
panel calculations were provided by Thad Michael.
The co-authors should be commended individually
for all the contributing efforts encompassing this work.
Chris Chesnakas conducted the LDV measurements.
Dave Fry conducted the PIV measurements. Martin
Donnelly conducted load measurements. Joel Park
prepared the NSWCCD report on the work, and Scott
Black performed the Crashback analysis using the
panel code.
REFERENCES
Chesnakas, C. J., Donnelly, M. J., Fry, D. J., and Park,
J. T., Performance of Propeller 4381 in Crashback,
NSWCCD-50-TR-2004/010, August, 2004.
Davoudzadeh, F., et. al., Coupled Navier-Stokes and
Equations of Motion Simulation of Submarine
Maneuvers, Including Crashback, ASME Fluids
Engineering Division Summer Meeting, Vancouver,
British Columbia, Canada, 1997.

FUTURE WORK

Chen, B., , Computational Fluid Dynamics of FourQuadrant Marine Propeller Flow, Ms.Sc. Thesis, The
University of Iowa. 1996.

Maneuvering behaviour during crashback has been


difficult to predict and often results in indeterminate
vehicle motions. This area is related to lateral propeller
forces during crashback, which is believed to be driven
by the transient behaviour of the ring vortex. Lateral
forces result from spatial nonuniformities in propeller
inflow, which for this case is the flow entering the
propeller from behind. Future plans include reorienting
the PIV laser sheet perpendicular to the shaft axis,
located aft of the propeller and utilize high speed Stereo
PIV. This will allow instantaneous measurement of the
inflow to the propeller. Also the frequency of the PIV
image frame rate will be increased to 32 frames /sec.
This will permit capturing the true time variation of the
flow. Simultaneous measurements will also be made of
the unsteady propeller bearing forces.
To continue understanding the local blade flows,
additional measurements are planned of the
instantaneous blade section flow. A PIV sheet will be
oriented to intersect the blade chord, and measure the
leading edge flow on the blade face. The leading edge
separation, or stall like conditions will be measured and
used to validate advanced computation methods. Blade
strain will also be measured simultaneously to correlate
blade flow and blade load.

Greeley, D.S. and Kerwin, J.E., "Numerical Methods


for Propeller Design and Analysis in Steady Flow,
"Transactions SNAME, Vol. 90, 1982.
Hecker, R. and K. Remmers, Four Quadrant Open
Water Performance of Propellers 3710, 4024, 4086,
4381, 4383, 4384, and 4426, David W. Taylor Naval
Research and Development Center Ship Performance
Department Test and Evaluation Report, 417-H-01,
1971.
Guide to the Expression of Uncertainty in
Measurement,
International
Organization
for
Standardization (ISO), Geneva, Switzerland, 1995.
Jessup, S., "Measurement of Pressure Distribution on
Two Model Propellers," DTNSRDC Report 82-035,
July 1982.
Jessup, S., "Propeller BLade Flow Measurements
Using LDV," ASME Summer Fluids Division
Meeting, Lake Tahoe, NV, June 1994.

19

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Jessup, S.D.," An Experimental Investigation of


Viscous Aspects of Propeller Blade Flow," Ph.D.
Thesis, School of Engineering and Architecture, The
Catholic University of America, 1989.
Jiang, C.-W., Huang, T.T., Ng, R., Shin, Y.S., ,
Propeller Hydrodynamic Loads and Blade Stresses
and Deflections During Backing and Crashback
Operations, SNAME Propellers/Shafting 91,
Virginia Beach, 1991.
Jiang, C.W., Dong, R.R., Liu, H.L. and Chang, M.S.,
"24-inch Water Tunnel flow Field Measurements
During Propeller Crashback," 21st ONR Symposium
on Naval Hydrodynamics, 1996, Trondheim, Norway,
pp. 136-146.
22nd Iternational Towing Tank Conference (ITTC),
Propulsion committee Report, Seoul, Korea, and
Shanghai, China, 1999.

20

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

thrust, and a worst case spanwise loading


distribution, which moves the loading towards
the tip. The bollard condition would have a more
conventional spanwise loading distribution, as
shown in Figure 10 and would produce extreme
flucuations in loads, as occurs in crashback.
Another factor is the blade geometry. Large aft
skew will worsen the blade stress distribution in
crashback.

DISCUSSION
Neil Bose
Memorial University of Newfoundland, Canada
This is an interesting paper because it
describes experimental measurements that are rarely
seen in the literature. I have a number of questions:

What do the authors think was the influence of


the tunnel walls during the test? Would they
expect the flow measurements to be the same or
similar if the tests were done in a towing tank?
Had the authors considered doing the tests at
zero advance conditions in a bigger tunnel or in a
tank?
In the third point of the conclusions, the authors
say that crashback would be expected to be the
critical blade loading condition and that peak
loads are at least twice those of time averaged
loads. I assume that they mean the critical design
condition from the point of view of blade
strength? Could the authors clarify this please.
Could they also perhaps give the ratio of these
peak loads to the maximum blade load expected
say under steady bollard pull conditions?
Could a brief description be given of the PIV
system that will have a frame rate of 32 /sec?
What sort of cameras will be used?

AUTHORS REPLY

There is likely an effect of the water tunnels


boundaries on the flow field. The 12inch
diameter propeller produces a flow disturbance
out to twice its diameter running in a 36WT
test section. An additional difference in the
configuration is the downstream drive used in
open water testing, and the upstream drive used
in the water tunnel, which should more closely
resemble the ship configuration. Measured load
differences can be seen in the Figure 8 and 16.
Testing in the towing tank offers more
complexity when performing PIV and LDV due
to longer support distances for the tranversing
and support hardware.

From the data in the paper, at bollard condition,


KT=0.6. At the crashback condition where the
unsteady loads are largest, KT=-0.3-0.4. The
predictions of peak loads from the analysis in the
later section of the paper are a 200%
augmentation over the mean values. This would
produce a peak KT similar to the bollard
condition. The crashback condition may be a
worst loading case due to the reversal of the

Dave Fry describes the proposed high frame rate


PIV system as follows. The planned PIV setup
requires a pair of YAG pulsed lasers and a pair
of high speed CMOS digital cameras. The lasers
are sold by Big Sky Laser Technologies Inc.
They sell a pair of CFR400 lasers on a single
platform along with beam combining optics.
These lasers can each provide about 200 mJ /
pulse of 532 nm light at repetition rates up to 30
Hz. The cameras that will be used are the
FASTCAM-ultima APX model sold by Photron
Limited. Two monochrome camera heads (10
bit; 1024x1024 pixel) will be used to view the
measurement plane in stereo. Each head is
mated to a standard processor with it's own 2.6
gigabyte memory (up to 8 gigabyte memories are
available). Camera systems (camera head plus
processor) are capable of taking and storing 2000
full resolution images at a maximum of 2000
frames per second. The images are subsequently
downloaded to a computer through a IEEE1394
interface. The camera system has a "random
reset" mode of operation that allows PIV
investigators to operate at very small times
between successive PIV image pairs (< 1 s).
The system can therefore produce image pairs
that are processed with cross-correlation analysis
techniques. We hope to investigate even smaller
time scales by using faster pulsed lasers. These
lasers operate in excess of 1000 Hz but have an
order of magnitude lower pulse energies.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Fabio Di Felice
Istituto Nazionale per Studi ed Esperienze di
Architettura Navale, Italy
The authors should be congratulated for
their extensive study of the complicated flow
fields around a propeller operating in off design
conditions. Several experimental tools have
been used to investigate the phenomena. LDV
and PIV are used in complementary way and the
advantage of PIV in obtaining instantaneous
snapshot of the flow field is used to evaluate
inflow conditions causing extreme blade loading
in crashback. In such case in front of the
propeller a large and complex stagnation area
exists (see figure 22). The behaviour of the
stagnation area is strictly related with the vortex
ring breakdown. To point out if there is a
periodicity in the vortex ring breakdown should
be interesting to see the power spectrum of the
velocity measured by LDV in some points
(upstream of the propeller and nearby the vortex
ring for example).
An important point, as stated by the
authors, is the effect of facility blockage that
seems to be important as demonstrated also by
figure 16 where the differences of KT and KQ
between towing tank and tunnel data are
increasing when reducing J. This is probably
due to the increasing blockage of the propeller.
May be the different position of the vortex ring
from J= -0.5 and J= -0.7 is also influenced by the
facility blockage.
As suggested by the authors future work
using time resolved PIV (nowadays high
resolution camera up to 1 Mpixels at 2000
frame/sec and suitable illumination systems are
available) will give a better insight on the flow
unsteadiness.

Facility blockage can possibly affect the


flow structure in crashback. A one foot diameter
propeller in a three foot diameter tunnel may still
be affected by tunnel effects. Testing a smaller
propeller or conducting tests in a larger facility
are possibilities, but can affect the overall level
of effort of research projects of this type.
The
discussers
comments
on
conducting PIV measurements at higher
frequency are timely. Measurements conducted
in August 2004 utilized high speed cameras and
lasers to obtain images at up to 1000 frames per
second. This provides the temporal resolution to
track individual vortices in the flow, but further
complicates the data analysis process. Hopefully
this work will be reported at a later time.

AUTHORS REPLY
The authors wish to thank the discussers
for their helpful comments. The idea of
analyzing the power spectrum of the velocity is
interesting. It is a challenge to analyze the
unsteady velocity field in a fashion that can
utilize the data obtained in the most efficient
manner.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Eric Paterson
Pennsylvania State University/Applied Research
Laboratory, USA
The authors present magnitude of unsteady
forces during crash back. Can they comment on
frequency content of the unsteady forces?
One of the objectives is to understand ship
motions during crash back, however, only CT & KQ
have been measured. Do the authors plan to measure
the remaining 2 components of force and 2
components of torque? These side forces are
obviously important for understanding this complex
problem.
AUTHORS REPLY
The measurements presented in this paper
did not include side force measurements. Recent tests
conducted in August of 2004 included additional PIV
measurements, and unsteady shaft forces and
moments, including side forces. These results will be
reported at a later date.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics


St. Johns, Newfoundland and Labrador, CANADA, 8-13 August 2004

Numerical and Experimental Investigation


of the Hub Vortex Flow of a Marine Propeller
Moustafa Abdel-Maksoud 1, Katrin Hellwig 2, Jrg Blaurock 3
(1Duisburg-Essen University, 2Potsdam Model Basin, 3Consultant, Hamburg)

ABSTRACT

INTRODUCTION

The shape of hub caps can have a strong influence on


the hub vortex and the cavitation inception of this
vortex. The delay or avoidance of hub vortex
cavitation is important not only for navy ships but
also for merchant ships. This can reduce the noise
level of navy ships and avoid vibration or damage in
the rudder region behind the propeller of merchant
ships.

The form of the hub cap of a propeller is an integral


component of the propeller geometry. During the
design process of a propeller, the shape of the hub
cap is rarely given considerable attention - although
the shape of the hub cap may have an appreciable
influence on propeller efficiency as well as on the
cavitation behaviour of the hub vortex.

Numerical and experimental investigations were


carried out to study the effect of the hub cap shape on
the hub vortex. The investigations were conducted for
a propeller which was arranged with three differently
shaped hub caps. The experimental tests included
cavitation observation model tests in a cavitation
tunnel and open water tests in a cavitation tunnel and
in a towing tank. LDV and PIV measurements were
carried out to obtain more information on the
mechanism of formation, the structure of the hub
vortex and the velocity field behind the hub.
Simultaneously, the velocity fields on the propeller
and behind the hub were calculated for model and
full-scale. The comparison between the calculated
and measured results for model scale confirms that
the applied numerical method is able to predict the
complicated velocity field behind a propeller.
The calculated pressure coefficients of the full-scale
results show a strong pressure reduction in the hub
vortex region in comparison to model scale. The
calculated minimum pressure-coefficient values
behind the propeller confirm that the cavitation
inception of the full-scale takes place much earlier
than in the model scale.

To achieve the high requirements of acceptable noise


level of a navy ship propeller, it is important to delay
all kinds of propeller cavitation including hub vortex
cavitation. In some cases hub vortex cavitation may
be the first to appear. The shape of the hub cap can
play an important role in influencing the
characteristics and intensity of the hub vortex.
Therefore, when designing the shape of the hub cap,
it is important to find an optimum compromise
between cavitation behaviour, efficiency and other
important aspects of the propeller design. For these
reasons detailed information on the local flow around
and behind the hub cap is needed to support the
design process.
At the moment, the cavitation bucket diagram of a
propeller is still determined experimentally. The
extrapolation methods of model results of hub vortex
inception to full-scale differ in various model basins.
The prognosis of hub vortex cavitation inception for
full scale is still not accurate. The reason is the
difficulty of taking into account the influence of the
Reynolds number, and water quality on the results.
Unclassified studies on the form of the hub cap were
very limited. This could possibly be attributed to
investigations being carried out within a project
conducted by a propeller manufacturer. Published
results cover only a very limited number of
parameters (Missalek, 1966).

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

AIM OF STUDY

EXPERIMENTAL STUDY

The main aims of the study were focusing on the


characteristics of the flow of hub vortices and
analysing the influence of the different cap shapes on
the propeller efficiency. Another important aim of the
study was to investigate the possibility of using CFD
results to assist correlation procedures for the
prognostic of the cavitation inception of the hub
vortex for full-scale. For this reason, the pressure
reduction in the core of hub vortex was calculated for
model and full-scale.

The model tests were conducted in the cavitation


tunnel of the Potsdam model basin. The cross section
area of the measuring section is equal - 850 x 850
mm. The following tests were carried out: propeller
open water test, observation of cavitation inception
and velocity measurements using LDV and PIV
methods. All tests were performed for every cap
shape. The open water tests were repeated in the
towing tank. The Reynolds number of the open water
test varied between 0.98x106 and 2.4x106.
The velocity measurements were carried out behind
the three shapes at two operation points J= 1.0 and
J=1.06. For the comparison with the numerical results
phase averaged values of the LDV measured data
were calculated in 2 degree steps, which means that
the mean values of the three velocity components
were available for LDV results at 180 angular
positions of the propeller (Abdel-Maksoud et. al.,
2002). Similar LDV measurements were carried out
with a FP propeller model also having 3 different cap
shapes at two cross sections each behind the caps
(Blaurock and Lammers 2001). These tests
demonstrated significant differences in the velocity
field caused by the different shaped caps, as also
found in the present tests.

All model tests as well as all calculations were


carried out for uniform inflow condition.

OBJECTS OF RESARCH
In the present study the viscous flow on a geometry
of a propeller with three different cap shapes was
investigated. A 5 bladed controllable pitch propeller
was selected for the investigations. The propeller has
a typical radial load distribution as normally used for
a navy ship to delay cavitation inception (strong deloaded at blade tip and root). The hub diameter ratio
and skew angle equal 0.2976 and 22.7 respectively.
The scale ratio of the propeller model is 12. Main
parameters of the propeller model (VP 1352) are:

EXPERIMENTAL RESULTS
D
Z
Pm/D
Ae/Ao

= 350.0 mm
=
5
= 1.279
= 0.72

The three hub cap shapes investigated have different


characteristics. The first has a concave shape and is
named CON. The second is a divergent hub cap,
named DIV, and the last is a mixed form between the
first and the second shape. The forward part of this
hub cap is concave and the after part is divergent.
Therefore, it has been named CONDIV.
The selected shapes of the hub cap are based on the
results of a previous intensive investigation
(Blaurock, 1987).
The geometry of the investigated propeller and the
hub cap shapes are shown in Figures 1 and 2.
All numerical and experimental tests were carried out
only for design pitch condition.

The results of the open water tests are shown in


Figure 3. The measured efficiency of the propeller
with CON and CONDIV hub cap are similar
(o=0.70) (CONDIV slightly better than CON, but
within the measuring accuracy). The measured
efficiency of the propeller with CON and CONDIV
hub cap is higher than that measured with DIV shape
(o =0.67). The same results have been obtained in
the towing tank.
Cavitation model tests were carried out at 63% and
42% air content. Figure 4 shows the cavitation
buckets of the propeller with the three different hub
cap shapes at 63% air content. The operation curve of
the propeller at different thrust coefficients is also
included. Hub vortex cavitation inception has been
defined as the condition when cavitation occurred in
5% of the observation time. The intersection points
between hub vortex cavitation buckets and the
operation curve of this particular propeller show that
the DIV-shape has the worst cavitation behaviour
(hub vortex starts at about 20 kts). The cavitation

2
Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

inception behaviour of the CONDIV-form is the best,


it starts at about 29.5 kts. The cavitation inception of
hub vortex of CON shape takes place at about 26 kts.
The experimental results show additionally high
dependency on the thrust loading coefficient. At
KT=0.25 the differences in the cavitation behaviour
of the different hub cap shapes are very small, but
this is an off-design operating point. According to
these results, it may be concluded that under other
boundary conditions (i.e. higher loaded propeller,
other radial blade load distributions etc.) the most
suitable hub cap form might be other than the shape
found in this particular case.
First occurrence of the hub vortex cavitation with hub
caps DIV and CONDIV could be observed at about
2 D to 4 D behind the cap. When decreasing the
cavitation number, the position of first occurrence of
the hub vortex cavitation moves towards the cap until
about 0.2 D behind it.
Cavitation of hub vortex of the cap CON was
observed at 0.2 D behind the hub only. The cavitation
did not touch the cap at any investigated cavitation
number. That is the reason why Figure 4 includes
only one curve for far and close cavitation inception
of the CON hub vortex. Cavitation occurred at the
same time in a region extending between 0.2 D to
about 3 D. The best cavitation behaviour of the CON
shape was measured in this range, when the propeller
was operating at J values higher than 0.8 or KT
values lower than 0.25 respectively, which is the
design operating range of the tested propeller.
All other cavitation types of the propeller were very
similar for the three different caps.
The experimental results showed that the effect of the
air content on the observed cavitation behaviour did
not have the same tendency for all operation points of
the propeller. At J-values of 0.8 or lower , the
cavitation of the hub vortex started at higher pressure
with a gas content of 63 % rather than with 42 % . At
higher J-values cavitation an opposite tendency was
observed, which means that hub vortex cavitation
inception occurred with 63 % at lower pressure rather
than with 42 %.
The results given in this paper are taken with 63%
gas content.

NUMERICAL COMPUTATION
The commercial CFX-TASCflow code has been
applied to solve the Reynolds-averaged NavierStokes equations. CFX-TASCflow is based on a
conservative finite volume method. CFX-TASCflow
uses a non-orthogonal, block-structured numerical
grid, in conjunction with Cartesian velocity
components. For a description of the application of
CFX-TASCflow method for propeller flow see
Abdel-Maksoud and Heinke, 2002.
The applied discretization method of the convective
terms in transport equations is based on the MassWeighted-Skew-Upwind-Differencing
method
(MWS) and the Linear-Profile-Skew-Upwind
differencing procedures (LPS). The accuracy of the
calculated convective terms is improved with the help
of the Physical Advection Correction (PAC) method.
Multi-grid technology is applied to reduce long-wavy
errors and to accelerate the solution of the algebraic
equation system. While the masses and three impulse
conservation equations in every iteration are coupled
solved, the turbulence equations are solved
individually. The coupled solution of the masses and
impulse conservation equations has many advantages
especially for a complicated flow behaviour and it
leads to a robust, reliable and quick algorithm.
To solve the Navier-Stokes equation for the propeller
the solution domain is divided into two regions. The
outside region is stationary and the region inside,
which includes the propeller, rotates. Between the
inside and the outside region a sliding non-matching
interface is applied (Abdel-Maksoud and Heinke,
2002). For turbulence modelling the SST model is
applied.
Especially in the separated flow region for example
behind the hub, more accurate results can be achieved
with the SST turbulence model. Although the
calculations were carried out for a propeller in
homogeneous flow, all propeller blades are
considered in the study. This was necessary to focus
the interaction between the root vortices of the
propeller blades and the unsteady behaviour of the
hub vortex flow.
The applied numerical grid Control contains about
1.3 Mio. control volumes and most of them are
located in the hub region in order to capture the detail
of the flow in this region. The numerical grid is

3
Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

shown in Figures 1 and 5.


The applied numerical method was validated for
propeller flow calculations (Abdel-Maksoud, Menter
and Wuttke, 1998). It was also applied to calculate
the unsteady interaction between hull and propeller
(Abdel-Maksoud, Rieck and Menter, 2000).
In most of the pervious studies the main interest was
the flow around the propeller blades. In the present
investigation the capture of flow behind the propeller
hub was main aim of the calculations. The flow in
this region is very complicated due to the interaction
between the root vortex of every propeller blade and
the separated flow region behind the hub. The
measured velocity results in model scale were used to
validate the numerical results in this region.

NUMERICAL RESULTS
The calculated efficiencies of the propeller show the
same tendency as the measured data. The efficiencies
of the propeller with CON and CONDIV hub caps are
close to each other and much higher than that with
DIV shape. The absolute efficiency differs from those
measured. This can be explained by the fact that the
number of grid cells on the outer radii of the propeller
blades was kept to a minimum in order to be able to
apply higher grid resolution at the inner radii of the
propeller blades and behind the hub. The total
number of cells was kept below a certain limit to
enable a large number of numerical computations to
be carried out within an acceptable time.
The flow behaviour around the CONDIV shape can
be seen in the Figures 6-8. The low pressure regions
in the flow are shown in Figure 6. These regions are
located on the suction side of the propeller blade and
directly behind the hub cap as well as at a certain
distance behind it. At this location the root vortices
are coincided together and form a strong hub vortex
with low pressure region. The streamlines near the
root region of the blades are shown in Figure 7. The
swirl component dominates the flow in this region. A
separated flow region is located directly behind the
hub cap. The shape of the streamlines is changed
suddenly behind the separated flow region as shown
in Figure 8.
A comparison of measured and calculated axial
velocity component in a longitudinal section along
the rotation axis is included in Figure 9. Results of
the LDV and the PIV measurements as well as the

numerical calculations for model scale are presented


in this figure. The axial velocity component is
normalised by the parallel inflow velocity. The dark
blue areas or contour lines indicate separated flow.
The red areas mean that axial velocity component is
equal to 90% of the normalised velocity or higher.
Good agreement was achieved between calculated
and measured velocity contours for each hub cap.
While the contours of the numerical and the PIV
results are for one angular position, the results of
LDV measurements are the mean value of many
measured data at the same angular position. This is
the reason for the unsteady character of the numerical
and the PIV results.
The LDV, PIV and numerical results show that the
shape variant DIV has the largest separation flow
region behind the hub cap. This region exists along
the axis of rotation of the propeller for all
investigated hub cap forms. The smallest separation
flow region is located behind the hub cap CONDIV.
The measured LDV velocity components and those
calculated were compared at different cross sections
plans behind the propeller. Figure 10 shows the
results for the axial velocity component at 0.1 D
behind the propeller. The results of the LDV
measurements are plotted in black and the numerical
results are plotted in blue.
The calculation results of the three geometry forms at
r/R= 0.7 agree well with those measured. In this cross
section, the maximum axial velocity component of
the propeller flow varies between 1.2 - 1.4 of the
inflow velocity. The highest acceleration occurs
behind the shape DIV. This is because the increase of
the diameter of the hub cap behind the propeller blade
reduces the available area for propeller stream. This
may be the reason for the reduced propeller
efficiency in combination with the hub cap shape
DIV.
In the region very close to the root of the propeller
blades r/R= 0.16, the flow is de-accelerated. The
calculated and the measured values show back flow
region behind the form variants DIV and CON. The
axial velocity component behind the form CONDIV
is still very low but it is higher than that behind the
other two shapes.

4
Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

The DIV shape produces a large separation region


behind it which enables the blade root vortices to
keep distance from each other for a long distance
behind the hub cap in comparison to CONDIV.
Behind CONDIV shape the individual root vortices
may unify with each other and cause as a
consequence locally higher circulation, which leads
to lower pressure, and hence earlier cavitation
inception. Furthermore, the flow becomes extremely
intermittent. Small differences in the local geometry
of the model propeller blades may initiate small
differences in the flow downstream of the different
blades. However, the calculation results show less
intermittent flow. This is due to the fact that the
geometry of the propeller blades for the all five
blades is identical. The capturing of this intermittent
behaviour of the flow behind the hub was the reason
for considering all propeller blades in the
computations and for applying a fine grid behind the
hub.

flow region behind all investigated hub cap shapes in


the model scale shows a different characteristic as
compared to the full-scale study.

Several factors may be responsible for the differences


between the model test and numerical results. Many
sources of uncertainty can reduce the accuracy of the
results of numerical computations for separated flow,
such as: discretization errors due to a limited number
of control volumes used and applied discretization
scheme and modelling error due to the limited
accuracy of the applied turbulence model and the
neglecting of the effect of the air content on the flow
in the computations.

Figure 13 shows the contours of the normalised


pressure pNORM of the three hub cap forms in the
longitudinal section along the axis of rotation.

An important source of uncertainty in carrying out


cavitation model tests is the difficulty to define
exactly the cavitation inception by observation. The
quality of the results depends on the experience of the
observer. Furthermore, there is a hysteresis effect
concerning the cavitation inception. This means the
cavitation inception pressure depends on the
experimental procedure i.e. whether the pressure in
the cavitation tunnel decreases or increases during the
model test. Even after taking all these shortcomings
into account, the results of the calculations seem very
promising.

CORRELATION
BETWEEN
THE
CALCULATED RESULTS FOR MODEL
AND FULL-SCALE
The velocity distribution in the longitudinal section
along the rotation axis is shown for model and fullscale in Figure 11 and 12. The shape of the separated

At full-scale in all investigated hub cap shapes, there


is a ring-shaped vortex directly behind the edge of the
hub cap. The diameter of this ring is proportional to
the diameter of the hub cap at the end. The numerical
results of the model scale show that the centre line of
the hub vortex does not lie on the axis of rotation of
the propeller, but oscillates around this axis with an
increasing amplitude. The oscillation of the hub
vortex in the model scale is a result of strong
interaction between the root vortices of propeller
blades due to the short distance between them. A
small disturbance is sufficient for two root vortices to
coincide together. This disturbs the symmetry of the
system and therefore it starts to oscillate. The
oscillation of the hub vortex of CON and CONDIV
shapes is stronger than that of the DIV.

pNorm =

p
0.5 (V + (0.7 D n ) 2)

(1)

The pressure reduction on the suction side of the


blades, the increase of pressure on the pressure side
and the low pressure region behind the hub cap can
be seen in Figure 13. The comparison of the
numerical results for the model and full-scale shows
that the pressure contours of the full-scale are much
smoother than those in the model. The reduction of
normalised pressure behind the full-scale is much
higher than in the model. This effect is much stronger
for the hub cap DIV in comparison to the other two
shapes.
Table 1 includes the normalised minimum pressure in
the centre of the hub vortices. The numerical results
show that the pressure reduction in the full-scale is
much stronger than in the model scale. The calculated
lowest pressure reduction for the first operation point
at full-scale takes place with the hub cap CON. That
is the case in both model and full-scale.
The calculated normalised pressure in the centre of
the hub vortices confirms that the use of the
McCormick formula is necessary for the
extrapolation of the experimental results to the fullscale (McCormick, 1962).

5
Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

The calculated pressure values were used to


determine the exponent of McCormick formula as
shown in Table 1.

(p )
(p )

(Rn )full - scale.


Norm full- scale =

(Rn )model
Norm model

(2)

Table 1: Calculated normalised minimum pressure


values in the centre of the hub vortices and the
corresponding McCormick exponent.

DIV
CON
CONDIV

Model
pNorm
-0.06
-0.045
-0.058

Full-scale
pNorm
-0.225
-0.131
-0.155
mean value

Exponent
m
0.40
0.32
0.31
0.34

The mean exponent value was determined for one


propeller geometry only and for one radial thrust
loading distribution. Therefore, the value of the
exponent m may be not applicable for all propeller
and hub cap geometry.
CONCLUSIONS AND RECOMMENDATIONS
The effect of the shape of the hub cap on propeller
efficiency as well as on hub vortex cavitation
inception is significant. It is therefore recommended
to pay more attention to the shaping of the cap of the
hub, especially when operating with low noise is
important.
The results of the study confirm that the highly
sophisticated velocity- and pressure field behind the
hub of an operating propeller can be predicted by
applying numerical methods for viscous flow
calculation. These methods are a good tool for
predicting the pressure reduction in the hub vortex
and therefore it is possible to apply CFD-methods to
find out the most suitable hub cap shape. This is valid
for model as well as for full-scale conditions.

Differently shaped hub caps may have a significant


influence on propeller efficiency. A divergent cap has
a negative influence. A moderate convergent
influence such as CON and the special shape
CONDIV may have a positive influence.
This statement is valid for propellers with a relatively
high hub diameter ratio, which is the typical case for
CP propellers. For fixed pitch propellers having an
hub diameter ratio of 0.2 or even less the difference
between different cap shapes can be expected to be
small and may be ignored. Hence the cap shape for
such propellers should be selected only with regard to
hub vortex cavitation inception of the hub vortex.

ACKNOWLEDGEMENT
The authors would like to express their gratitude to
the Federal Office of Defence, Technology and
Procurements, Department Ships and Naval
Equipment for the support of this study.

LIST OF SYMBOLS
Ae/Ao
D
dh
J
KT
KQ
kn
m
n
Pm
PS
pNorm
r
R
Rn
SS
V
Z
o

=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=

expanded area ratio of the propeller


propeller diameter
hub diameter
advance ratio = V/(n *D)
propeller thrust coefficient
propeller torque coefficient
knots
exponent in McCormics formula
number of revolutions
mean pitch of propeller
pressure side
normalised pressure
local radius
propeller radius = D/2
Reynolds number
suction side
mean inflow velocity
number of propeller blades
propeller efficiency = J/(2*) *KT/KQ
density of fluid (water)
cavitation number

Furthermore, the results of the calculations showed


that the cavitation inception of hub vortices with
model tests (at low Reynolds numbers) should be
extrapolated to full scale using the formula as
suggested by McCormic.

6
Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Basin, Report No. 2839, , October 2002 (in German,


Classified).

REFERENCES
Abdel-Maksoud, M., Menter, F. R., Wuttke, H.,
"Viscous Flow Simulations for Conventional and
High Skew Marine Propellers," Ship Technology
Research, Vol. 45, No. 2, 1998.
Abdel-Maksoud, M., Rieck, K., Menter, F. R.,
"Unsteady Numerical Investigation of the Turbulent
Flow Around The Container Ship Model (KCS) with
and without Propeller," Gothenburg 2000, A
Workshop on Numerical Ship Hydrodynamics,
Gothenburg, Sweden, 2000.
Abdel-Maksoud, M., Heinke, H., "Scale effects on
Ducted Propellers," 24th Symposium on Naval
Hydrodynamics, Fukuoka, Japan, July 2002
Abdel-Maksoud, M., Hellwig, K., Blaurock, J.,
Schmidt, D., Jaksic, D., "Correlation of Propeller
Cavitation Design of Hub Cap," Potsdam Model

Blaurock, J., "Influence of the Propeller Hub Cap


Shape on Hub Vortex Cavitation," 6. US-GEHydroacoustics Symposium, Annapolis, MD.,1987
(Classified).
Blaurock, J., Lammers, G., "Velocity Field behind
the Propeller Hub with different shaped Caps,"
HSVA Report No. 1559, Hamburg, Sept. 2001 (in
German).
McCormic, jr. B.W. "On Cavitation Produced by a
Vortex Trailing from a Lifting Surface," Journal of
Basic Engineering, Sept. 1962
Missalek, R., "Influence of the Form of the Hub Cap
on Open water and Cavitation Characteristic," VWS
Berlin, Report Nr. 325,1966 (in German).

DIV-shape

CON- shape

CONDIV- shape
Figure 1: Investigated propeller geometry

Figure 2: Geometry of hub cups

7
Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DIV - KT
CON - KT
CONDIV - KT

DIV - 10KQ
CON - 10KQ
CONDIV - 10KQ

DIV - ETA0
CON - ETAO
CONDIV - ETAO

1,1
1,0
0,9

10K Q

KT, 10KQ, 0 [-]

0,8
0,7
0,6

0,5

KT

0,4
0,3
0,2
0,1
0,0
0,0

0,1

0,2

0,3

0,4

0,5

0,6

0,7

0,8

0,9

1,0

1,1

1,2

1,3

J [-]

Figure 3: Open water test

Figure 4: Cavitation bucket curves

8
Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Figure 5: Numerical grid

Figure 6: Low pressure regions

Figure 7: Streamlines in the root region

Figure 8: Streamlines of the hub vortex

LDA

PIV

CFD

DIV

CON

CONDIV

Figure 9: Comparison of measured and calculated axial velocity component, longitudinal section

9
Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DIV, r/R=0.71

DIV, r/R=0.16

CON, r/R=0.71

CON, r/R=0.16

CONDIV, r/R=0.71

CONDIV, r/R=0.16

Figure 10: Comparison of measured and calculated axial velocity component, cross section

10
Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Model

Full
scale

DIV

CON

CONDIV

Figure 11: Comparison of calculated axial velocity contours for model and full-scale

Model

Full
scale

DIV

CON

CONDIV

Figure 12: Comparison of calculated axial velocity vectors for model and full-scale

Model

Full
scale

DIV

CON

CONDIV

Figure 13: Comparison of calculated pressure field for model and full scale

11
Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Stuart Jessup
Naval Surface Warfare Center, Carderock Division,
USA
I would like to congratulate the authors for
addressing a topic for which Navy propeller
designers have great interest. Design methods for hub
vortex performance are presently very crude. The
design community can benefit from any experimental
or computational efforts in this area. The conclusion
of the application of McCormick type sealing
supports our general approach.
Our experience with the use of divergent
hub cones is similar in respect to performance loss.
We installed such a hub cone on our PC-1 patrol boat
propeller which resulted in an efficiency loss. One
positive result was an improvement in thrust
breakdown. The divergent cone appears to raise the
static pressure in the propeller plane, thus reducing
the amount of blade cavitation at high speed. Can the
authors comment on this aspect?
Also, can the authors explain why the RANS
computations used in the paper have required a full
propeller disk simulation? Would it be possible to
compute the hub vortex pressures by modeling only
one blade passage? Would accuracy be sacrificed by
making that simplification?
We appreciate the contribution that this
paper has made for the propeller designer tasked with
addressing hub vortex cavitation requirements.

the geometry of a propeller is always a compromise


depending on the numerous sometimes very opposite
requirements.
The consideration of all propeller blades in
the numerical study was necessary to study the
interaction between the root vortices of the propeller
blades and the unsteady behaviour of the hub vortex
flow. The numerical results show that interaction
between the root vortices in model scale is much
stronger than in full-scale. Therefore it is important to
consider all blades of the propeller in model scale
calculations. The consideration of all propeller blades
in full-scale calculations was necessary to have a
reasonable comparison between model and full-scale
results.

AUTHORS REPLY
Thank you very much for your kind
comments.
The main interest of this task was focussed
on the structure of the velocity field behind the hub,
the cavitation onset and the influence of the different
hub shapes on the propeller efficiency. Hence, the
influence on thrust break down and the corresponding
influence on blade cavitation have not been
investigated.
We agree with you that the divergent cone
raises the static pressure in the propeller plane. A
well known proceeding in design of propellers for
fast vessels is to reduce the hub diameter at half boss
length and to increase it again toward the hub end.
The aim of this proceeding is to raise the pressure at
the route fillet locally to delay route cavitation. This
is the same physical mechanism, which can be
observed on the divergent cap.
Unfortunately this positive effect cannot be
achieved, when we look for late cavitation onset of
the hub vortex. Necessarily, we have to accept that

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Soave Massimo
INSEAN, Italy
I have appreciated the work presented
especially because it has confirmed a similar
investigation carried out at the CEIMM (Italian Navy
Cavitation Tunnel) a few years ago. All the results
presented are perfectly congruent with our previous
work with the exception of the proposed correlation
between model and full scale; extensive comparison
of model cavitation with full scale investigation
carried out by the Italian Navy Cavitation Tunnel
have several times shown that the hub vortex
cavitation (at least the type attached to the hub) is
subject to a very small scaling effect. Could the
authors of this paper say if they have validated their
numeric theory with full-scale sea investigation?
AUTHORS REPLY
Thank you very much for your encouraging
comments. In the paper, the viscous flows around a
propeller geometry in model and full-scale were
calculated and the estimated minimum pressures in
model and full-scale were compared. The comparison
shows that it is possible to apply the formula
developed by McCormick to calculate minimum
pressure in the hub vortex region of the full-scale
from the model scale results.
The calculated results of the velocity
distribution behind the different hub cap forms were
compared with the measured data in model scale.
As you know, in full-scale case, the
observation of the hub vortex is much more difficult
than the tip vortex due to many reasons. In addition it
should be mentioned that the location of lowest
pressure was found not attached to the propeller cap
but at a significant distance behind it. Just this area is
extremely difficult to observe by optical means and
even cavitation of the hub vortex is existent there.
Unfortunately, there were no measurements
available for a validation study of full scale results.
We would appreciate a lot if it would be possible to
get validation data for this part of the study.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Christopher Chesnakas
Naval Surface Warfare Center, Carderock Division,
USA
Was the root flow at the blade examined
either experimentally (through oil flow) or
numerically for flow separation, and if so, did the
shape of the boss cap have an effect on the root
separation?
AUTHORS REPLY
A so called wet paint (or oil flow) test has
not been carried out at any of the tested conditions.
Although this proof is not available, it should be
expected that especially the divergent (DIV) cap
shape will retard slightly the flow close to the hub
surface and hence the pressure will increase in this
region. (See also the contribution of S. Jessup and the
corresponding answer).
The LDA and PIV results are available only
for the flow behind the propeller blades. Therefore, it
was not possible to study the effect of the hub cap
shape on the flow in the root region.
The numerical results do not show any flow
separation near the roots of the blades. Perhaps the
reason is the applied ideal inflow condition to the
propeller in the study. The propeller was investigated
in parallel flow and the operation point was close to
the design condition.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Hoyte C. Raven
MARIN, Netherlands
The paper concludes that numerical methods
are a good tool for predicting the pressure reduction
in the hub vortex, and can be used to find the most
suitable hub cap shape. However, the calculated
minimum pressures in Table 1 seem not in agreement
with the cavitation inception test results. What
evidence do you have that you can actually predict
the pressure in the hub vortex?
AUTHORS REPLY
The calculated velocity distribution behind
the different hub cap forms were compared with the
measured data in model scale. The results of the
calculated minimum pressure for the CON and DIV
forms agree well with the cavitation inception test
results . The reason for the deviation of the results of
CONDIV is that the calculated minimum pressure of
this hub cap form is strongly effected by the unsteady
behavior of the hub vortex due to the highly
interaction between the root vortices of the blades.
Because this effect takes place only in model scale, it
should be expected that a good agreement can be
achieved for full-scale case.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics


St. Johns, Newfoundland and Labrador, CANADA, 8-13 August 2004

A B-Spline Based Higher Order Panel Method for Analysis of


Steady Flow around Marine Propellers
C.-S. Lee1 , G.-D. Kim1 and J. E. Kerwin2
(1 Chungnam National University, Korea, 2 Massachusetts Institute of Technology, U.S.A.)
ABSTRACT

method has been successfully applied to propeller problems by many researchers(Lee 1987, Hoshino 1989). A
good discussion on the potential-based panel method may
be found in Kerwin et al(1987). The existing potentialbased method, called the low-order panel method, assumes that the potential is constant over a panel, and
hence to get the velocity distribution on the body surface,
the method requires a finite difference scheme which inevitably introduces a numerical differentiation error. This
error is most significant at/near the trailing edge and the
tip of the lift-generating surface, and leads ultimately to
the degradation of the accuracy of the low-order method.
The main purpose of the present paper is to develop
a higher-order panel method, which improves the prediction of the velocity and pressure in these regions. We
will employ B-spline basis functions(Rogers and Adams
1989, Piegl and Tiller 1996) to represent both the geometry and the potential. Since the derivatives of the basis functions can be obtained exactly, there is no need to
use a numerical differentiation scheme to get the velocity
from the potential, and hence it is obvious that the inherent limit of the low-order panel method would no longer
exist. The order of the B-splines to represent the body
and the potential can also be increased without limit, and
hence the solution of any order can be obtained.
The most detailed description of the higher-order
panel method based on B-splines was first given by Hsin,
Kerwin and Newman(1993) at M.I.T. for the analysis of
the flow around two-dimensional bodies. Their theory
is mathematically complete and the numerical procedure
is reported robust. Maniar(1995) then extended to the
three-dimensional higher-order panel method, following
the similar line of approach as Hsin et al. The threedimensional higher-order panel method proved to be very
accurate for their test cases of spherical floating bodies
and floating offshore structures with vertical cylindrical
columns. In their approach, the integrals of the influence functions are expressed in terms of polynomials of
a parametric coordinate, and the polynomial coefficients
are derived from B-spline basis functions. Although the
polynomial representation is sufficient for most bodies
of interest, it is in general not appropriate to expand in

A higher order panel method based on B-spline representation for both the geometry and the solution is developed for the analysis of steady flow around marine
propellers. The self-influence functions due to the normal dipole and the source are desingularized through the
quadratic transformation, and then shown to be evaluated
using conventional numerical quadrature. By selecting
a proper order for numerical quadrature, the accuracy of
the present method can be increased to the machine limit.
The far- and near-field influences are shown to be evaluated based on the same far-field approximation, but the
near-field solution requires subdividing the panels into
smaller subpanels continuously, which can be effectively
implemented due to the B-spline representation of the geometry. A null pressure jump Kutta condition at the trailing edge is found to be effective in stabilizing the solution
process and in predicting the correct solution. Numerical
experiments indicate that the present method is robust and
predicts the pressure distribution on the blade surface, including very close to the tip and trailing edge regions,with
far fewer panels than existing low order panel methods.
INTRODUCTION
Panel methods have been widely accepted as a useful tool
for aerodynamic and hydrodynamic design, since the pioneering work of Hess and Smith(1964). A large number
of different panel methods have been developed for a variety of applications(Hess 1975), leading recently to application to the analysis of steady performance of marine
propellers(Hess and Valarezo 1985). Until Morino(1974)
introduced a panel method based on Greens formula in
which the primary unknown is the potential, most of the
previous works were based on the velocity-based formulation in which the boundary condition on the body surface is satisfied through the direct computation of the velocity. Morinos potential-based formulation is known
to be more stable and hence more suitable to numerical
computation than the velocity method, since the potential is one order less singular than the velocity. Morinos
1

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

polynomials the non-uniform rational B-spline(NURBS)


surfaces, which are the de-facto standard in industry to
represent complex surfaces like ship hulls.
Our ultimate target geometry in developing yet another three-dimensional higher-order panel method is the
NURBS surface. In a recent report(Lee and Kerwin
2003), we introduced a very general numerical procedure
for the evaluation of influence coefficients for the analysis of the lifting flow around two-dimension foils. The
approach was made possible by removing the singularities in the induction integrals for both the normal dipole
and source panels. The desingularization process for
two-dimensional influence functions was proven robust
through a systematic test with bodies of various profiles.
The numerical procedure of Lee and Kerwin(2003) is extended in this report to the analysis of the flow around
three-dimensional steady lifting bodies, including marine
propellers.
The desingularization procedure in threedimensional problem is taken from the first pioneering
work of Maniar(1995). Instead of following Maniars
approach relying upon the polynomial expansion of
various quantities, we adopt a conventional numerical
integration method such as Gauss-Legendre quadrature.
The approach is very simple mathematically and numerically, and yet there is no sacriface in the accuracy in
evaluating the induction integrals. The coding requires
far less efforts than that of Maniar.
Numerical procedures are tested with a simple Bspline surface and validated by comparison either with
the exact solution or with the low-order panel method
represented by sufficiently large number of panels. The
method is then applied to propellers with experimental
results available. The comparison of the predicted pressure distribution on the blade surface indicates that the
new numerical B-spline based higher-order method can
be a better alternative for the analysis of the steady flow
around marine propellers.

Let us consider a three-dimensional surface patch expressed in parametric form


~x = ~x(u, v)

(1)

where u and v are two independent parameters which


monotonically increase along the respective parametric
spaces. We will define the u parameter space first, and
the v parameter space may subsequently be defined in the
same manner.
The range of u = [0, umax ] is called the usable
g
parametric space, and we will divide this space into N
spans of equal parametric lengths, placing internal knots
g 1}. The usable parametric space
at u = {1, 2, , N
is then generalized to form a space by including (p + 1)
zeroes and (p + 1) umax values at the ends of the usable
parametric space resulting in a uniform knot vector of the
form
~ g = {0, 0, 0, 0, 1, 2, , N
g, N
g, N
g, N
g }T
U
(2)
or in normalized form for umax = 1.0
~ g = {0, 0, 0, 0, 1/N
g , 2/N
g , , 1, 1, 1, 1}T
U

(3)

in the particular case when p = 3.


We may also define in the same manner the normalized parametric space for v, resulting in a uniform knot
vector of the form
~ g = {0, 0, 0, 0, 1/M
g , 2/M
g , , 1, 1, 1, 1}T
V

(4)

in the particular case when q = 3 with q + 1 multiplicity


at both ends.
We will adopt the (p, q)-th degree B-spline basis
functions to represent a surface patch and then express
the surface as a weighted sum of tensor product B-spline
basis functions as
~x(u, v) =

v
v 1 M
N
1
X X

i=0

i (u)M
j (v)
~xvi,j N

(5)

j=0

i (u) are the p-th degree B-spline basis functions,


where N
j (v) the q-th degree B-spline basis functions, ~xv the
M
i,j
v and M
v the number
geometric control vertices, and N
of geometric control vertices in (u, v) parametric direci (u) and M
j are in general rational
tions, respectively. N
functions of the parameters u and v, but in this paper we
will consider only the non-rational (integral) B-splines.
The approach developed here can be directly applied to
geometry represented by NURBS without any difficulty.
Our ultimate goal is to solve for the velocity potential
(u, v) along the surface of the body. However, instead
of treating the potential directly, we will represent the potential as a weighted sum of tensor product B-spline basis
functions in a similar form as for the geometry as

B-SPLINE GEOMETRY AND POTENTIAL REPRESENTATION


The boundary surface may be composed of patches, over
each of which two parameters (u, v) are taken as independent variables for describing the geometry and the potential by B-spline representations. The patches may consist of multiple numbers of blade surfaces, trailing wakes
and hub surfaces, without any overlapping between them.
Since the geometric characteristics and the solution properties are similar, only one set of blade and its trailing
wake surfaces will be considered in the numerical formulation. It is straightforward and intuitive to further include
influences of the remaining patches and interactions between them.

(u, v) =

v
v
NX
1 MX
1

i=0

2
Copyright National Academy of Sciences. All rights reserved.

j=0

vi,j Ni (u)Mj (v)

(6)

Twenty-Fifth Symposium on Naval Hydrodynamics

where Ni (u) and Mj (v)are the B-spline basis functions,


vi,j the potential control vertices, and N v and M v the
numbers of potential control vertices in (u, v) directions,
respectively. The numbers of potential control vertices
(N v ,M v ) and the basis functions {Ni (u),Mj (v)} may be
different from the corresponding quantities for the geometry, but the usable parametric spaces of the geometry and
the potential should be identical. With the introduction of
the potential vertices, the unknowns of the hydrodynamic
problem are now the values of the potential vertices, vi,j ,
which are not the potential in the physical sense. And the
potential knot vectors in (u, v) are defined, in a similar
way as in (3) and (4) as
~
U
~
V

Y
X

Geometric vertices

= {0, 0, 0, 0, 1/N , 2/N , , 1, 1, 1, 1}

= {0, 0, 0, 0, 1/M , 2/M , , 1, 1, 1, 1}T (7)

when p = q = 3. The numbers of spans or panels in


(u, v)-directions in the usable parametric spaces for the
potential will be (N , M ), and are related with the numbers of potential control vertices by N v = N + p and
M v = M + q.
Figure 1 shows a typical blade surface patch generated with the geometric control vertices and the internal
knot vectors which serve as the potential panel boundary. The blade geometry is generated by the (3, 3)-degree
B-spline basis functions, that is (p, q) = (3, 3) with
v, M
v ) = (11, 11) and (N , M ) = (8, 8).
(N
Figure 2 shows a typical wake surface patch generated. The wake, being a thin sheet of zero thickness, can
be represented easily by fitting in a least square sense the
B-spline net to the offset provided by the modeling based
on the hydrodynamic principle. For example, see Greeley
and Kerwin(1982).

Discretized panels

Figure 1: A blade surface patch generated by


v, M
v) =
B-splines with (p, q) = (3, 3), (N

(11, 11) and (N , M ) = (8, 8)

Y
X

DISCRETIZATION OF INTEGRAL EQUATION

8
7

0.5

6
5

The potential on the body surface is represented by the influences of the normal dipoles and the sources distributed
on the body surface consisting of the blade and the hub
(SB SH ) and the wake surface(SW ) as

ZZ

G
=
+
G dS
()
2
n
n
SB SH
ZZ
G
+
()
dS
(8)
n
SW

3
2

-0.5

1
0

-1
-1

Wake vertices

0
1

1
8
7

0.5

6
5

3
2

-0.5

0
-1
-1

the unit vector


where is the perturbation potential, n
normal to the body and wake surfaces, and G the Green
function defined as
1
G=
4r

Wake panels
1

Figure 2: A wake surface patch generated by B w, v , M


w, v ) =
splines with (p, q) = (3, 3), (N
w,
w,
(84, 11) and (N
, M ) = (81, 8)

(9)

The strength of the dipole in the wake is expressed by


the potential jump at the trailing edge. In the potential
3
Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

based formulation, the normal derivative of the potential


is specified on SB SH from the kinematic boundary
condition as

~r
=
nU
(10)
n
~r = U
~A
~ ~r is the velocity of the circumferwhere U
ential mean oncoming stream past a fixed blade at radius
r, being defined by the circumferential mean advance ve~ A and the rotational speed of the propeller .
~
locity U
As discussed in the preceding section, the blade surface patch may be discretized into (N , M ) panels using the usable parametric space of the knot vectors (7).
The wake surface patch, which sheds downstream from
the trailing edge, can be represented by a set of M
streamwise strips with the constant normal dipole on each
spanwise location.
Discretization into a set of (N , M ) panels on the
blade and M strips in wake will then yield
ZZ
X
G
+

dS
2
n
S,
,
X ZZ
X ZZ
G

dS =
GdS
()
+
n
w
S
S, n
,

summation for the dipole over the panels on the blade


in (13) includes the case of i = and j = . In the
low order panel method, this term drops out, since the effect is already considered by the subtended angle of the
hemisphere surrounding the point where the potential is
evaluated. In the higher-order panel method, there are
additional effects from the curvature of the geometry and
the higher order variation of the potential in addition to
the subtended angle effect. This will be described in detail later.
DESINGULARIZATION OF INDUCTION INTEGRALS
When the control point falls within the panel boundary,
special care is necessary to avoid the singular behavior
in computing the influence integral, especially when the
coefficient of B-spline basis functions is constant in (13).
The induced potentials, I0S and I0D , due to a source and
a dipole of constant strength, respectively, are of the following form
ZZ
1
1
dS
(14)
I0S =
4
r
ZZ
1

n ~r
dS
(15)
I0D =
4
r3

(11)
Noting that there are only (p + 1, q + 1) nonzero
basis functions at each span defined by the space between
adjacent knots in (7), we may rewrite the potential (6) as
a (p + 1, q + 1) term summation as follows
(u, v) =

q
p X
X

v, N (u)M (v)

where the distance vector ~r is defined by a vector extending from the control point to the source point on the panel
surface and is expressed by a series of B-spline tensor
products as
X
(u)M
(v)
~r =
(~xv, ~xc )N

(12)

a=0 b=0

a,b

where the subscript (s, t) are the span indices satisfying


the relation (u, v) ([us , us+1 ), [vt , vt+1 )) in the knot
vectors (7), = s p + a and = t q + b. We will add
the subscripts (, ) to the span indices (s, t) to indicate
the associated panel indices (, ) as in s and t , when
necessary. If the above equation (12) is inserted into (11),
we obtain for the control point on the (i, j)-th panel the
following

|~r|

(16)

The derivatives of the above with respect to the parametric variables can be obtained analytically from the Bspline derivatives as follows:
~r
u

1 X
{
Ni (ui )Mj (vj )vi ,j }
2
a,b
X ZZ
X
G
+
dS
{
N (u)M (v)v, }
n
S, a,b
,
X ZZ
X ZZ
G

+
()
dS =
GdS
n
n
w
S
S,

~r
v

X
(u)
N
(v)
(~xv, ~xc )
M
u

(u) M (v)
(~xv, ~xc )N
v

a,b

(17)

a,b

Since r vanishes on the surface, the integrals (14)


and (15) are singular, but integrable. In order to remove
the zeroes from the denominators, we partition the square
parametric space into four triangles as shown in Figure 3.
We will repeat here the original desingularization procedure of Maniar(1995) for completeness. The apex of each
triangle will be located at the collocation point and their
base at the sides of the square. We denote the triangles

(13)
where i = si p + a, j = tj q + b, = s p + a
and = t q + b. It should be noticed that the (, )
4

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

will give

v
Transformation
(0)
for
(2)
( uc , vc )

u - uc =
v - vc =

(0)

u - uc

u
=0

v
=

u
=1

v
=

(24)

(0)

(1)

The derivatives of the distance vector ~r with respect to the


transform variables may then be computed by applying
the chain rule as

(3)
u

~r

~r

Figure 3: A typical triangulation of the parametric space and the quadratic transformation, after
Maniar

For (1) ,
For (2) ,
For (3) ,

u uc
v vc
u uc
v vc
u uc
v vc
u uc
v vc

=
=
=
=
=
=
=
=

~r
r

(18)

k=0

~r
~r

where J is the jacobian of the surface element dS and can


be evaluated as

~r
~r

(21)
J =

S
I0,
(0)

v vc =

(22)
=

Without loss of generality, we may describe the procedure to get the integral in the rightmost triangle (0)
and this can be equally applicable to the other triangles.
The quadratic transformation for (0)
u uc =

(26)

~
~ + Q
Q

~
Q

(27)

Substituting (26) and (27) into the source integral (19) for
(0) will lead to

and the unit normal vector to the surface element n


can
be computed by
~r
~r

/J

~
Q(u,
v)
Q(u, v)

(20)

n
=

(25)

~ Note that is non-negative within the


where Q = |Q|.
triangle of transformation. This quantity Q is proportional to the ratio of the distance r, between the collocation point and the singular point, and the coordinate ,
and never becomes zero since is factored out from ~r in
(26). When the singular point falls on the -axis and if
the variable is substituted by the curvilinear distance,
the quantity Q will approach to unity, as shown by Lee
and Kerwin(2003) for the the 2-dimensional case.
From (26), we can also compute the vector normal to
the surface element as

With the partitioning, the integrals (14) and (15) may now
be obtained by adding the contribution of each triangular
section as
ZZ
3
3
X
X
1
1
S
Jdd (19)
I0S =
I0,
(k) =
4
(k) r
k=0
k=0
ZZ
3
3
X
X
1

n ~r
D
D
Jdd
I0 =
I0,(k) =
4
r3
(k)
k=0

which may be used in computing the jacobian (21) and


the normal vector (22).
Maniar(1995) then showed that the distance vector ~r
can be expressed by a product of the transform variable
~
and a finite nonzero vector quantity Q

by (k) for k = 0, , 3. On each triangle Maniar introduced the quadratic transformations


For (0) ,

~r u ~r v
+
u
v
~r u ~r v
+
u
v

~r
~r dd

r
(0)

!
ZZ

~
~ dd
1
Q
Q
~

Q+

Q
(0)

1
4

ZZ

(28)
where the integrand of the last expression is non-singular.
Similarly, the quadratic transformation applied to the

(23)
5

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

dipole integral for (0) gives

of the square, respectively, wn and wm the weights of


Gauss
quadrature in - and -directions, and Ng and Mg
D
I0,
the order of the Gauss quadrature in both directions, re(0)
ZZ
spectively. It will be shown later that the order of Gauss
~r
~r ~r
1

3 dd
=
quadrature for the self-induction computation should be
4

r
(0)

!
(Ng )Self 4 and (Mg )Self 4. Application of the simiZZ
~
~
~
1
Q
Q
Q
~
dd lar equation to the other triangles of the square will com=
Q +

||3 Q3
(0)
plete the computation of the self-induced potentials due
(
ZZ
to the source and the dipole.
~ Q/
~
~
Q
Q
1
=
4
Q3
(0)
)

!
SELF-INDUCED POTENTIAL OF DIPOLES OF
~
~
Q
Q
1
~
dd

Q
(29) HIGHER ORDER

Q3
When the control point falls on the panel surface, the inThe numerator of the first term in the integrand is always duction integral due to the normal dipole of higher order
zero, whereas the denominator is nonzero unless = 0. requires a special treatment as described above. The conWe expect that this term will result in the solid angle tribution of the higher orders is, however, less singular
of 2, when the contribution from all four triangles are than that due to the constant term in dipole strength, and
added together as described by Maniar(1995). This is the hence the same argument leading to (32) can be applied
factor which leads to the /2 term in the integral equa- to evaluate the dipole-induced potential of higher-order
tion. With the nonzero denominator in the second term as
in the integral, we observe that the integrand is a reguD
Iself
lar function. The significance of the above expressions
(28) and (29) is that the singular part is removed from the
3 ZZ
3
X
X
G
D
original equations (19) and (20), and hence we can apply

=
I
dS
(k) =
n
(k)
a numerical quadrature to perform integrations.
k=0
k=0
3 ZZ
To compute the integrand of (29) using more familiar
X
X
G
v, N (u)M (v)
=
Jdd
physical quantities rather than Q, we use the relation
n
(k) a,b
k=0
~
~
~
~
~ Q
Q
(33)
~ 1 = ~r ~r ~r + Q Q/ Q

Q)
(
3
3
3

Q
r
Q
(30) The last line of (33) may be rewritten as
All the variables in the first term in the right-hand side
D
of (30) are well defined. The second term is identically
Iself
ZZ
zero, since 6= 0 in our numerical integration for (29).
X
G
v
=
,
N (u) M (v)
Jdd
Equations (28) and (29) may now be computed by
n
(k)
k,a,b
applying Gauss-Legendre quadrature with global variX
ables as follows
=
v,
k,a,b
1 R L T B
S

ZZ
I0,
'
(0)
1

n ~r
4
2
2
Jdd
N
(u)
M
(v)

Mg 1 Ng 1
4
r3
X X
~r
(k)
~r 1

wm wn

X
1 R L T B

r n,m
v,
'
m=0 n=0
4
2
2
k,a,b
(31)
)

1 R L T B
X
D

~
r
I0,
'
(0)
wm wn N (u)M (v) 3 J
4
2
2
r
n,m
m,n

Mg 1 Ng 1
X X
~r
~r ~r
(34)
wm wn

r
n,m
m=0 n=0
(32) Note that computation of (34) does not require any special treatment due to the removal of singularity at the
where L and R denote the parameters at the control origin, and hence the same order of the Gauss-Legendre
point and the rightmost side of the square, respectively, quadrature may be maintained as the order for the conB and T the parameters at the lower and upper ends stant dipole integral (32).
6
Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

SELF-INDUCED POTENTIAL OF SOURCES OF


HIGHER ORDER

triangles is not necessary. Knowing the exact expression


for the strength of the source from the boundary condition, we may directly compute the values required at the
Gauss integration points. The source-induction integral
for the far field point may be evaluated as follows:
X ZZ

IfSar =
GdS
S, n
,
X ZZ

G J du dv
=
S, n
,
X uR uL vT vB
'
2
2
,

)
X

wm wn
(37)
GJ
n
n,m
m,n

Since the source strength /n in (13) is known


from the kinematic boundary condition, the evaluation
of the self-induction due to sources of higher order is
rather straightforward. The self-induction integral due
to sources can be evaluated by applying Gauss-Legendre
quadrature as follows:
S
Iself
3
X

S
I
(k) =

k=0

3 ZZ
X
k=0

(k)

GdS
n

X 1 R L T B
4
2
2
k

X
1
J
wm wn
n r
n,m
m,n

'

(35)

Note the summations over and is substituted by (35)


when the (i, j)-th control point falls upon the self panel,
that is when i = and j = .

FAR-FIELD APPROXIMATION OF THE INDUCED POTENTIAL DUE TO NORMAL


DIPOLES OF HIGHER ORDER

NEAR-FIELD INDUCTION INTEGRALS


The far-field approximation of (36) and (37) is valid when
the field point, where the potential is wanted, is sufficiently far from the panel. When the field point is not
sufficiently far from the panel, the panel may be subdivided into 4 or more subpanels and the induction integrals
may be evaluated for the subpanels applying the far-field
formulation. This partitioning technique is due to Maniar(1995) and will be introduced here for completeness.
We first introduce the criterion, Cr, for the measure of
the far distance, which will be compared with the square
of the ratio of the distance d, from the centroid of the
panel to the field point, and the characteristic length of
the panel `. If (d/`)2 < Cr, the field-point is said in the
near-field.
As shown in Figure 4, if the field point falls in the
near-field, the panel is subdivided into four subpanels by
the average parametric values of all sides of the square,
and the distance from the center of each subpanel to the
field point d is compared with the characteristic length of
the subpanel `. For subpanels which meet the far-field
criterion (d/`)2 > Cr, the induction integrations are performed using the far-field approximation formula. For
other subpanels, the partitioning continues until finally
all the subsequent subpanels meet the far-field criterion.
The partitioning requires computation of geometric
quantities for the newly generated subpanels such as the
characteristic length ` and the distance from the source
point on the panel to the field point. Since all the geometric variables can be computed from their B-spline representation like (16) accurately, the partitioning method is
really accurate, fast and yet very simple. In practice, the

When the control point is away from the dipole panel, it


is possible to apply the far-field approximation directly in
parametric spaces.

IfDar
X ZZ
,

X ZZ
,

S,

X
S, a,b

'

,,a,b

v, N (u)M (v)

(Z Z

v,

,,a,b

G
dS
n

v,

G
Jdu dv
n

G
N (u)M (v)
Jdu dv
n
S,

uR uL vT vB
2
2

)
G
wm wn N (u)M (v)
J
n
n,m
m,n

(36)

where J = |~r/u ~r/u|. Note the summations over


and is substituted by (34) when the (i, j)-th control
point falls upon the self panel, that is when i = and
j = .
FAR-FIELD APPROXIMATION OF THE INDUCED POTENTIAL DUE TO SOURCES OF
HIGHER ORDER
When the control point is away from the source panel,
the integrand is not singular, and hence partitioning into
7

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

condition after some manipulations as

(a)

(c)

2U +
+
2V +
x
y
y

2W +
+
z
z

2U +
+
2V +
x
x
y
y

2W +
=0
(40)
+
z
z

(b)

This equation is nonlinear in perturbation velocity, and


it is necessary to resort to an iterative procedure. With
the introduction of an iterative index for the present step t
and a short hand notation for the quantities at the previous
iteration as

(d)

Figure 4: Continuous partitioning of the parametric space from (a) to (d), after Maniar
dynamic array allocation technique can be applied effectively to handle the storage need for continuously increasing number of subpanels which is unknown until all the
subpanels meet the far-field cirterion.

(t1)

2U +
x
(t1)

2V +
y

(t1)

2W +
z

(41)

Equation (40) may be rearranged with new symbols as


KUTTA CONDITION

To solve the lifting problem, it is necessary to specify


the Kutta condition at the trailing edge. The Kutta condition can be either of the kinematic or of the dynamic
type or both. For the steady lifting problem, the strength
of the wake sheet dipole at the j-th potential control
vertex is equivalent to the potential jump at the trailing
edge on the same radial control vertex, i.e., according to
Morino(1974),
vj = +,v
,v
= vN v 1, j v0, j
j
j

E
+F
+G
x
y
z

=0

(42)

The perturbation velocity components are expressed


in terms of the potential derivatives with respect to the
parameters as shown in Appendix A, and hence the first
term in (42) may take the following form

(38)

where the superscripts + and denote the suction and


pressure sides of the trailing edge, respectively. This
equation and the induction due to the wake sheet dipole
allow the potential jump at the trailing edge, which will
lead to the circulation around the blade and the net thrust
on the blade.
Another form of the Kutta condition is the zero pressure jump condition in the wake sheet. This is equivalent
to the statement that the magnitudes of the total velocity
on both sides of the trailing edge should be identical in
the steady flow, that is,
~ r + + )2 = (U
~ r + )2
(U

+F
+G
x
y
z

E
D

E
D

E
D

v

n
y
v
n1
y

u
y
v

y
u
y
v

z
u
z
v

n1

n2

E y
u

u D n1

z
v
n2
z
u
z
v

z
u
n2

v
(43)

Similarly the second and third terms become

F
y

(39)

~ r = (U, V, W ) and = (/x,


Representing U
/y, /z), we may restate the zero pressure jump

F x
= v
D n0
x
F u

D x
v

8
Copyright National Academy of Sciences. All rights reserved.

u + D

z
v
n2
z
u
z
v

u
n0

z
u
n2

v
(44)

Twenty-Fifth Symposium on Naval Hydrodynamics

G
z

G x
v
D n0
x
G u
D x

=
+

y
v
n1
y
u
y
v

G x
u

u D n0

y
u
n1

since N v = N + p and M v = M + q, and hence


we have to place more than one collocation points on
each panel, where the no flux kinematic boundary condition (10) on the blade surface should be satisfied. Based
on experience on applying the present method to the 2dimensional flow by Lee and Kerwin(2003), we will select four collocation points per panel at the intersections
of = 0.5 and 0.5 and = 0.5 and 0.5 when the
panel boundaries are limited by the normalized local parameter = 1 and 1 and = 1 and 1. The total number of collocation points will then be NCP =
4 (N M ). Note that, since NCP > NV, the simultaneous equations are now overdetermined and have
to be solved by a least square approach.
Substituting the relations (33) and (36), (35) and
(37), and the Morinos condition (38) into the integral
equation (13) and then applying this equation (13) to
NCP collocation points, we will obtain for the determination of NV potential vertices the following relation

(45)

Since the potentials in (43), (44) and (45) can be differentiated with respect to the parameters, the first three
terms in (42) may now be recast as

+F
+G
x
y
z

E F G
1 x y z
D v v v u
n0 n1 n2
x y z

1 u u u
G
E F
D
v
n0 n1 n2
x y z

u
u
1 u
y
x
z
D v v v n
G
E F

E
=

a,b

G
n

(46)

Sv = T

Substituting (46) into (42), we obtain the final form


of the dynamic Kutta condition equation as

X N
M v

E
M + F N
,

u
v
a,b

X N

E
M + F N
v,

u
v
a,b

= G

+ G

(48)

where A is NCP NV non-square matrix, v the unknown potential vertex strength vector and B the forcing
vector with NCP elements.
The dynamic Kutta condition (47) with all terms expressed in terms of the strength of the potential vertices
vi,j can be expressed in matrix form as

+ F
+ G
u
v
n
X N
X
M v
E
N
M v, + F

u
v ,
E

a,b

Av = B

(49)

where S is 2M NV non-square matrix and T 2M 1


matrix for the present problem. Since we want to impose
more weights on the dynamic Kutta condition equation at
the trailing edge than on the equation obtained from the
kinematic condition, the problem we have is to minimize
B Av subject to the constraint Sv = T.
Using Lagrange multipliers L, Lee and Kerwin(2003) showed that the final solution for the potential
vertices v is as

v = (AT A)1 (AT B ST L)

(47)

(50)

with the Lagrange multiplier L being

The significance of the above equation (47) is that the


velocity and hence the dynamic form of Kutta condition
can be expressed by a linear combination of the strengths
of the potential vertices vi,j . We will show later that the
dynamic boundary condition may be used as a constraint
equation when we determine the strength of the potential
vertices by a least square approach.

L = (S(AT A)1 ST )1 (S(AT A)1 AT B T)

(51)

NUMERICAL RESULTS AND DISCUSSIONS


In this section, we present sample computations to validate the present numerical procedure. We first select
a sphere positioned in uniform flow, for which analytic
solution is available, and then select a circular wing to
check the validity of the present in solving the lifting
problem. The final example will be the propeller operating in uniform oncoming flow with a constant rotational
speed.

LEAST SQUARE SOLUTION WITH A CONSTRAINT


The number of unknown potential vertices NV = (N v
M v ) is greater than the number of the panels (N M ),
9

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Flow around a sphere: For a sphere in a uniform


flow of infinite extent, the potential along the center
plane of symmetry is given by = 0.5 cos , where
is the azimuthal angle. Figure 5 shows the sphere represented by N M = 8 4, that is by 32 panels.
Note that each panel expressed by B-splines represents
the curved surface correctly. The potentials predicted by
the present method and by the analytic solution are compared in Figure 6. As expected the difference is indistinguishable. This result indicates that the numerical procedure functions correctly at least for a non-lifting blunt
body.

wing with tmax at root = 0.1


Circular
= 5o

Z
Y

Figure 7: A circular wing represented by B v, M


v ) = (13, 7)
splines with (p, q) = (3, 3), (N

and (N , M ) = (10, 4)

Y
X

10% thickness-chord ratio at the midspan, and is tapered


to zero quadratically as tmax (y) = to {1.0 (2y/s)2 },
where to is the maximum thickness at the root, s the
span, y the spanwise coordinate. Figure 7 shows the
circular wing represented with N M = 10 4.
Figure 8 shows the circulations distribution predicted by
the present method with (p, q) = (4, 3), (N , M ) =
(20, 10) and Jordans solution. Since Jordans result
is applicable only to the zero-thickness wing of circular planform, the circulation by the present method can
not be validated. The fact that circulation distribution is
lower than the zero-thickness wing explains the thickness
effect of the present circular wing. The skewed leading
edge of finite thickness is likely to reduce the loading
away from the mid-span compared to the zero-thickness
wing. The low order solution by Ramsey(1995) shows
however almost complete agreement with Jordans solution. It is not clear at present which numerical procedure
produces the correct result. Extensive numerical study,
between the present higher order method and Ramseys
low order method, is needed for validation of the present
procedure. Figure 9 shows the pressure distribution at selected spanwise stations, including a station very close to
the tip, that is at 0.49/0.5 = 98 % of half span. The pressure distribution, showing the typical flat plate loading
near the tip, is considered reasonable.
Propeller performance analysis: The propeller
4119 is known to have a simple symmetric blade contour and has abundant experimental data(Jessup 1989),
and has been tested for the verification of many numerical codes; hence DTRC 4119 propeller is a natural choice
to validate the new high order panel method. Figure 10
shows the blade and trailing wake surfaces represented
by B-splines. The principal particulars and offset of the
DTRC 4119 propeller can be found in Jessup(1989).
Convergence test: To test the convergence of the new
numerical method, numbers of chordwise and spanwise

Figure 5: A sphere represented by B-splines with


v, M
v ) = (11, 7) and
(p, q) = (3, 3), (N

(N , M ) = (8, 4)
0.5
0.4
0.3

Present Method
analytic

0.2

0.1
0

-0.1
-0.2
-0.3
-0.4
-0.5

90

180

270

360

Figure 6: Comparison of potentials predicted by


the present method (N , M = 8, 4) and the analytic solution

Flow around a circular wing: To validate the


present procedure, we select a circular wing positioned
in uniform oncoming flow at an angle of attack = 5o .
For a circular wing with zero thickness, an analytical solution for the circulation distribution in the spanwise direction is available. Our wing of the circular planform has
10

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

4
2.5

3.5

100.0*/2RU

100.0*/2RU

1.5

2.5

2
N M = 8 8
N M = 16 8
N M = 16 16
N M = 24 16
N M = 24 24

1.5

Present Method
Jordan
0.5

0.5

0
0

0.25

0.5

0.75

0.4

0.6

0.8

r/R

y/R

Figure 11: Convergence characteristics of radial


circulation distribution for various number of panels (J S = 0.833)

Figure 8: Comparison of circulations predicted


by the present method with N M = 20 10,
(p, q) = (4, 3) and Jordans solution
Circular Wing

panels, N and M respectively, are varied for DTRC


4119 propeller at Js = 0.833. Figure 11 shows the influence of the panel numbers on the circulation distribution. The figure shows in general good converged behavior with N = 24, M = 16 except a slight unstable behavior at the inner radii near the hub. This is
considered caused by the change of the relative distance
of the control points on the innermost strip to the hub
panels. Although the influence is local, further study is
necessary to improve the behavior in this region. Figures 12(a) through (d) show the pressure distribution at
selected radii for various panel numbers. The pressure
convergence is considered achieved with N = 16 and
M = 8.
Figure 13 shows the thrust and torque coefficients
at the same advance coefficient Js = 0.833 for various
panel number combination, with the converged loadings
with N = 24 and M = 16. Because the number of
control points on each panel is four, with the selection of
N = 24, M = 24, the total number of control points
where the boundary condition be satisfied will be 2304,
which is sufficiently large to minimize the least square
errors in determining 729 (27x27) unknowns.
Results and Discussion: Figure 14 shows the radial
distributions of the circulation computed by the present
method and the lifting surface theory of MIT(PSF-2), and
compares them with the experiments of Jessup. The experimental circulation of Jessup was derived through the
integration of the velocity measured by the laser Doppler
velocimeter just behind the trailing edge of the propeller
and is labed Measured in the figure. Added in the figure
is the result is corrected considering the viscous effect. It
is observed that the high order panel method correlates

0.49

NACA 0010
angle of attack 5 degree

-Cp

0.45

0.00

0.25

0.5

0.75

Figure 9: Pressure distribution at selected spanwise stations predicted by the present method with
N M = 20 10, (p, q) = (4, 3)
Y

X
Z

Figure 10: Geometry of propeller and wake


model generated by B-splines

11
Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

0.5
Pressure distribution on DTRC 4119, J=0.833

Pressure distribution on DTRC 4119, J=0.833


0.8

0.45

0.4

r/R=0.7

r/R=0.3

0.7

0.3

0.6
0.5

0.2

0.4

0.1

0.4
0.35

0.3
0

-Cp

-Cp

0.2
0.1

-0.1

0
88
16 8
16 16
24 16
24 24

-0.2
-0.3
-0.4

88
16 8
16 16
24 16
24 24

-0.3
-0.4

0.25
0.2

-0.5

-0.5
-0.6

0.3

10KQ

-0.2

-0.1

0.25

0.5

0.75

-0.6

0.25

s/C

0.5

0.75

s/C

KT

0.15
0.1

(a) pressure distribution


at 0.3R

(b) pressure distribution


at 0.7R

0.05
0

168

88

1616

2416

2424

NUMBER OF PANELS
Pressure distribution on DTRC 4119, J=0.833

Pressure distribution on DTRC 4119, J=0.833


0.4

0.4

r/R=0.98

r/R=0.9
0.3

0.2

0.2

0.1

0.1

-0.1

of

-0.1

-0.2

-0.2
88
16 8
16 16
24 16
24 24

-0.3
-0.4

88
16 8
16 16
24 16
24 24

-0.3
-0.4

-0.5
-0.6

Figure 13: Convergence characteristics


KT , 10 KQ for various number of panels

-Cp

-Cp

0.3

-0.5
0

0.25

0.5

0.75

-0.6

Measured
0

0.25

0.5

0.75

3.5

s/C

s/C

(c) pressure distribution


at 0.9R

(d) pressure distribution


at 0.98R

100.0*/2RU

Measured, Viscous corrected

Figure 12: Convergence characteristics of pressure distribution for various number of panels

2.5

1.5

Present Method
Experiment(Jessup)
Experiment(Jessup)

PSF-2

0.5

well with the corrected circulation over all radii.


Figures 15 through 18 show the pressure distribution in the chordwise direction at various radii r/R =
0.3, r/R = 0.7, r/R = 0.9. In these figures, the
present results are compared with experiments of Jessup
and the lifting surface analysis code, PSF-2, of Greeley and Kerwin(1982), and the low order panel code of
Hoshino(1989). First, due to the presence of the hub,
which is not included in the analysis, the lifting surface
prediction deviates far away from the experimental values, whereas the low and high order panel methods agree
each other. The difference between the panel method and
the experiments in Figure 15 is presumably due to the
viscous flow effect around the hub ahead of the propeller,
which is not present in the potential flow analysis.
In Figures 16 and 17, the agreement in pressure distribution is generally good between the experiment and
the theories. Most significant differences arise at the leading edge as shown in Figure 16, where the pressure varies
drastically from the stagnation pressure to some finite
value. In high order method, the curvature effect at/near
the leading edge is considered in the computation, but this
effect is included neither in the low order method nor in
the lifting surface method.
In Figure 17, the pressure on the lower surface of
the blade near the trailing edge at r/R = 0.9 is flat in

0.4

0.6

0.8

r/R

Figure 14: Circulation distribution of radial direction on DTRC 4119 propeller: Comparison
with experiments and PSF-2 prediction (J S =
0.833)
0.6

r/R=0.3

0.5
0.4
0.3

-Cp

0.2
0.1
0

-0.1

Present Method
-0.2

Experiment (Jessup)

-0.3

Hoshino (Panel)

-0.4

PSF-2

-0.5

0.25

0.5

0.75

s/C

Figure 15: Pressure distribution on DTRC 4119


propeller at r/R = 0.3: Comparison with experiments, PSF-2 and Hoshinos prediction (J S =
0.833)

12
Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

0.3
0.3

r/R=0.7

r/R=0.9

0.2
0.2
0.1
0.1

-Cp

-Cp

-0.1

-0.2

Present Method
Present Method

-0.1

Experiment (Jessup)

-0.3

Experiment (Jessup)

Hoshino (Panel)
-0.4

-0.5

0.25

0.5

Hoshino (Panel)

-0.2

PSF-2

PSF-2
0.75

s/C

0.25

0.5

0.75

s/C

Figure 16: Pressure distribution on DTRC 4119


propeller at r/R = 0.7: Comparison with experiments, PSF-2 and Hoshinos prediction (J S =
0.833)

Figure 17: Pressure distribution on DTRC 4119


propeller at r/R = 0.9: Comparison with experiments, PSF-2 and Hoshinos prediction (J S =
0.833)
0.3

experiment whereas is convex downward in the computations. To see whether this is due to any discrepancy of
the numerics, another Figure 18 is presented comparing
the predicted pressures by various numerical codes. It is
clearly seen that there exist correlations between the theories. This may imply first that there may be an error in
measurement errors or in data processing. Another possibility is related to the treatment of the Kutta condition
in numerical form.
Figures 19 shows the pressure distribution at r/R =
0.99, where the prediction by the low order panel method
has been practically impossible due to the abrupt change
of both the geometry and the flow characteristics. The
solution in this region may provide a critical information
for prediction of the tip vortex cavity flow solution. The
unnatural behavior of the pressure distribution at the trailing edge is likely to be influenced by the wake sheet location, which is fixed by a set of wake geometry parameters. This may be improved by introducing the real wake
model or by iterative procedure to locate the force-free
wake sheet.
Figure 20 show the velocity vectors and streamlines
on both suction and pressure sides of the blade. The
streamlines show a qualitatively similar picture to the experimental results of Jessup reproduced in Figure 21. It
is clear that the streamlines in the tip region direct toward
the hub as they move downstream on the suction side of
the blade, and outward on the pressure side. This crossflow component will form a strong vortex core when the
flow detaches from the blade tip.
Figure 22 shows the pressure contours on suction and
pressure sides of the blades. The pressure coefficient is
non-dimensionalized by the uniform inflow speed.
As a final step for DTRC propeller 4119, the open
water characteristics at various advance coefficients from

r/R=0.9
0.2

-Cp

0.1

-0.1
Present Method
Experiment(Jessup)
KRISO(Panel)
PSF-10(Panel)
VSAERO(Panel)

-0.2

0.25

0.5

0.75

s/C

Figure 18: Pressure distribution on DTRC 4119


propeller at r/R = 0.9: Comparison with experiments, prediction of other panel methods (J S =
0.833)
0.5

r/R=0.99
0.4

0.3

-Cp

0.2

0.1

Present Method

-0.1

-0.2

0.25

0.5

0.75

s/C

Figure 19: Pressure distribution on DTRC 4119


propeller at r/R = 0.99 (J S = 0.833)

13
Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Pressure side (or Face)


Z

0.98R
0.97R

0.9R

0.8R

Suction side (or Back)


X

0.98R
0.97R

(a) Pressure side

0.9R

0.8R

Figure 20: Velocity vectors and stream-lines on


DTRC 4119 propeller (J S = 0.833)
(b) Suction side

Js = 0.5 to Js = 1.1 were computed and compared with


experiments(Jessup 1989). In this calculation, the surface
drag coefficient of CF = 0.004 was used, and the number of panels were N = 16 and M = 8. Although
the correlation is fairly good all over the Js range, the experimental torque at Js = 0.5 is slightly larger than the
prediction. This may be due to the increase of drag due
to flow separation at the low Js value.

Figure 21: Velocity vectors and stream-lines on


DTRC 4119 propeller by Jessup(1989) (J S =
0.833)
be evaluated by Gaussian quadrature through desingularization. By adjusting the degree of the
quadrature the induction integrals are shown to be
able to reach any desired accuracy.

CONCLUSIONS
When the control point falls very close to the panel,
a continuous subdivision method can be applied to
evaluate the induction-integrals. Sub-division into
a smaller panels does not lose the accuracy of the
geometry and the potential representation.

A B-spline based high order panel method is developed for the analysis of the propeller performance.
Both the geometry and the velocity potential are represented by B-splines, thus increasing the accuracy
of geometry and flow representation.

The simultaneous equation for determination of the


potential is overdetermined and may be solved by
least square approach with the dynamic Kutta condition equation as constraints.

Depending on the relative position of the control


point to the singularity panel, the induction integrals
are divided into self-induction, near-field induction,
and far-field induction integrals, and different numerical procedures to evaluate these integrals are
shown accurate and effective.

Numerical experiments show that the new B-spline


based high order panel method is robust, can handle
the thin trailing edge and the tip region flow. Smaller
number of panels may be used for practical purpose
without sacrificing the accuracy.

The self-induction due to the high order normal dipole and source distribution is shown to
14

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Pressure side (or Face)


Z
411 3
7 510 91112
1311
15

12
14 14

16

161

16

1 8 15
13

14

Level
25
24
23
22
21
20
19
18
17
16
15
14
13
12
11
10
9
8
7
6
5
4
3
2
1

17

17 1 8

17

15

16

23

25

17

17 17

1923
18 17

16

18

16

18

19

16

18

22 14
2

16

14

18

221

12

17
1818
20

23

17

11

10

15

14

21
2223

17

22

15
19

17

15

19

15 15

20

14

13

13

17

1616

192020

2521

18

15

14

151 111315
16 3

Cp
0.3000
0.2667
0.2333
0.2000
0.1667
0.1333
0.1000
0.0667
0.0333
0.0000
-0.0333
-0.0667
-0.1000
-0.1333
-0.1667
-0.2000
-0.2333
-0.2667
-0.3000
-0.3333
-0.3667
-0.4000
-0.4333
-0.4667
-0.5000

X
1
611089 6 9

99
8 79612
14
14
15

51 72
2
13 17
24 21
15 252
02

12
11

17

21

17
18 18

12

12

11

11

13
14
15

1111

11

111347

1
19

25
5 6 01 9
4
01 2
11 4224
2
23

Extensive numerical tests are needed for validation


of the new method for propellers of various geometry.

14
15

ACKNOWLEDGMENTS

Cp
0.3000
0.2667
0.2333
0.2000
0.1667
0.1333
0.1000
0.0667
0.0333
0.0000
-0.0333
-0.0667
-0.1000
-0.1333
-0.1667
-0.2000
-0.2333
-0.2667
-0.3000
-0.3333
-0.3667
-0.4000
-0.4333
-0.4667
-0.5000

This research was initially supported from 1 June 1998


- 31 August 1998 by the Massachusetts Institute of
Technology under Contract N0016798M1038 from the
Carderock Division, Naval Surface Warfare Center of the
United States Navy, Dr. Ki-Han Kim, Project Monitor,
and subsequently supported by Advanced Ship Engineering Research Center administered by the Korea Science
and Engineering Foundation, Grant No. R11-2002-10402001-0. Financial supports are greatly appreciated.

11

25

8
14 9
18

17

1 91
23 9

23

10

23

12

18

88

20

11

10

6
4

18

14

19

13
25
14 1919 18

17

10

15

11

10

15
1724

Level
25
24
23
22
21
20
19
18
17
16
15
14
13
12
11
10
9
8
7
6
5
4
3
2
1

13

13

23
17
14

13

14 14

13

14

513
11 1010
13

To improve the accuracy of the new high order panel


method, the trailing wake surface should be placed
to be force-free, either through the improved wake
modeling or through the iterative search of the true
surface location satisfying the kinematic and dynamic boundary conditions on the wake sheet.

Suction side (or Back)

11
13

If the blades are represented by NURBS(NonUniform Rational B-Spline) surfaces, the panel
method can be easily extended to adopt the NURBS
surface.

Figure 22: Non-dimensionalized pressure distribution on DTRC 4119 propeller (J S = 0.833)

REFERENCES
Greeley, D.S. and Kerwin, J.E., Numerical Methods
for Propeller Design and Analysis in Steady Flow,
Transactions of SNAME, Vol. 90, 1982, pp. 415-453.
Hess, J.L., Review of Integral-Equation Techniques
for Solving Potential-Flow Problems with Emphasis on
the Surface-Source Method, Computational Methods in
Applied Mechanics and Engineering, Vol. 5, 1975, pp.
145-196.

1.2

DTRC 4119
1

Experiment
Present Method

0.8

Hess, J.L. and Smith, A.M.O., Calculation of Nonlifting


Potential Flow About Arbitrary Three-Dimensional
Bodies, Journal of Ship Research, Vol. 8, No. 2, 1964,
pp. 22-44.

o 0.6
10KQ
0.4
KT
0.2

0
0.4

0.6

0.8
J

Hess, J.L. and Valarezo, W.O., Calculation of Steady


Flow about Propellers by Means of a Surface Panel
Method, 23-rd Aerospace Science Meeting, AIAA,
January 14-17, 1985, Reno, Nevada.

1.2

Figure 23: Comparison of open water performance of DTRC 4119 propeller with experiments

Hoshino, T., Hydrodynamic Analysis of Propellers


in Steady Flow using a Surface Panel Method, Journal
of the Society of Naval Architects of Japan, Vol. 165,
1989, pp. 55-70.
15

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

in (12) and (5), we may use the chain rule to get


Hsin, C.-Y., Kerwin, J.E. and Newman, J.N., HIPAN2:
A Two-Dimensional Higher-Order Panel Method Based
on B-Splines, theory and program documentation, 1993,
Department of Ocean Engineering, M.I.T.

Jessup, S. D., An experimental investigation of


viscous aspects of propeller blade flow, Ph.D. Thesis,
The Catholic Univ. of America, 1989.

=
=
=

x
x u
x
x v

x n0

+
+
+

y
y u
y
y v

y n1

z
z u
z
z v

z n2

+
+
+

(52)

where ~x = (x, y, z) and n


= (n0 , n1 , n2 ). We can
compute the derivatives of ~x with respect to the parameters (u, v) using (17) and the derivatives of with respect
to the parameters similarly. /n on the last line of (52)
is known from the kinematic boundary condition. Using
Kramers rule, we can solve (52) for the perturbation velocity components as
y z

u
u
u

1
y
z
=
v
v
v
x
D

n
n
1
2
n
x z

u
u

1 u

x
z
=
v
v
y
D v
n0 n n2
x y

u
u
1 u

y

x
=
(53)
v
z
D v v
n0 n1 n

Kerwin, J.E., Kinnas, S.A, Lee, J.-T. and Shih, W.-Z., A


Surface Panel Method for the Hydrodynamic Analysis of
Ducted Propellers, Transactions of SNAME, Vol. 95,
1987, pp. 93-122.
Lee, J.-T., A Potential Based Panel Method for the
Analysis of Marine Propellers in Steady Flow, PhD
Thesis, Department of Ocean Engineering, M.I.T., 1987.
Lee, C.-S and Kerwin, J.E., A B-Spline Higher
Order Panel Method Applied to Two-Dimensional
Lifting Problem, Journal of Ship Research, Vol. 47, No.
4, 2003, pp. 290-298.
Maniar, H.D., A three dimensional higher order
panel method based on B-splines, PhD Thesis, Department of Ocean Engineering, M.I.T., 1995.

where the determinant D is defined as


x y z

u
u u
y
z
D = x
v
v
v
n0 n1 n2

Morino, L. and Kuo, C.-C., Subsonic potential


aerodynamic for complex configuration: A general
theory, AIAA Journal, Vol. 12, No. 2, 1974, pp.
191-197.
Piegl, L. and Tiller, W., The NURBS Book, 2nd
Ed, Springer-Verlag, Berlin, Heidelberg, 1996.
Ramsey, W. D., Boundary Integral Methods for
Lifting Bodies with Vortex Wakes, PhD Thesis, Department of Ocean Engineering, M.I.T., 1995.
Rogers, D.F. and Adams, J.A., Mathematical Elements for Computer Graphics, 2nd Ed, McGraw-Hill,
New York, 1989.

APPENDIX A. EVALUATION OF PERTURBATION


VELOCITY
To compute the perturbation velocity on the body surface,
we need to take derivative of the perturbation potential
with respect to the physical coordinates. Since the perturbation potential and the physical coordinate ~x are
both represented as functions of the parameters (u, v) as
16
Copyright National Academy of Sciences. All rights reserved.

(54)

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION

DISCUSSION

Ki-Han Kim
Office of Naval Research, USA

Prof. Spyros A. Kinnas


University of Texas at Austin, USA

I would like to congratulate the authors for


their successful efforts in extending the potential
flow theory to predict the tip flow that is known to be
one of the most difficult areas.
There are no propeller analysis methods
based on the potential flow theory (lifting surface or
low-order panel methods) that are currently available
that can predict the tip flow accurately and reliably.
Several researchers (for example, Pyo and Kinnas
(1997)) in the past tried to accurately model the tip
flow using the potential flow theory with moderate
success.
In practice, therefore, we compute the flow
quantities, including the velocity, pressure and
circulation distribution, up to a certain limit (for
example 0.9 radius) and extrapolate these quantities
to the tip. The authors presented a new B-Spline
based higher-order panel method with a systematic
validation that showed a good convergence and
physically correct flow behavior close to the tip area
(0.99 radius) for relatively simple test cases.
I have a couple of suggestions for the
authors to consider pursuing in the future:

I would like to comment the authors on


presenting a rigorous BEM technique for the
prediction of the inviscid flow around wings and
propeller blades. I have the following comments:
a)

In comparing their results with those of the


lifting surface theory (Jordan) for a circular
planform wing of zero thickness, they present
their BEM results for a thickness to chord ratio
of 10%. In order to make their comparison with
Jordans theory more relevant, given the fact that
the BEM formulation becomes degenerate for
zero thickness, they should present a comparison
of Jordans circulation distribution with that
which is extrapolated to zero thickness from a
series of runs from their method at various
thickness to chord ratios (e.g. 2%, 5%, 10%)

b) In the paper of Kinnas et al (1993) it was found


that applying an iterative pressure Kutta
condition (Kinnas and Hsin, 1992) at the trailing
edge of the blades could produce non-physical
results in some cases (like in the case of a
circular planform wing) if the trailing wake
surface was NOT aligned (like in the presented
method) with the resulting flow. In fact at the
same paper it was proven (by theory and by
results) that aligning the trailing wake made the
application of the iterative pressure Kutta
condition less necessary (i.e. a Morino type of
condition was sufficient).

(1) Since NURBS (Non-Uniform Rational B-Spline)


representation of complex surfaces is widely
used, as the authors correctly pointed out, extend
the current non-rational (integral) B-spline
method to NURBS surfaces,
(2) Continue validation for progressively more
complex geometry and flow, including highlyskewed propellers, and
(3) Extend the steady method to unsteady method
including cavitation

c)

Although viscous flow methods are


increasingly becoming popular in the propeller
community, accurate and reliable potential flow
methods are always welcome considering the
computational cost, hardware and trained manpower
availability.

In their comparison with PSF-2 (Fig. 14 in the


paper) it should be noted that PSF-2 has been
run without a hub (as it appears to be from the
behavior of the shown circulation distribution at
the hub) and that the thickness/loading coupling
(Kinnas, 1992) has not been included.

REFERENCES
Kinnas, S.A., Pyo, S., Hsin, C.-Y., and Kerwin, J.E.,
"Numerical Modeling of Propeller Tip Flows," Sixth
International Conference on Numerical Ship
Hydrodynamics, pp. 531-544, Iowa City, IA, August
1993.

REFERENCES
Pyo, S. and Kinnas, S.A., "Propeller wake sheet rollup modeling in three dimensions," Journal of Ship
Research, Vol. 41, pp. 81-92, 1997.

Kinnas, S.A. and Hsin, C.-Y., "A Boundary Element


Method for the Analysis of the Unsteady Flow

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Around Extreme Propeller Geometries,"


Journal, Vol. 30, pp. 688-696, 1992.

AIAA

Kinnas, S.A., "A General Theory for the Coupling


Between Thickness and Loading for Wings and
Propellers," Journal of Ship Research, Vol. 36, pp.
59-68, 1992.
AUTHORS REPLY
The authors first of all express sincere
thanks to all discussions and comments.
Dr Kim suggested the extension of the
present work in various aspects. As he pointed out
NURBS is the industry-standard in representing the
complex propeller geometry. The present formulation
using (integral) B-spline surface can be easily
extended to NURBS-represented geometry. The
potential however may better be expressed by the
integral B-spline, because the weight factor
introduced in NURBS expression can be another
source of uncertainty. Once we build up experience
with the integral B-spline expressed potential,
NURBS may provide another flexibility in
expressing flow quantities.
Dr Jessup asked our plan to improve the
wake model to be force free. Because we have
increased the accuracy of the solution on the blade
surface, notably near the trailing edge and blade tip
regions, B-spline represented wake for geometry and
potential will be our next consideration. This will be
another challenge, but we exptect the B-spline
represented dipole wake can be handled relatively
easily than the existing alignment method. The issue
related to Kutta condition raised by Prof Kinnas may
be studied at the same time.
The numerical procedure in the present
formulation has been systematically tested for series
of propellers. The correlation is considered fairly
good for the normal blade geometry, but not
validated with propellers of extreme geometry. Once
the wake alignment procedure is completed, we will
conduct comparative studies with the highly skewed
propellers, as suggested by Dr Kim.
Prof Kinnas suggested the computation with
several thickness ratios to compare with Jordans
solution to the lifting surface problem for the circular
wing of zero-thickness. We are planning to perform
that computation in near future.
The present method can be extended to
solve the unsteady problem relatively easily, but will
be very difficult to be applied to the cavity problem,
if not at all impossible.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics



Symposium on Naval Hydrodynamics


St. Johns, Newfoundland and Labrador, CANADA, 8-13 August 2004

Prediction of Performance of Ducted and Podded Propellers


Spyros A. Kinnas, Hanseong Lee, Hua Gu, Apurva Gupta
(The University of Texas at Austin, USA)

ABSTRACT
The numerical modeling of the flow around two types of
high efficiency propulsors, ducted and podded propellers,
is performed via coupling of a Vortex Lattice Method
(VLM) based potential solver (MPUF3A) with a Finite
Volume Method (FVM) based Euler solver (GBFLOW).
The coupling of GBFLOW with MPUF3A provides a
way of analyzing the interactions among propeller, pod
and strut, or between propeller and duct. In the coupling
of MPUF3A and GBFLOW, MPUF3A is first solved for
the potential flow distribution around the propeller, and
the pressure distributions on the propeller blades are evaluated. The pressure distributions are then converted into
body force terms which represent the existence of the
propeller in GBFLOW. GBFLOW solves the Euler equations, with the converted body force terms and the appropriate boundary conditions, in the whole fluid domain
which includes the appendages such as duct, pod and
strut. The effective wake to the propeller and the body
forces keep being updated as iterations continue until the
loading of the propeller converges. The numerical computations are carried out for a ducted propeller, and a pull
or twin types of podded propulsors, and the method is
validated by comparing the numerical results with those
measured in experiments.
INTRODUCTION
The multi-component propulsors like ducted and podded
propellers have been widely used in modern commercial
ships on account of their high efficiency, maneuverability
and sea-keeping ability. A ducted propeller, with the influence of duct, can either accelerate flow inside the duct
and provide additional thrust (accelerating type) or improve the pressure distribution inside the duct and delay
the occurrence of cavitation on the blade (decelerating
type). A podded propulsors, with the proper design of
the pod and strut, can also improve the overall performance of the propulsion and reduce the fuel consumption
of a vessel. In addition, podded propulsors and ducted
propellers can be used in dynamic positioning systems of

floating structures by being mounted on the bottom of the


hull, and by being able to rotate around a vertical axis,
and provide thrust in all horizontal directions. In accordance with the wide application of these two advanced
propulsor systems, the accurate prediction of their performance is in high demand.
Ducted propeller
Most of numerical methods for the modeling of ducted
propellers have been developed based on the potential
methods such as lifting line method, vortex lattice method
and boundary element method.
The actuator disk model of a propeller was coupled with
surface vorticity method to predict the performance of
a ducted propeller by Gibson & Lewis (1973). In their
method, the singularities were distributed on the actual
duct surfaces to represent the duct more exactly and
thus to introduce the nonlinear thickness effect in the
calculation of hydrodynamic characteristics. Glover &
Ryan (1973) represented the propeller by lifting line and
the duct by distributing surface vorticity on the actual
duct surface. Later, Falcao de Campos (1983) had coupled the actuator disk model with an Euler solver to account for the axisymmetric shear flow effects of a radially varying inflow field. The research on ducted propeller has advanced by developing of lifting surface theory which could account for the complex blade geometries more accurately, and this method was applied in
steady (Van Houten 1986), and in unsteady (Feng &
Dong 1985) flows.
A potential based panel method, which could account
for non-axisymmetric interaction between propeller and
duct, was developed by Kerwin et al. (1987). Later, their
method was extended to determine the optimum loading
distribution on blades inside of a duct which was represented in nonlinear theory by Kinnas & Coney (1992).
Hughes et al. (1992) compared the experimental results
with those from a coupled propeller lifting surface and
duct/hub surface panel analysis code over a wide range
of advance coefficients. Furthermore, Hughes & Kinnas
(1993) and Hughes (1993) modeled a ducted propulsor

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

in unsteady potential flow using a time marching panel


method for all components. In their method, the duct and
hub were represented using the generalized image model,
which accounted for the effects of the duct and hub on
the propeller blades without solving the boundary value
problem for the duct and hub at each time step.
Recently, viscous flow methods have been applied to
properly analyze the gap flow at the tip of the propeller
inside of a duct. Kerwin et al. (1994, 1997) and Warren
et al. (2000) developed a hybrid method which coupled a
RANS method for the duct with a lifting surface method
for the propeller. Incompressible viscous flow methods
which model both the duct and the propeller using the
RANS equations have been developed by Sanchez-Caja
et al. (2000) and Abdel-Maksoud & Heinke (2002).
Podded propulsor
A podded propulsor can be classified into a push type
which propeller operates downstream of the strut, a pull
type which propeller operates upstream of the strut, or
a twin type in which one propeller operates upstream of
the strut and the other downstream of the strut. Podded
propulsors which involve contra-rotating propellers have
also been designed. More recent forms of podded propulsors may be found in Atlar (2004). According to the report of HYDROCOMP Inc. (1999), the podded propulsors usually have a large hub and often have been designed to mitigate noise and vibrations by off-loading the
tip and root areas. Although the off-loading can decrease
propeller efficiency by a few percent, however, the overall efficiency gain of a podded propulsor system over a
conventional propeller can be increased in the order of
2% to 4%. This observation leads to the need for the development of numerical tools which can reliably predict
the overall performance of a podded propulsor system.
Computational modeling of podded propulsors involves
adapting computational grids around complex geometries. Ghassemi & Allievi (1999) developed potential
flow methods, and Sanchez-Caja et al. (1999) applied viscous flow methods to calculate the flow around podded
propulsors. More recently Hsin et al. (2002) developed
a design tool for pod geometries based on a coupled viscous and potential flow method. A coupled Euler solver
with a potential flow method was applied for the performance analysis of podded propulsors by Kakar (2002)
and Kinnas et al. (2004).
Solution method
In this work, a general method of coupling of a Vortex Lattice Method (VLM) with a Finite Volume Method
(FVM) is applied to model the flow around podded and

ducted propellers. For both cases, MPUF3A, a Vortex Lattice Method based potential solver, solves the
flow in the vicinity of the propeller, and GBFLOW, a
Finite Volume Method based Euler solver, solves for
fluid domain where the appendages (hub, pod/strut or
duct) are included. The interactions among propeller(s),
pod and strut (or duct), and the inflow (which can be
non-axisymmetric in general) are considered by applying
MPUF3A and GBFLOW with the effects of one on the
other being considered iteratively.
First, MPUF3A solves the potential flow problem around
the propeller and hub subject to the effective wake, and
calculates the pressure, forces and moments on each propeller blade. It should be noted that MPUF-3A can also
evaluate unsteady (leading edge or mid-chord) sheet cavitation on the blades. The initial effective inflow to the
propeller is taken as an uniform (or a given nominal nonaxisymmetric) wake, and the inflow wake is updated by
using the resulting velocities from GBFLOW as iterations
continue. The calculated pressure distribution is then
converted to body force terms which represent the existence of each propeller in GBFLOW. The Euler equations
are then solved with the appropriate boundary conditions.
Once the Euler equations are solved, the effective wake is
determined by subtracting the propeller induced velocity
computed using the result of MPUF3A, from the total velocity computed in GBFLOW near the propeller plane.
The iteration process between MPUF3A and GBFLOW
continues until the loading of the propeller converges.
The present method is applied to ducted propellers and
podded propulsors (a pull and a twin type), and the predicted forces are compared to those measured in experiments.

GOVERNING EQUATIONS FOR EULER SOLVER


The flow field around a duct or pod and strut is solved
by using a 3-D steady Euler solver (GBFLOW) based on
a finite volume method and the artificial compressibility
method (Chorin 1967). The details of this method are
described in Choi & Kinnas (2001) for steady flow, and
in Choi & Kinnas (2000, 2003) for unsteady flow.
The non-dimensional continuity and the momentum
equations in a ship fixed coordinates system are as follows:





 
  

(1)

where the column matrices, U, F, G, H, and Q, are de-

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

fined as follows.

 

 









 


 


 


r


 


duct











 





 







 



gap




R prop

mid chord line



blade




(2)

hub




where is the artificial compressibility factor.


A vertex based scheme finite volume method (Choi &
Kinnas 2001) is used to solve the governing equation 1.
In addition, Nis lax-Wendroff method (Ni 1982) is introduced for the time discretization and a fourth order artificial viscosity is applied to help the stability of
the solution. The body force terms in the column matrix  of equation 2 are calculated by integrating the
pressure difference between the face and the back sides
of the blade surface,   , over the area of the lifting
surface which is intersected by the finite volume cells.
The non-dimensionalized 3-D body force terms applied
in GBFLOW are evaluated as follows (Choi & Kinnas
2001).


!
"$#&%


')(


*,+

Figure 1: A configuration of ducted propeller and geometric parameters. From Kinnas et al. (2002).

MPUF3A GBFLOW

duct
body
force

effective
wake
plane

blade
pressure
effective
wake

Figure 2: Iterative solution method for ducted propeller


problem.

(3)
Far Field Boundary


*,+

where
is the non-dimensionalized pressure force
calculated from the propeller potential flow solver
",#
%
(MPUF3A). , and ' are the volume cell and the advance ratio based on ship speed, respectively.
Inflow Boundary

Outflow Boundary

DUCTED PROPELLER
A typical configuration of a ducted propeller is shown
together with the main geometric parameters in Fig. 1. In
the figure, the duct sectional angle and the inflow angle
of attack are denoted as - and -/. , respectively, and the
propeller radius, 021&.4351 , is defined as a distance from hub
centerline to midchord of the blade tip. The gap size is
expressed as a percentage of the propeller radius which
is measured at the midchord of blade, as shown in the
figure. Notice that the propeller radius at the tip varies
along the chord since the blade tip is placed at a constant
distance from the duct inner surface. The technique of
adapting the blade geometry to the duct and hub surface
is described in Kinnas et al. (2002).

Solid Boundary
Repeat Boundary

Axis Boundary

Figure 3: Flow domain around the duct, and boundary


conditions in the Euler solver.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Interaction between duct and propeller

Far field boundary

In the present method, a vortex lattice method based


MPUF3A and a finite volume method based GBFLOW
are coupled to solve the fluid flow around ducted propellers. A MPUF3A solves the performance of propeller
with the effect of hub and duct which are modeled using a simplified image method. An uniform inflow (or
given nominal non-axisymmetric inflow) is provided as
a guessed effective wake at the first iteration step, and
MPUF3A predicts the propeller forces based on the provided inflow. The pressure distributions computed from
MPUF3A on the propeller blades are then converted to
the body force terms which represent the existence of
propeller in GBFLOW. The duct and hub are modeled
as parts of boundary conditions in GBFLOW. Once the
Euler equations are solved, the total velocity and pressure distributions inside the entire fluid domain are determined. Then, the effective wake can be generated at
the propeller plane by subtracting the propeller induced
velocities from the total velocities. At the next iteration
steps, the updated effective wake is used by MPUF3A
to update the force distributions on the propeller blades.
This iterative procedure, as shown in Fig. 2, continues
until the forces and moments of the propeller blades converge within a specified tolerance.
Grid and boundary conditions in GBFLOW
In the present work, only axisymmetric duct and hub geometries are considered. Therefore,

a 2-D grid is generated first, and then rotated along -axis to create the 3D
grid. Since the duct is surrounded by fluid in 2-D grid,
this domain is a doubly-connected. A C-Type grid with
a cut at the trailing edge of the duct, as shown in Fig. 2,
makes the domain simply-connected and thus simplifies
the numerical implementation. The boundary conditions
in GBFLOW are specified on each boundary of the domain, as shown in Fig. 3. The velocities and pressures
along the cut (repeat wake boundary) are averaged from
those at neighboring vertices, and the actual duct trailing wake does not follow this cut. The applied boundary
conditions on each boundary are as follows:
Inflow boundary










Outflow boundary

 








(4)









 


Axis boundary







    
   



 

 







(6)

(7)

where  
: cell nodes next to the axis, and

  is the number
of cells in the circumferential direction.
Solid boundary

!" #


where ! 








 .

#


(8)

Repeat boundary: the values of the velocities and


pressure are averaged from those at the adjacent
nodes in the radial direction.
Image method
Kerwin et al. (1987) pointed out that, when the tip gap
is small, the flow around the duct varies between blades.
They named such effects as duct wall effect, or hub wall
effect in the case of the hub. However, different from the
panel method applied in Kerwin et al. (1987), the present
method represents the propeller blades as the body forces,
and the flow variation around the duct in the circumferential direction would be inaccurate. In order to include
the duct or hub wall effect, a simplified image model with
respect to inner duct surface (or hub surface) is applied in
MPUF3A. An assessment of the accuracy of this simplified image model, as compared with the exact image
model, is provided in Kinnas & Coney (1992).
The locations of image vertices are determined as follows
(Kinnas et al. 2002):


)( $  $

'
( $

( $&$&%%
$ % and ( $ %
0

(9)
(10)

are the radial( and axial coordinates


where 0
of the image respectively. 0 $ and $ are the coordinates
of blade vertices, and( 0 ' denotes the radius of the inner
surface of the duct at $ .

Gap model


(5)

The predicted propeller forces and cavity patterns via an


inviscid flow model (e.g. MPUF-3A) seem to compare

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

ULTIMATE
WAKE

4
DUCT
INNER
SURFACE

GAP
VORTEX
PANELS

3
100

BLADE
VORTEX
PANELS

2
Y

20x 9
20x18
20x27
20x18, w/o gap model

Z
X

WAKE
VORTEX
PANELS

0
0.2

0.4

GAP
WAKE
PANELS

Figure 4: Vortex lattice model for ducted propeller with


blade, wake and gap vortex loops.
images

0.6

0.8

r/R

Figure 7: Convergence of predicted circulation distri%




butions
with number of spanwise panels:  
,




, and gap size = 0.38 %
5

gap vortices
tip

gap wake vortices

3
100



inner
surface


duct

2
0% gap
0.38% gap
1% gap
2% gap

blade
L.E.

wake

0
0.2

0.4

0.6

0.8

r/R

T.E.

Figure 5: Vortex loop structure in the gap model.


KB
GAP KBC
GAP DBC

Figure 8: Effect of gap percentage on predicted circu%



 
lation
distributions by present method:  
, and
 

magnified view

100

CQ=0.0
CQ=0.3
CQ=0.6
CQ=0.85
CQ=1.0
CQ=1.5

0
0.2

Figure 6: Arrangement of control points on the blade and


gap panels.

0.4

0.6

r/R

0.8

 

Figure 9: Effect of discharge coefficient,


, on pre%
dicted circulation distributions by present method:  

 
, and gap size = 1%

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

well with those of the viscous model at high Reynolds


number, over a wide range of advance ratios, as reported
in Kinnas et al. (2002). In the case of ducted propellers,
however, the viscous flow in the gap between the blade
and the duct can affect the overall performance significantly. From flow visualizations, measurements in experiments (Chesnakas & Jessup 2003), and detailed viscous
flow calculations (Brewer et al. 2003), a strong tip vortex
would develop from the leading edge part of a square tip
blade, similar to that off the leading edge of a delta wing,
which will interact with the vortex which is coming off
the trailing edge of the tip. This flow pattern acts like
a winglet and increases the loading of the blade at the
tip. There would be no finite loading at the tip for this
case, since MPUF3A produces zero loading at the tip of
the blade unless an image model is used and zero gap is
utilized. To capture the details of the tip gap flow, a viscous flow solver with high concentration of grids at the
tip region is required (Brewer et al. 2003). In the present
method the detailed viscous flow in the gap region is not
addressed. However, the effect of the gap flow on the
loading and performance of ducted propeller is accounted
for by using the orifice equation model as proposed in
Kerwin et al. (1987). This model has been successfully
coupled with panel methods by Hughes (1997) and Moon
et al. (2002).
By using Bernoullis obstruction theory (White 1986), the
flow through an orifice can be approximated as follows:
 

(11)


 

is the empirical discharge coefficient, is the


where

total flow volume per unit time across the
 gap, and 
is the pressure difference across the gap. and are the
density of the fluid and the clearance height of the gap,
respectively. With normalization, the relative flow, ! . ,
across the gap becomes

!
.








(12)
1

where

!
.



and


 


1

  


(13)

The discharge coefficient


is taken equal to 0.85 in
the present method, while the suggested value in Hughes
(1997) and Moon et al. (2002) is 0.84.
Vortex loops are built inside the gap to implement the gap
model in MPUF3A, and extended from the tip panels of
the blade to the duct inner surface. According to Kelvins
theorem, these vortex loops cannot end in the fluid domain and must shed vorticity in the same way the trailing wake of the propeller blade does. The vortex loop

arrangement on the blade, its image and within the gap


is depicted in Fig. 4, and a 3-D view of the actual blade
and gap vortex loops with the control points in the gap is
shown in Fig. 5.
Equation 12 determines the flow velocities at the gap
panels, and thus is applied on the gap kinematic boundary condition (KBC) points. Since the pressures on the
blade are evaluated at blade dynamic boundary condition (DBC) points in MPUF3A, the gap DBC points are
also introduced on the gap panels. The arrangement of
KBC and DBC points on gap panels is similar to that on
the blade panels, as shown in Fig. 6. Unlike the DBC
points on the blade, where the dynamic boundary condition (unsteady Bernoulli equation) is applied to evaluate
the pressures on the blade, in the case of the gap panels, the pressures at the DBC points are extrapolated from
the evaluated pressures on the blade. Then, the pressures
at the gap KBC points are determined via interpolation

of the pressures at the gap DBC points, so that 
1
of
equation 12 is determined. The original MPUF3A solves
for the strengths of the blade vortices by requiring zero
normal velocity at the blade KBC points. In the case
of the gap panels the strengths of gap vortex loops are
determined by applying the conditions required by equation 12. An iterative process is required to apply
equation

12 at the gap KBC points, since the values of 
1 depend
on the propeller loading, as well as the gap loading.
A Ka 4-70 Wageningen B-series propeller with NSMB
19A duct is considered to validate the current model. The
blades, and a pitch to propeller diamepropeller has four
  . The four bladed ducted propeller

ter ratio of
%


is subject to uniform inflow at advance ratio of ' 
and the gap of 0.38% of the propeller radius. The dependence of the circulation distribution with respect to the
number of panels in the spanwise direction is shown in
Fig. 7. The panel numbers in the chordwise direction are
kept same as twenty, while varying in the spanwise direction from nine, to eighteen and twenty seven. The circulation distribution predicted by the original MPUF3A
without the gap model is also plotted in the same graph.
Figure 7 shows clearly that the present gap model increases the loading of the blade, especially at the blade
tip. The effect of the gap size on the circulation distribution is shown in Fig. 8. As expected, the loading of the
propeller blades decreases as the gap size increases. Notice that the circulation distribution for gap of 0.38% is
very close to that for zero gap. This indicates that, if the
gap is small enough, the gap is practically sealed and the
ducted propeller acts like that with zero gap. Finally, the
effect of the discharge coefficient on the predicted circulation distribution is studied in the case of a gap  of 1%,
and the results are depicted in Fig. 9. Since a zero
implies that there is no flow across the gap and thus the blade

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

tip is attached to the duct inner surface, the predicted circulation distributions have to be the highest
and same as
 
that of zero gap, as shown in Fig. 8. As
increases, the
effect of the viscous gap flow becomes smaller (in other
words the gap flow is inviscid and that can lead to very
high velocities through the gap), and the predicted loading at the blade tip decreases. Then, the propeller loading
converges to that of potential solver without gap model.
 
In the present work, however, a moderate value of
is chosen from Hughes (1997) and Moon et al. (2002) to
determine the performance of ducted propellers.

GBFLOW-3D --- Solution on the Center Plane


1.0

U
1.50
1.34

1.18
1.02
0.86
0.70
0.54
0.38
0.22
0.06
-0.10

Validations
Figure 10: Grid and axial velocity contours around duct
3
predicted by GBFLOW: NACA0015 duct with - 
4

Panel Method (MIT)


GBFLOW-3X
GBFLOW-3D

2
-Cp

In order to validate the present method, the numerical calculation is first performed for the bare duct without propeller. Since no propeller is present in this case, the computation is performed only by GBFLOW.
The first case is for an axisymmetric duct for which Kerwin et al. (1987) has computed the pressure distributions on the duct surface using the surface panel method.
The duct section has a NACA0015 thickness distribution
without camber, and the duct radius at the leading edge is
equal to the duct chord. The angle of attack of the duct
  3 . Figure 10 shows the 2-D grid and
section is - 
the predicted axial velocity distributions around the duct.
The comparison of the pressure distributions on the duct
section between the predicted from the present method
and from the panel method is shown in Fig. 11. In the
figure, GBFLOW-3X and GBFLOW-3D stand for the axisymmetric and the fully 3-D version of GBFLOW, respectively. The predicted results from both Euler solvers
compare quite with those from the panel method.
The second case is for DUCT II subject to open flow at
an angle of attack. The experiments for the pressure measurement on DUCT II surface were conducted by Morgan
& Caster (1965). The section of the duct has a modified
NACA 66 thickness distribution with maximum thickness
to chord ratio of 0.10, and a NACA 0.8 meanline with
maximum camber to chord ratio of 0.04. The sectional
 3
angle of attack is - 
, and the ratio of the duct chord
to the duct inner surface diameter at the propeller plane is
0.8. Figure 12 shows the pressure contours and streamlines around DUCT II when the inflow angle of attack is
3
. Due to the inflow angle of attack, the pressure distribution is no longer axisymmetric. The predicted pressures at the top and the bottom sections of the duct are
close to those measured by Morgan & Caster (1965), as
shown in Fig 13.
Oosterveld (1970) conducted systematic series tests on
ducted propellers. Ducted propeller with Ka 4-70 propeller and nozzle (or duct) 19A is modeled with the
present method. A Ka 4-70 propeller has four blades of

-1

-2

0.2

0.4

x/C

0.6

0.8

Figure 11: Comparison of pressure distributions predicted by panel method (Kerwin et al. 1987) and present
3
method: NACA0015 duct with - 
Y

GBFLOW-3D --- Solution on the Center Plane


1.0

P
0.51
0.42
0.32
0.22
0.12
0.03
-0.07
-0.17
-0.27
-0.36
-0.46

Figure 12: Pressure contours and streamlines predicted


by GBFLOW: DUCT II at inflow angle of attack of -/. 
3

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

=0






























     
  
 
 




























































 




























     

 
 
   


  
   
GBFLOW (inner surface)
  
   
GBFLOW
(outer
surface)
  
EXPERIMENT (inner surface) 
   
  
EXPERIMENT
(outer
surface)

 
 





























   



 


 

























































































   
 


 


 
























































































   

-1.2
-1
-0.8
-0.6
-0.4

C
Pp

   -0.2


  0
0.2
0.4

0.6
0.8

Kaplan type, and the chord at the blade tip is finite. The
nozzle 19A has an axial cylindrical part to adapt the finite
blade tip at the propeller plane, and the back side of the
duct is mainly a straight line (cone in 3-D). The pitch of
the propeller varies from 0.8 at the
root to 1.4 at the tip,
%

and the design advance ratio is ' 
! . The gap between duct inner surface and propeller tip is 1mm, which
corresponds to 0.38% of the propeller radius.
 The thrust
#
of the nozzle is normalized as "$#   % '&)(+* , which is
the same way the propeller thrust is normalized. The frictional force on the duct is considered
as follows:


1
1.2

=180

0.2

0.4

",#

  
X

X/C

0.6

0.8

  
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
  

 
 
 
  


 

























  


























 
 
 



 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 



  
































 
   

  



























































































 
 
 

 






 
 

 
 







GBFLOW (inner surface)










 
 
 

























GBFLOW (outer surface)
 
 

 
 






 
 
 
 

EXPERIMENT (inner surface)






 
 
 

 
 















EXPERIMENT
(outer
surface)





















 
 


 
  
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
  
 
  




 

 































 
  
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
































-1.2
-1
-0.8
-0.6
-0.4

CPp p
C

 
 
 
 
  -0.2
















 
   0










"

180

1
0

0.2

0.4





 
 
 


 X
 



X/C



0.6

0.8

Figure 13: Comparison of pressure distributions predicted by present method with those measured in experi

 3
  3 (bottom): DUCT II
ments at 
(top) and 
3
at inflow angle of attack of - . 
. From Kinnas et al.
(2002)
P/D=1.0

measurement
current method

0.4

10KQ

0.3

KT,10KQ

#2#


"

0.8

0.2

KTT

0.1

-0.1
0.4

KTN

0.5

0.6

0.7

0.8

Figure 14: Thrust and torque coefficients measured in experiments and predicted by the present method with the
gap model for Ka 4-70 propeller/Nozzle 19; P/D=1.0.

1
' /
0

(14)

0.4
0.6

where  is a frictional coefficient taken as a function of


%
the Reynolds
number of the duct, ' is the advance ra1
tio, and / is the normalized surface area of the duct with
respect
to the square of the propeller radius. The value

of  is determined using the ITTC formula. The total
thrust of the ducted propeller system is the combination
of the blade thrust, the nozzle thrust and the frictional
correction part, as follows:

0.2

1.2

.- 


"

43
"

5- 
.

(15)

where "$# is the propeller thrust. The predicted (including the gap effects) and measured forces and moments
are plotted in Fig. 14. In this case, the present method
appears to predict the forces well overall, with the agreement becoming somewhat less satisfactory at lower advance ratios.
Dyne (1973) performed series tests on ducted propellers
using the propeller P1452 and varying duct geometries,
and measured the thrust and the torque on the ducted propellers. Propeller P1452 and DUCT No. 15 are chosen to
validate the present method. The ducted propeller has the
following geometric characteristics: the angle of attack of
 3 , and the maximum camber to chord
the duct is - 76

ratio of the duct is 0.060. A propeller diameter,
pitch
 and



6 8:9;9
to diameter ratio of propeller P1452 are



6
and
, respectively. In addition, a radial
clearance between propeller tip and duct inner surface is
1.0mm. Since the propeller chord length at the tip is equal
to zero, which leads to a varying gap ratio, the gap model
is not applied in this case. In order to compare the results in the fully wetted case, Dynes close-to-wetted run
  is considered. The body force distribution preat < 
dicted by the present method, and the duct geometry are
shown in Fig. 15. The swirl distribution inside the duct
is shown in Fig. 16. The predicted thrust and torque are
compared with those measured in experiments in Fig. 17.
In this case the present method seems to overestimate the
propeller thrust and torque. The cause of the discrepancy
is still under investigation.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

PODDED PROPULSORS

Body Force Distribution in GBFLOW-X

Qbx
4.35
4.11
3.88
3.64
3.41
3.17
2.94
2.70
2.47
2.23
1.99
1.76
1.52
1.29
1.05
0.82
0.58
0.35
0.11
-0.12
-0.36

1.5

0.5

0.5

1.5

Figure 15: Body force distribution for ducted propeller


predicted by GBFLOW: P1452 propeller + DUCT No.15
GBFLOW-X Solution

w
0.01
-0.00
-0.02
-0.04
-0.06
-0.08
-0.10
-0.11
-0.13
-0.15
-0.17
-0.19
-0.20
-0.22
-0.24
-0.26
-0.28
-0.30
-0.31
-0.33
-0.35
-0.37
-0.39
-0.40
-0.42
-0.44

1.6
1.4
1.2

1
0.8
0.6
0.4
0.2
0

0.5

1.5

Figure 16: Swirl velocity contours inside DUCT No.15


and P1452 propeller, predicted by the present method
4

measurement
current method

10KTT, 10KTD, 100KQ

100KQ
3

10KTT

With the growing use of podded propulsor units, it is important to model the flow around the pod and strut in the
presence of the propeller. The prediction of the flow is
essential in calculating the pressures on the body, which
when integrated can give the total forces, which are of
importance, in determining the overall performance of a
podded propulsor.
Interaction between propeller and pod
The interaction of the propeller with the pod and strut
requires very extensive calculations. The interaction between the three components is considered iteratively via
coupling of MPUF3A and GBFLOW, as described in the
case of ducted propellers. MPUF3A solves for the pressures, forces and moments in the vicinity of the propeller. The propeller forces from the potential solver are
represented in the Euler solver as body forces. At the
initial iteration, an uniform inflow wake (or given nonaxisymmetric nominal wake) is assumed as the effective
wake in the potential solver. By placing body forces at
appropriate locations and by applying the proper boundary conditions, GBFLOW solves for the fluid flow in the
whole domain using Euler equations. The pod and strut
are represented using the wall boundary condition inside
the domain. It should be noted that in the present method
the hull on which the pod is mounted is presented as an
infinite plate, on which a flow tangency condition is applied. The effective wake distribution is then updated
for the next MPUF3A prediction. The iterations between
MPUF3A and GBFLOW proceed until the convergence
criterion is satisfied. The pictorial representation of the
iterative procedure for the podded propulsor is shown in
Fig. 18. As shown in Fig. 18, the present method can
accommodate 2 blade rows, like contra-rotating, twinrotating or stator-rotor combinations. Computations are
done separately for each propeller (Kinnas et al. 2004)
and the body forces are transfered to GBFLOW at the
appropriate locations. Iterations are performed between
GBFLOW and all the blade rows solved in MPUF3A until the convergent solutions for both propellers are obtained.

10KTD

Grid generation
0

0.4

0.5

0.6

0.7

Figure 17: Thrust and torque coefficients predicted by the


present method and measured: P1452 propeller + DUCT
No.15

The hub radius of the podded propeller is not constant at


the location of the propeller, hence it is essential to adapt
the blade geometry to the pod surface. It is also important that the wake generated by the propeller be adapted
to the pod surface. Figure. 19 shows the wake geometry in the case of two propellers rotating in the same

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Xeff

Xeff

3-D Euler Solver

STRUT

Effective wake

Body Force

Potential Solver

POD

Validations

Effective wake

Body Force

Potential Solver

Figure 18: Pictorial representation of the coupling of the


Finite Volume Method and Vortex Lattice Method
Y

Direction of
Flow

grid is cylindrical around the pod up to a certain radius,


and non-cylindrical beyond it.

Figure 19: Wake geometries for twin propeller system,


rotating in the same direction, also showing the wake
alignment with the pod

direction, as generated by MPUF3A. One propeller is


placed upstream of the strut, and the other downstream
of the strut. It should be noted that although the wake
of the upstream propeller seems to intersect the strut in
Fig. 19, this wake geometry is only modeled in MPUF3A
but not in GBFLOW. The propeller is represented with
body forces in GBFLOW, while the strut is represented
as a solid boundary. Thus within GBFLOW the upstream
propeller wake sheet location and its intersection with the
strut will be accounted for through the solver.
A fully 3-D grid is required to solve for a pod with a strut
in GBFLOW, since the presence of the strut makes the
problem non-axisymmetric. The computational modeling
of the pod and strut requires adaptive computational grids
around non axisymmetric pod and strut. The generated

The present method is validated by comparing the results with those measured in experiments for a modified
KCA-110 series propeller with a constant P/D over the
radius. Results are presented for pull type and twin rotating propellers subject to uniform inflow, and for twin
rotating propellers subject to inflow with yaw angles of
attack. More details, including validations with other inviscid or viscous flow methods in the case of pods, as well
as grid dependence studies of the results from the present
method, are given in Gupta (2004).
The numerical results for pulling propeller at zero degree
angle of attack are compared with the measured values.
Figure 20 shows the predicted pressure contours as well
as the body force distributions for pull type configuration in GBFLOW. The pressures increase at the leading
edge of the pod and strut. The forces are calculated by
integrating the pressures over the surface of the pod and
strut. The axial force acting on the whole unit (the propeller, the pod and strut)% is compared at the four different
values of advance ratio ' . Figure 21 shows the comparison of the total axial forces calculated from the present
method for a pulling propeller with those from the experiments performed by Szantyr (2001). The predicted axial
forces compare well with those at higher advance ratios
but seem to underpredict the force at lower advance ratios. The frictional force on the pod was
considered by

using the empirical ITTC formula for  .
The second case is for the twin type podded propulsor
which one propeller is placed in front of strut and one aft
of it, and both propellers rotate in same directions. The
diameters of both propellers are same as 0.182m and hub
ratios are 0.352. The propeller pitch to diameter ratios of


   , respectively.
front and aft propellers are
and
Figure 22 shows the body force distribution and axial velocity contours predicted by the present method. The flow
is accelerated after it passes the first propeller plane, and
further accelerated after the second propeller which is aft
of the strut. The tangential velocity contours for twin rotating propellers are shown in Fig. 23. The tangential velocity after the front propeller plane is negative below the
pod and positive on the top surface. Due to rotation of the
propeller, the flow at the top surface of the pod goes into
the paper, while at the bottom surface comes out of the
paper. At the aft propeller, the tangential velocities at the
top and bottom of the pod increase due to rotation of the
propeller. The total forces predicted for the twin rotating
propellers are compared with those from Szantyr (2001),
as shown in Fig. 24.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Tangential Velocity in the Center Plane

ZX

body force cells

-0.23 -0.05 0.12 0.30 0.47 0.65 0.83 1.00 1.18 1.36 1.53 1.71 1.88 2.06 2.24

-0.60 -0.51 -0.43 -0.34 -0.26 -0.17 -0.09 0.00 0.09 0.17 0.26 0.34 0.43 0.51 0.60

Figure 20: Body force and pressure contours predicted by


%

GBFLOW for a pull type podded propulsor: ' 
0.2

Figure 23: Tangential velocity contours predicted by


%

GBFLOW for twin type podded propulsor: ' 
! for
both propellers
0.4

0.3

Total Axial Force (F X)

Total Axial Force (F X)

0.15

0.1

0.05

0.1

Experiment (Szantyr, 2001)


Present method
0
0.4

0.5

0.6

0.7

0.2

Experiment (Szantyr, 2001)


Present method
0.8

0.9

JS

Figure 21: Comparison of total axial force for a pull type


podded propulsor

0
0.4

0.5

0.6

0.7

0.8

0.9

JS

Figure 24: Comparison of total axial force for twin type


podded propulsor
-0.09 0.08 0.25 0.42 0.59 0.76 0.93 1.10 1.27 1.44 1.61 1.78 1.95 2.12 2.29
Y
X

0.25 0.36 0.46 0.56 0.67 0.77 0.87 0.98 1.08 1.18 1.29 1.39 1.49 1.60 1.70

Figure 22: Body force distribution and axial velocity


contours for twin type podded propulsor in GBFLOW:
%

' 
for both propellers

Figure 25: Axial velocity contours predicted by


GBFLOW for twin type podded propulsor at yaw angle
 %

 
of attack of 3 : ' 
for both propellers.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

0.4

The above method was also successfully applied to predict the performance of podded propulsors, and the results compared reasonably well with those measured in
experiments in the cases of pull and twin type propellers.
More validations with other experiments as well as with
other methods are required to assess the present method
more thoroughly.

FX

0.3

0.2

Present Method
Szantyr(2001)

0.1

0
-20

-15

-10

-5

10

ACKNOWLEDGMENT

15

20

Yaw angle (degrees)

Figure 26: Comparison of total axial forces predicted by


present method and measured in experiments for twin
%

 
type podded propulsor: ' 
for both propellers.
Finally, the twin rotating podded propulsor subject to inflow with yaw angle of attack is used to validate the
present method. The axial velocity contours and 3-D
view of streamlines around the pod and strut are shown
in Fig. 25. Inflow to the podded propulsor is with yaw
  3 , and the streamlines in the figure
angle of attack of
show clearly the inflow yaw angle. Figure 26 shows the
comparison of total forces measured in the experiment
of Szantyr (2001) with those calculated from the present
method for a twin rotating propeller system at various
yaw angles of attack. Both propellers have an advance
%

 
ratio of ' 
.
CONCLUSIONS
A general computational method, based on the coupling of a potential flow solver applied on the propeller
blades with an Euler solver applied on the flow field,
has been developed to predict the performance of podded
propulsors and ducted propellers. The coupled method
of GBFLOW and MPUF3A is capable of capturing the
interactions between the appendages and the propeller
blades.
Comparisons of the predicted thrust and torque with some
known experiments for ducted propellers showed good
agreement to each other. However, since the gap model
applied to the present method depends on the choice
of the empirical discharge coefficient, more validations
studies with experiments, as well as comparisons with the
results from fully viscous flow solvers, are needed to assess the validity of the gap model. A possible extension
could include the treatment of the leading edge vortex at
the tip in the current method

Support for this research was provided by Phase III,


and IV of the Consortium on Cavitation Performance
of High Speed Propulsors with the following members:
AB Volvo Penta, American Bureau of Shipping, Daewoo Shipbuilding and Marine Engineering Co. Ltd.,
El Pardo Model Basin, Hyundai Maritime Research Institute, Kawasaki Heavy Industries, Michigan Wheel
Corp., Naval Surface Warfare Center Carderock Division
and the Office of Naval Research (Contracts N0001401-1-0225 and N00014-04-1-0287), Rolls-Royce Marine AB, Rolls-Royce Marine AS, VA Tech Escher
Wyss GMBH, W rtsil Propulsion Netherlands B.V. and
W rtsil Propulsion Norway AS. Partial support of this
work was also provided by the Office of Naval Research
under its NNRI program, through Florida Atlantic University (Subagreement TRD67).


References
Abdel-Maksoud, M. & Heinke, H. J. (2002), Scale effects
on ducted propellers, in Twenty-Fourth Symposium
on Naval Hydrodynamics, Fukuoka, Japan.
Atlar, M., ed. (2004), First International Conference on
technological Advances in Podded Propulsion, School
of Marine Science and technology, University of Newcastle, UK.
Brewer, W. H., Newman, J. C., Burgreen, G. W. & Burg,
C. O. E. (2003), A design method for investigating cavitation delay, in The 8th International Conference on
Numerical Ship Hydrodynamics, Busan, Korea.
Chesnakas, C. & Jessup, S. D. (2003), Tip-vortex induced cavitation on a ducted propulsor, in International Symposium on Cavitation Inception, ASME
FED Summer Meeting, Honolulu.
Choi, J.-K. & Kinnas, S. A. (2000), An unsteady threedimensional euler solver coupled with a cavitating propeller analysis method, in The 23rd Symposium on
Naval Hydrodynamics, Val de Reuil, France.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Choi, J.-K. & Kinnas, S. A. (2001), Prediction of nonaxisymmetric effective wake by a 3-D Euler solver,
Journal of Ship Research 45(1), 1333.
Choi, J.-K. & Kinnas, S. A. (2003), Prediction of
unsteady effective wake by a euler solver/vortexlattice coupled method, Journal of Ship Research
47(2), 131144.
Chorin, A. J. (1967), A numerical method for solving incompressible viscous flow problems, Journal of Computational Physics 2, 1226.
Dyne, G. (1973), Systematic studies of accelerating
ducted propellers in axial and incline flows, Symposium on Ducted Propellers, RINA pp. 114124.
Falcao de Campos, J. (1983), On the calculation of ducted
propeller performance in axisymmetric flows, Technical Report 696, Netherlands Ship Model Basin, Wageningen, The Netherlands.
Feng, J. & Dong, S. (1985), A method for the prediction of unsteady of unsteady hydrodynamic performance of the ducted propeller with a finite number of
blades, Technical Report 85006, China Ship Scientific
Research Center, Wuxi,China.
Ghassemi, H. & Allievi, A. (1999), A Computational
Method for the Analysis of Fluid Flow and Hydrodynamic Performance of Conventional and Podded
Propulsion Systems, Oceanic Engineering International 3(1), 101115.
Gibson, I. & Lewis, R. I. (1973), Ducted propeller analysis by surface vorticity and actuator disk theory, in
Proceedings of the Symposium on Ducted Propellers,
RINA, Teddington, England.
Glover, E. & Ryan, P. (1973), A comparison of the theoretical and experimental performance of a ducted propeller system, in Symposium on Ducted Propellers,
The Royal Institution of Naval Architects, Teddington,
England.
Gupta, A. (2004), Numerical prediction of flows around
podded propulsors, Masters thesis, The University of
Texas at Austin.
Hsin, C.-Y., Chou, S.-K. & Chen, W.-C. (2002), A new
propeller design method for the pod propulsion system, in Proceedings of Twenty-Fourth Symposium on
Naval Hydrodynamics, Fukuoka, Japan.
Hughes, M. (1993), Analysis of Multi-component Ducted
Propulsors in Unsteady Flow, PhD thesis, M.I.T., Department of Ocean Engineering.

Hughes, M. J. (1997), Impementation of a special procedure for modeling the tip clearance flow in a panel
method for ducted propulsors, in Propellers/Shafting
97 Symposium, Virginia Beach, VA.
Hughes, M. & Kinnas, S. (1993), Unsteady flows around
multi-component integrated propulsors, in Forum on
Unsteady Flows, (FED-Vol. 157), Fluids Engineering
Division, ASME, pp. 2131.
Hughes, M., Kinnas, S. & Kerwin, J. (1992), Experimental validation of a ducted propeller analysis method,
Journal of Fluids Engineering 114(2), 214219.
HYDROCOMP Inc. (1999), Modeling tractor-style azimuthing podded drives, Technical Report 127, HYDROCOMP Inc.
Kakar, K. (2002), Computational Modeling of FPSO Hull
Roll Motions and Two-component Marine Propulsion
Systems, Masters thesis, Department of Civil Engineering, The University of Texas at Austin.
Kerwin, J. E., Kinnas, S. A., Lee, J.-T. & Shih, W.-Z.
(1987), A surface panel method for the hydrodynamic
analysis of ducted propellers, Transactions of Society
of Naval Architects & Marine Engineers 95.
Kerwin, J., Keenan, D., Black, S. & Diggs, J. (1994), A
coupled viscous/potential flow design method for wake
adapted multi-stage, ducted propulsors using generalized geometry, Transactions of Society of Naval Architects & Marine Engineers 102.
Kerwin, J., Taylor, T., Black, S. & McHugh, G. (1997), A
coupled lifting-surface analysis technique for marine
propulsors in steady flow, in Propellers/Shafting 97
Symposium, Soc. Naval Arch. & Marine Engnrs., Virginia Beach, VA, pp. 115 (Paper No. 20).
Kinnas, S. A., Choi, J.-K., Lee, H. S., Young, Y. L.,
Gu, H., Kakar, K. & Natarajan, S. (2002), Prediction
of cavitation performance of single/multi-component
propulsors and their interaction with the hull, Transactions of The Society of Naval Architects & Marine
Engineers .
Kinnas, S. A. & Coney, W. B. (1992), The generalized
image model - an application to the design of ducted
propellers, Journal of Ship Research 36(3), 197209.
Kinnas, S. A., Gu, H., Gupta, A. & Lee, H. S. (2004),
Numerical prediction of the performance of podded
propulsors and ducted propellers, in The 13th offshore
symposium : The Application of Emerging Technologies Offshore, Texas Section of The Society of Naval

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Architects & Marine Engineers, Houston, TX, pp. 19


34.
Moon, I. S., Kim, K. S. & Lee, C.-S. (2002), Blade tip
gap flow model for performance analysis of waterjetpropulsors, in International Association for Boundary
Element Methods, Austin, TX.
Morgan, W. & Caster, E. (1965), Prediction of the aerodtnamics characteristics of annular airfoils, Technical
Report Report 1830, David Taylor Model Basin.
Ni, R.-H. (1982), A multiple-grid scheme for solving the
Euler equations, AIAA Journal 20(11), 15651571.
Oosterveld, W. C. (1970), Wake adapted ducted propellers, Technical Report No. 345, Netherlands Ship
Model Basin, Wageningen, Netherlands.
Sanchez-Caja, A., Rautaheimo, P. & Siikonen, T. (1999),
Computation of the incompressible viscous flow
around a tractor thruster using a sliding-mesh technique, in Proceedings of Seventh International Conference on Numerical Ship Hydrodynamics, France.
Sanchez-Caja, A., Rautaheimo, P. & Siikonen, T. (2000),
Simulation of incompressible viscous flow around
a ducted propeller using a rans equation solver, in
Twenty-Third Symposium on Naval Hydrodynamics,
Valde Reuil, France.
Szantyr, J. (2001), Hydrodynamic model experiments
with pod propulsors, Oceanic Engineering International 5(2), 95103.
Van Houten, R. (1986), Analysis of ducted propellers
in steady flow, Technical Report 4.76-1, Airflow Research and Manufacturing Corp., Watertown, MA.
Warren, C., Taylor, T. & Kerwin, J. (2000), A coupled viscous/potential-flow method for the prediction
of propulsor-induced maneuvering forces, in Propellers/Shafting 00 Symposium, Soc. Naval Arch. &
Marine Engnrs., Virginia Beach, VA.
White, F. (1986), Fluid Mechanics, 2nd edn, McgrawHill Book Company.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Prof. Mehmet Atlar
University of Newcastle, United Kingdom
The authors present numerical modelling of
the flows around two types of unconventional
propulsors: ducted and podded propellers which have
substantial stationary components which are
subjected to complex hydrodynamic interaction as
well as large viscous drag forces under the effect of
propellers action. In order to tackle the complex
interaction problem the authors present a
computationally efficient, hybrid and iterative flow
model.
Within the framework of the method
presented in this paper and the hydrodynamic
interaction phenomenon involving these devices, say
for a pulling type podded propeller, one can consider
three types of interaction: Firstly, the effect of the
propeller flow on the pod and strut would appear as
an increase of the axial and tangential velocities
combined with an increase of the pressure field
behind the propeller. This effect is represented by the
body forces in the Euler solver of the authors paper.
Secondly, the presence of the pod and strut in the
slipstream of the propeller will slow down the inflow
velocity and thirdly, the latter effect will also deform
the propeller wake. The reduced flow effect can be
taken into account by the measured velocity field
(nominal wake) in the presence of the pod/strut
housing as input to the lifting surface code. However,
it is not common practice to measure the wake in the
presence of the pod/strut housing unless for a
particular reason. Can the Authors shed a light on
how this interaction is modelled effectively using
only hull inflow data in the absence of the pod/strut
housing? As far as the slipstream deformation is
concerned, it is not clear in the paper, at least to the
discusser, how this effect is taken into account in the
potential propeller solver and/or in the Euler solver. It
will be therefore helpful if the authors can give a
brief description details for the account of this effect.
In fact Fig. 19 of the paper indicates no visible
deformation of the wake as it intersects the strut.
However the physics of the phenomena may present
rather complex interaction problem for heavily
loaded propulsor as one can see in Fig. 1 from the
cavitating tip vortex trajectories for a pulling type
podded propeller operating in bollard pull condition.
The degree of slipstream deformation and its
modelling can be argued depends upon the axial
distance from the propeller and hence relative
position of the strut to the propeller, which may vary
depends upon particular pod type, requiring careful
modelling. Within this respect although there is

reasonable measured slipstream data for conventional


propellers,
similar
type
information
for
unconventional propellers, particularly for podded
propeller, is scarce. In a recent study the discusser
and his colleagues presented an experimental
investigation into the velocity field of a podded
propeller by LDA and the use of this information for
the assessment of propulsor wash and interaction
between the hull and propulsors [2].

Fig 1. A view of cavitating slipstream trajectories of


a podded propeller operation in bollard pull condition,
Atlar et al [1]
Finally, in comparison to RANS based flow
solvers, the present approach makes the major
assumption that viscosity plays a rather insignificant
role in the above mentioned interaction phenomenon.
This may be justified near the propeller since the
interaction in this region will be dominated by
potential vortex flow. However, in the power
prediction of these propulsors, it is customary to treat
the propulsor as a whole unit including its large
stationary elements and the propeller. Within this
framework the accurate prediction of the resistance of
these large non-stationary parts, which is mainly
viscous, under the effect of the operational propeller
is important. For example, once again considering a
pulling type podded propulsor, an accurate estimation
of the thrust for the whole unit, which includes the
drag of the pod and strut under the effect of the
propellers action, in full-scale is critical. The viscous
pressure drag of these stationary parts under the
effect of propeller, can be much larger than the
frictional drag component, as reported in e.g. [3], and
their accurate predictions at large Re numbers in fullscale is essential. Within this respect it is perhaps
worthwhile the authors to look into coupling of their
potential flow solver with RANS. This will also
provide ability to optimise the pod housing
effectively from the drag point of view.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

1. Atlar, M., Aryawan, I., Prasetyawan, I. and Sasak,


N., Cavitation in Ice-Milling with a Podded
Propulsor, Proceedings of FEDSM03 , 4TH
ASME_JSME Joint Fluids Engineering Conference,
Honolulu, Hawaii, USA, July 6-11, 2003.
2. Wang, D., Atlar, M., Glover, E.J. and Paterson, I.,
Experimental Investigation of Flow Field around a
Podded Propulsor using LDA, Proceedings of TPOD, First Intl Conference on Technological
Advances in Podded Propulsion, University of
Newcastle, UK, April 14-16, 2004.
3. Lobachev, M.P. and Tchitcherine, I.A. The FullScale Resistance Estimation for Podded Propulsion
System by RANS Method, Proceedings of
International Symposium on Ship Propulsion, SP
2001 Lavrentiev Lectures. The St.-Petersburg State
Marine Technical University, The Inst. of Marine
Engs. and Central Marine research & Design Inst.
Saint-Petersburg, Russia, June 19-21, 2001

quite close to the time-average of the unsteady 3-D


flow as predicted by a fully unsteady version of
GBFLOW. This is accomplished by allowing for
non-axisymmetric loading on the propeller blades in
GBFLOW. In this way the unsteady interaction
between the tip vortex of the upstream propeller and
the strut is not captured. On the other hand the
propeller wake is adapted to the pod geometry and is
aligned to the circumferentially averaged flow. To
model the tip-vortex/strut unsteady interaction a fully
unsteady wake alignment method (similar to that of
Lee & Kinnas, 2004) should be applied in MPUF-3A,
and the unsteady version of GBFLOW must be
utilized.
The authors agree with Prof. Atlar that the
viscous effect on the pod (form drag) could be
important (in terms of the total thrust of the unit) for
some operating conditions and some types of podded
propulsors. We plan on coupling our potential solver
with a RANS code in the near future, in order to
assess the effect of the form drag of the pod.

AUTHORS REPLY

REFERENCE

We would like to express our thanks to Prof.


Atlar for his valuable comments and suggestions on
our paper and on podded propulsors in general.
In the first question, he asked how the
interaction was modelled using the hull inflow data in
the absence of pod/strut. In our method, the Euler
solver (GBFLOW) which solves for the flow field
around the pod and strut, and the potential flow
solver (MPUF3A) which solves for the propeller flow
subject to a given inflow, are coupled in an iterative
manner. The iterations between the two methods are
performed until the propeller forces converge within
a desired accuracy. As an input to MPUF3A in the
first iteration, we have used a uniform inflow (i.e. the
inflow far upstream of the podded propeller) or
nominal wake as the inflow wake, and this inflow
is updated (effective wake) during the iterations. So,
in other words what we call nominal wake in this
paper is NOT the inflow in the absence of the
propeller, but rather is the inflow in the absence of
BOTH, the propeller and the pod/strut. The effective
wake is calculated in GBFLOW by solving for the
total flow field (less the induced velocities) around
the pod and the strut with the inclusion of propeller
body forces, and thus the resulting flow from
GBFLOW already includes the effect of pod and strut.
In regards to the second question on the
propeller slipstream deformation, in the present work
the time averaged interaction between the propeller(s)
and the pod/strut are considered. As shown in Choi &
Kinnas (2003) in the case of open propellers this 3-D
version of GBFLOW predicts a 3-D flow which is

1. Lee, H. and Kinnas, S.A. Unsteady Wake


Alignment for Propellers in Non-axisymmetric
Flows, Journal of Ship Research (in printing), 2004.

REFERENCES

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Dr. Vladimir I. Krasilnikov


MARINTEK - Norwegian Marine Technology
Research Institute, Norway

axial, KY transverse, MZAX strut axial moment) are


compared with the measurements at J=0.299 in the
range of heading angles [45;+45] in Fig.2. These
1,0

Pull, zero heading


Experiment
BEM Calc.

0,9

KTP, KQP, 0P

0,8
0,7
0,6
0,5
0,4
0,3
0,2
0,1
0,0
0,0 0,1 0,2 0,3 0,4 0,5 0,6 0,7 0,8 0,9 1,0 1,1 1,2 1,3

J=V/(nD)

Fig.1: Pulling propeller characteristics at zero


heading.
0,6

Pull, J=0.299

0,12
0,10

0,5

KX

0,4

KX, KY

This interesting paper continues the line of


work conducted by Prof. Kinnas and his group on
numerical modeling of marine propulsors using a
coupled Euler/Vortex-Lattice Method algorithm,
which is, undoubtedly, a step forward compared to
the potential velocity field iteration methods widely
used for these purposes. Among the advantages of
this approach the main is that it is capable of correct
solution of the vortical inflow to predict the effective
wake field (assuming, of course, negligible role of
viscosity between the propeller and the inlet
boundary of the Euler flow domain).
I have a few questions regarding numerical
implementation of the authors method.
1) The effective wake distribution is defined
at the plane located upstream of the leading edge of
the propeller assuming its small change through the
site of propeller. It is justified for a blade row
installed on the shaft, but needs more attention in the
case of podded propeller where tapered pod end and
close location of the strut may cause a significant
change of the velocity components in the axial
direction. On the other hand, the tangential velocity
induced by propeller also changes rapidly across the
propeller disk. Did the authors investigate these
effects and their contribution to the final solution?
2) Applying a VLM to the analysis of the aft
propeller or strut/rudder located in the wake of the
forward stage (for example, to estimate cavity
pattern) we use Bernoulli equation in order to define
pressure. In this case it needs, however, a proper
correction since Bernoulli constant reveals a jump
across the forward propeller plane. In [2] the authors
solved this problem for the rudder behind propeller
applying Bernoulli equation between a point on the
rudder surface and a point on the same streamline far
downstream, which is very similar to the approach
employed by the discusser in [1]. What do they do in
the case of the aft propeller and in the case of strut
located between two blade rows?
In the second part of my discussion I would
like to propose a comparative calculation with the
pulling single-screw thruster investigated in [1]
experimentally and using a velocity based BEM,
which features a simple velocity field iteration
technique. The results of prediction of the
performance of propeller operating as a part of the
unit (KTP without pod/strut drag, KQP and 0P) in the
range of J-s corresponding to the zero heading are
given in Fig.1. The total forces on the unit (KX

0,08

0,3

0,06

0,2

0,04

0,1

0,02

0,0

0,00
-0,02

-0,1
-0,2

-0,04

KY

-0,3
-0,4

MZAX

solid lines - EXP.


dash lines - CALC.

-0,5
-50

-40

-30

-20

-10

10

-0,08
-0,10

-0,6
-60

-0,06

20

30

40

50

-0,12
60

Heading angle , [deg.]

Fig.2: Measured and calculated maneuvering forces


on the pulling unit at J=0.299.

results have been obtained in [1] using a quasi-steady


calculation with the assumption about non-deformed
propeller wake, which follows propeller axis.
Although additional corrections have been made to
account for the viscous pod/strut drag. Could the
authors present their results with the coupled
Euler/VLM algorithm for the same conditions?
I would like to congratulate the authors for
such thorough and thought-provoking paper.
REFERENCES
[1] Achkinadze, A.S., Berg, A., Krasilnikov, V.I. and
Stepanov, I.E. Numerical Analysis of Podded and
Steering Systems Using a Velocity Based Boundary
Element Method with Modified Trailing Edge, Proc.
of the Propellers/Shafting 2003 Symposium, Virginia
Beach, VA, Sep 17-18, 2003, pp. 12/1-22.
[2] Lee, H., Kinnas, S., Gu, H. and Natarajan, S.
Numerical Modeling of Rudder Sheet Cavitation

Copyright National Academy of Sciences. All rights reserved.

MZAX

DISCUSSION

Twenty-Fifth Symposium on Naval Hydrodynamics

Including Propeller/Rudder Interaction and the


Effects of a Tunnel, Proc. of the 5th International
Symposium on Cavitation (CAV2003), Osaka, Japan,
Nov 1-4, 2003.
AUTHORS REPLY
The authors would like to thank Dr.
Krasilnikov for his comments on our paper and for
providing us with the geometry and measured
performance of a pulling type propeller.
In the response to his first question, we
agree that the location where the effective wake is
calculated may be important in the prediction of the
effective wake velocity, especially in the case the pod
geometry changes significantly along the axial
direction. In order to investigate the effect of the
location on the propeller performance, the effective
wakes were evaluated at different locations in front
of the downstream propeller for the twin podded
propeller configuration presented in the paper. The
predicted mean forces are shown in the next table.
The propeller mid-plane location is at x=1.1899, and
the blade leading edge is at x=0.76.
Effective wake Location
K
K
0.7507
0.7151
0.6543

0.1248
0.1213
0.1267

0.2426
0.2357
0.2454

As seen from the table, the predicted forces


do not vary significantly. Nevertheless, the authors
are developing a new procedure in which the
effective wake will be evaluated at the actual
propeller control points, instead of just upstream of
the propeller plane.
In regards of the second question, as Dr.
Krasilnikov mentioned, in the case of a rudder
downstream of a propeller we applied the Bernoulli
equation between a point on the rudder surface and a
point on the same streamline far downstream. In the
present work however, the strut as well as the pod are
being modelled as a solid boundary in the Euler
solver, and thus the pressure is being determined
directly from the solution. However, for the aft
propeller, we need to evaluated the pressure using the
Bernoulli equation, and the pressure upstream of the
downstream propeller has to be taken into account,
especially when we need to predict sheet cavitation
on the blades. Figure 1 shows the pressure contours
inside the fluid domain (please note that the shown
pressure is only the difference from the hydrostatic
pressure). The locations of for and aft propellers are
depicted as well in the figure. As shown in the figure,
the pressure just upstream of the downstream
propeller is very small, which means that hydrostatic

pressure practically applies at that station. Thus,


based on this observation we apply the Bernoulli
equation for the aft propeller using hydrostatic
pressure as the inflow pressure. In fact the authors
plan to include this pressure distribution upstream of
the aft propeller, in case it varies significantly from
hydrostatic.

Figure 1: Pressure contours in the fluid domain: twin


type podded propulsor.
As a part of his second question, Dr.
Krasilnikov proposed a comparative calculation for a
pulling type thruster, the performance of which was
investigated experimentally and numerically. Figure
2 shows the propeller performance for a range of
advance ratios corresponding to zero heading. In
general, our method predicts very well the
performance of the propeller for higher values of the
advance ratio. However, for lower advance ratios the
current method seems to under-predict the propeller
forces. It should be noted that the shown results from
the method of Achkinadze et al use a semi-empirical
approach to increase the loading of the propeller at
lower advance ratios, and thus their method appears
to predict the forces more accurately. It is the
authors belief that a modification of the current
method which models the effects of a leading edge
vortex in a rigorous fashion should be more
successful in predicting the forces at lower advance
ratios.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Figure 2: Performance of pulling type podded


propulsor at zero heading.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Moustafa Abdel-Maksoud
Duisburg-Essen University, Germany
Congratulations for the impressive work! It
connects the advantages of using boundary element
methods and field methods.
Boundary element methods are very
valuable for propeller design, and the short
calculation time makes the combination with an
optimisation method very effective.
Of course, the application of these methods
is not recommended when it is important to consider
geometrical boundaries very close to the propeller as
a duct or a pod. But this limitation can be avoided by
combination with a field method such as Euler or
RANSE method. Therefore, the presented paper is
very valuable not only for academic reasons but also
for the practice.
In this context I have some questions:

Why did you choose an Euler method and not a


RANSE one? The effort to combine the
MPUF3A with an Euler or RANSE solver using
body force method is nearly the same. I think
that a RANSE solver will deliver much more
information, especially friction losses.

The problem of the gap region of a ducted


propeller is very complicated. The using of the
gap model is a good compromise. I `think, the
problem of this model is the determination of the
values CQ and Cp. The extrapolation of the
pressure on both sides of the blade is very critical
due to the strong local variation of the pressure
in this region. The value of CQ depends on thrust
loading condition. As ducted propellers are used
at high loading condition, do you have any
recommendations for the values CQ and Cp?

There are many possibilities for implementation


of the body forces method for a podded
propulsor in a Euler method such as :
1.

2.

body forces are considered at the actual


location of chord length of the propeller
blades (variation of the thrust loading in the
radial and circumferential directions)
body forces are distributed over a disc
(variation of the thrust loading only in the
radial direction)

The first possibility means that the


numerical calculation has to be treated as an unsteady
problem. The second possibility is not accurate

enough for this application. Do you use any method


to approximate the variation of the propeller thrust in
the circumferential direction?
Thank you again for your very interesting
work.
AUTHORS REPLY
We would like to thank Prof. AbdelMaksoud for reviewing our paper and for making
valuable comments.
Regarding the use of an Euler solver vs. a
RANS solver in the present method, the authors agree
that the RANS solver would provide more
information, such as friction losses and the details of
the viscous flow in the tip gap, than the Euler solver.
However, it is well known that a RANS solver
requires a significantly larger number of cells (thus
significantly larger memory, storage, and CPU time)
in order to capture the boundary layers on the duct
surface as well as the viscous flow in the tip gap. At
this stage of our research we have decided to use an
Euler solver, since the current method seems to
predict ducted propeller performance well over a
wide range of advance ratios, and at the same time to
be computationally efficient. On the other hand, to
assess the error in the numerical solution when using
Euler vs. RANS, we plan on coupling our method
with a RANS solver as well. In the case of open
propellers the authors have shown (see Kinnas et al
2002) that in the vicinity of the propeller the flow is
primarily governed by inviscid vorticity dynamics
and that an Euler solver (applied from a station
upstream of the propeller at which the boundary layer
on the hull is known from measurement or a RANS
calculation) should be adequate.
In response to the second question, first we
wish to clarify that this simplified semi-empirical
model is only being used in order to predict the effect
of the viscous flow in the gap on the propeller
loading and performance. From (Hughes 1997, Moon
et al. 2002), a C Q value of 0.85 is suggested.
However, this value of C Q may depend on the blade
gap percentage, the tip gap shape, and the blade
loading. The authors plan on determining the
appropriate value of the discharge coefficient by
calculating the viscous flow around a stationary foil
or a rotating blade in the vicinity of a solid wall and
by correlating the circulation distribution and forces
on the blade with those produced by the current
model. The proposed comparison will be made over a
variety of geometries and operating conditions in
order to study their effect on the value of the
discharge coefficient. In regards to Cp, the details of

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

the algorithm to evaluate it are given in (Gu and


Kinnas 2003). In fact the effects of the gap desingularize the pressure distribution at the tip and
that makes the extrapolation process of the blade
pressures into the gap more successful.
In regards to the third question, our present
model for the body force is similar to that mentioned
in the discussers first possibility, i.e. the propeller
body force vectors in the present Euler solver have
components in all directions and their magnitudes can
vary in the axial as well as in the circumferential
direction. This is the often called time-averaged 3-D
model. Choi (2000) and Choi and Kinnas (2003)
have also developed a model in which the unsteady
3-D Euler equations are solved with respect to a
coordinate system which rotates with the propeller.
In the same publications it is shown that the current
time-averaged 3-D version of the Euler solver
(used in the present paper) produced a flow field
which was quite close to the time-average of the
unsteady 3-D flow field as predicted by the fully
unsteady version of the Euler solver.
REFERENCES
1. Gu, H. and Kinnas, S.A. Modeling of ContraRotating and Ducted Propellers via Coupling of a
Vortex-Lattice with a Finite Volume Method,
Propeller & Shafting 2003 Symposium, SNAME,
Virginia Beach, VA, Sept. 2003.
2. Choi, J-K, Vortical Inflow-Propeller Interaction
using unsteady Three-Dimensional Euler Solver,
Ph.D Thesis, Department of Civil Engineering, The
University of Texas at Austin, August, 2000.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics


St. Johns, Newfoundland and Labrador, CANADA, 8-13 August 2004

Experimental Uncertainty Analysis for


Ship Model Testing in the Ice Tank
Ahmed Derradji-Aouat (National Research Council of Canada, Institute for Ocean
Technology, St. Johns, Newfoundland, Canada)
ABSTRACT

INTRODUCTION

Historically, until late 1980s, only marginal


work on Experimental Uncertainty Analysis (EUA)
was reported by ocean/marine test facilities. During the
1990s, the International Towing Tank Conference
(ITTC) and the International Ship and Offshore
Structure Congress (ISSC) have recommended and
supported the application of Uncertainty Analysis
(UA)
in
both
experimental
and
numerical/computational fields.
The work presented in this document deals
exclusively with Experimental Uncertainties (EU) in
the results obtained from testing of model ships in a
typical ice tank testing facility. Up to now, in the
literature, there are no standards to quantify and/or
minimize uncertainties in ice tank testing.
The objective of this work is to develop a
method of analysis for EU in typical ice tank ship
experiments. In fact, this objective is a task for the 24th
ITTC Specialist Committee on Ice (2002-2005).
To achieve this objective, experiments for
ship resistance in ice were conducted at the Institute for
Ocean Technology of the National Research Council of
Canada (www.iot-ito.nrc-cnrc.gc.ca/) using a model
for the Canadian Icebreaker The Terry Fox. The
data obtained from these tests was used to develop a
procedure for EUA in ice tank ship resistance tests.
From the project management point of view,
the ice tank test program was divided into several
phases to accommodate the planning for opportunity
testing. So far, three phases of testing have been
completed. Phases I and II of the test program were
already documented (Derradji-Aouat, 2004).
In this paper, the results from Phase III are
reported. The methodology developed to quantify EU
in the test results is presented and validated. Also,
comparisons of test results and analyses from the
previous tow phases (Phases I and II) of the test
program are compared to those from Phase III.

Experiments for ship model resistance in ice


were conducted at the Institute for Ocean Technology
of the National Research Council of Canada (www.iotito.nrc-cnrc.gc.ca/). These tests were conducted for the
ITTC 23rd and 24th ice specialty committees (mandate
period 1999-2002 and 2002-2005, respectively). One
of their main tasks is to develop a procedure for
Experimental Uncertainty Analysis (EUA) in ice tank
testing. So far, three phases of ship resistance in ice
testing have been completed. The test matrix for all
three phases is given in Table 1.
The results and analysis from the experiments
in Phase III test program are presented in this report.
However, for clarity and completeness, a brief
summary regarding the previous tow phases is given as
follows:
Phase I test program was documented by
Derradji-Aouat et al. (2002) and Derradji-Aouat (2002)
in two IOT internal technical reports: The 1st report
dealt with presenting the experimental program and
test results, while the 2nd report dealt with developing a
methodology to quantify EU in the test results.
The documentation for Phase II test program
is also presented in two IOT reports (Derradji-Aouat
and Coff, 2003, and Derradji-Aouat, 2003). In Phase
II, the same test matrix as in Phase I was repeated
(Table 1). The only difference is the target thickness of
the ice. In Phase I, all tests were conducted for only
one target ice thickness (40 mm), while Phase II tests
were conducted for two additional ice thicknesses (25
mm and 55 mm). Together, Phases I and II test
programs provided information for three different ice
thicknesses.
In Phase III, the same test matrix as in Phase I
was completed (Table 1). The difference between
Phase I and Phase III test programs is that in Phase I,
the ship model was attached to the carriage using the
tow post (Figure 1a), while in Phase III, the model was
attached to the carriage using the PMM (Planar Motion
Mechanism, Figure 1b).

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

1998), and GUM (2003). The necessary mathematics


used to develop an EUA procedure for ice tank testing
is given in Appendix B.
SHIP RESISTANCE IN ICE
Since the objective of this paper is to present a
procedure for EUA in the results of ship resistance
tests in ice tanks, a summary for the standard
calculations of ship resistance in ice is given as
follows:
The standards for ship resistance in ice
(ITTC-4.9-03-03-04.2.1) give the equation for the total
resistance in ice as the sum of 4 individual
components:
Figure 1a: Ship model in the tow post test setup.

R =R +R +R +R
t
br
c
b
ow

(1a)

where Rt is the total resistance, Rbr is the resistance


component due to breaking the ice, Rc is the
component due to clearing the ice, Rb is the component
due to buoyancy of the ice and Row is the resistance
component in open water.
In order to quantify each component, the test
plan should include tests in level ice, tests in pre-sawn
ice, creeping speed tests, and tests in open water. The
open water tests provide values for Row, while the
creeping speed tests give Rb. In the pre-sawn ice tests,
the ice breaking component Rbr = 0, and therefore:

R =R +R +R
t
c
b
ow

(1b)

Since Row and Rb are already known (from the


open water and the creeping speed tests), thus:

Figure 1b: Ship model in the PMM test setup.


All three phases involved experiments in ice
and in open water. A total of sixteen (16) different ice
sheets were tested. Table 1 gives the nominal ice
thickness and the target flexural strength for each
sheet. It should be noted that all experiments in ice
were very long test runs (the model was towed at
constant speed throughout most of the useable length
of the ice tank, 76 m). Appendix A provides a
summary for Phase III test program.

= R R R
t
b
ow

(1c)

where Rt, in Eq. 1c, is the measured resistance in the


pre-sawn ice test runs.
From tests in level ice, the total resistance Rt
is measured, and the ice breaking component, Rbr, is
calculated as (from Eq. 1a):
= R R R R
t
c
b
ow

(1d)

EXPERIMENTAL UNCERTAINTY ANALYSIS

A literature review for the history and


development of EUA in marine/ocean testing facilities
was given by Derradji-Aouat (2002).
Mathematically, the EUA procedure presented
in this report is based on the equations provided by
Coleman and Steel (1998). The latter is in harmony
with the guidelines of ISO (1995), ASME (PTC-19.1,

EUA - PROPOSED PROCEDURE FOR ICE


TANK TESTING

br

The proposed procedure was developed on the basis of


one hypothesis and one requirement.
1. Segmentation hypothesis, and
2. Steady state requirement.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

2.
Segmentation Hypothesis
For the test runs in ice, several reasons have
contributed to the decision for keeping the speed of the
ship model constant throughout most of the useable
length of the ice tank (about 65 m Appendix A). The
main one is the hypothesis that the time history from
one long ice test run can be divided into segments, and
each segment can be analyzed as a statistically
independent test. The hypothesis states that:

3.

Condition # 1 is based on the fact that the ITTC


procedure for resistance tests in level ice (ITTC-4.903-03-04.2.1) requires that a test run should span over
at least 1.5 times the model length. For high model
speeds (> 1 m/s), however, the ITTC procedure
requires test spans of 2.5 times the model length.

The history for a measured parameter (such as tow


force versus time) can be divided into 10 (or more)
segments, and each segment is analyzed as a
statistically independent test. Therefore, the 10
segments in one long test run are regarded as 10
individual (independent but identical) tests.

Condition # 2 is based on the fact that in EUA, for an


independent test, a population of at least 10 data points
is needed to achieve the minimum value for the factor t
(Eq. B.2 - Appendix B). The gain in any further
reduction in the value of t, by having more than 10
segments, is minimum (Derradji-Aouat, 2003).

Coleman and Steel (1998) reported that, in


statistical uncertainty analysis, a population of at least
10 measurements (10 data points) is needed.
Uncertainty is calculated using the mean and the
standard deviation of that population.
However, in ice tank testing, conducting the
same test 10 times is very costly and very time
consuming. Therefore, the principle of segmenting a
time history of a measured parameter over a long test
run into 10 segments, results in significant savings in
costs and efforts. In this case, EU are calculated from
the means and standard deviations of the individual
segments.
In order to further illustrate the segmentation
hypothesis, the following example is used (Figure 1a).
In this example, the measured tow force history is
obtained in test run # 1 (level ice), ice sheet # 2
(nominal ship speed of 0.2 m/s and nominal ice
thickness of 40 mm).
The segmentation hypothesis calls for
dividing the long time history (Figure 1a) into 10 equal
(more or less equal) segments, as illustrated in Figure
1b. Using the tow force segments, the first step is to
calculate the mean and standard deviation for each
segment. The second step is to calculate the mean of
the means and the standard deviation of the means of
the segments.
These two steps of calculations are repeated
for all test runs in ice (continuous, pre-sawn and
broken ice test runs, all six test runs, Appendix A )
It should be cautioned that the segmentation
hypothesis is valid only if the following 3 conditions
are satisfied (Derradji-Aouat, 2004):
1.

Each segment should include at least 10 events for


ice breaking (10 ice load peaks) or at least 10
collision events (in case of pack ice test runs), and
General trends (of a measured parameter such as
tow force versus time) are repeated in each
segment.

Condition # 3 is introduced to ensure that the overall


trends in a measurement are repeated in each segment.
This condition serves to provide further assurance into
the main hypothesis (the 10 segments in one long
test run are regarded as 10 individual, independent but
identical, tests). Fundamentally, if the trends are not,
reasonably, repeated, then the segments could not be
analyzed as independent but identical tests.
Table 1: Test matrix.

Each segment should span over 1.5 to 2.5 times


the length of the ship model,

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

200

PICE-2, Run 1, V = 0.2m/s

Tow Force (N)

150
100
50
0
-50
90

140

190

240

Tim e (s)

290

340

Figure 2a: Example for a measured tow force history (PICE-2 refers to ice sheet #2).
120

200

Segment 4
Tow Force (N)

Tow Force (N)

80
40
0

90

100

Time (s)

110

110

Time (s) 130

140

Segment 7

100

50
0

50
0
-50

-50
140

120

150

Segment 6

100

150

Time (s) 160

150

170

170

180

Time (s)

190

150

Segment 8

100

200

Segment 9

100

Tow Force (N)

Tow Force (N)

50

120

Tow Force (N)

Tow Force (N)

150

50
0

50
0

-50

190

200 Time (s) 210

220

150

-50

230

220

250

Segment 11

100

50
0

50
0

-50

-50
260

Time (s)

270

200

280

280

290

Time (s)

300

200

Segment 12

150

310

Segment 13

Tow Force (N)

150
100

100
50
0
-50
305

Time (s) 240

150

100

250

230

200

Segment 10

Tow Force (N)

Tow Force (N)

100

-50

-40

Tow Force (N)

Segment 5

150

50
0
-50

315

Time (s) 325

335

335

345

Time (s)

355

365

Figure 2b: Division of the tow force time history (in Figure 2a) into segments.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

The time histories measured in creeping speed


tests are not subjected to the segmentation hypothesis.
Furthermore, it is recognized that the division of the
results of a test run into segments is valid only for the
steady state portion of the measured data (only the
steady state portion of the measured time history is to
used for the segmentation). This is required to
eliminate the effects of the initial ship penetration into
the ice (transient stage) and the effects of the
slowdown and full stop of the carriage during the final
stages of the test run (also transient stage).

Theoretically, if the time history of a


measured parameter is changing drastically, then the
segments could not be analyzed as identical tests.
The steady state requirement, therefore, calls for a
corrective action to account for the effects of nonuniform ice thickness, non-homogenous ice
mechanical properties and small fluctuations in

Steady State Requirement


In ice tank testing, for any given ice sheet, the
ice properties are not completely (100%) uniform
(same thickness) and homogeneous (same mechanical
properties) all over the ice sheet. This is attributed,
mainly, to the ice growing processes and refrigeration
system in the ice tank. Figure 3 shows examples for
measured spatial variability of ice thickness and
flexural strength of an actual ice sheet in the tank.

carriage speed on the test measurements.


Figure 3b: Measured 3-D profiles of typical spatial
variation of the flexural strength of ice in the tank.

Figure 3a: Measured 3-D profiles of typical spatial


variation of ice thickness in the tank.
In addition to the spatial variability of the
material properties of ice, during an ice test run, the
carriage speed may (or may not) be maintained at
exactly the required nominal constant speed. The
control system maintains the carriage speed constant.
However, when ice breaks, small fluctuations in
carriage speed may take place.
Because of this inherent non-uniformity of ice
sheets, the non-homogeneity of ice properties and the
small fluctuations in the carriage speed, steady state in
the time history of a measurement may not be
achieved. For example, in Figure 1a, the tow force did
not become completely steady after the initial transient
stage.

To identify whether or not the time history for


a measured parameter has reached its steady state, the
following procedure was recommended. The time
histories for the measured parameters, in all ice test
runs, were plotted along with their linear trend lines
(Derradji-Aouat and van Thiel, 2004). A linear trend
line with zero slope (or very close to zero) indicates
that a steady state in a measured parameter is achieved.
After drawing the linear trend lines through
all measured tow forces, it was observed that, in the
majority of cases, a true steady state was never
achieved. For example, the linear trend lines show that
the tow force time histories sloped over a range of
0.002% (ice sheet # 2, Run #3) to 5.2% (ice sheet # 4,
Run #1). Derradji and van Thiel (2004) showed that,
although the slopes of the trend lines varied within
only 5.2%, they led to some significant changes in the
magnitude of tow forces over the 65 m towing
distance. They suggested that the non-steady state
condition may be attributed to one (or all) of the
following 3 factors:
1. A changing carriage speed (or small fluctuations
in carriage speed) during testing,
2. Non-uniform ice thickness,
3. Non-uniform mechanical properties of the ice
(flexural/compressive strengths, elastic modulus,
density of ice, ).

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

The contribution of each factor was


investigated by Derradji Aouat and van Thiel (2004),
and they concluded that the effect of changing carriage
speed can be ignored (that is factor # 1). The effects of
the other two factors are given as:

The value of (ROpen Water) is obtained from the


correlation obtained from the baseline open water test
results (Appendix A).
For a given ice sheet, with nominal thickness
ho, the following equation is used to correct for the
total ice resistance (Derradji-Aouat, 2003):

NON-UNIFORM ICE THICKNESS


Mean ice thickness profiles were calculated,
each mean profile is the average of 3 measured ice
thickness profiles (along the CC, NQP and SQP). Each
profile is a series of ice thickness measurements (every
2 m) along the length of the ice tank.
The linear trend lines, through the mean
profiles, indicate that the ice thickness varied within a
range of 0.69% to 2.64%.
To correct for the effects of non-uniform ice
thickness on the two force measurements, the
following correction methodology and rational were
followed (Derradji-Aouat, 2002):
Note that, in this forgoing discussion, mean
ice resistance values are used to show how the EUA
method is conceptualized and developed. The same
correction methodology are used for maximum ice
resistance values (Derradji-Aouat, 2002). Also, note
that ice thickness corrections are applied only to the
resistance due to the ice. Therefore, the total ice
resistance (RTotal Ice) is equal to the measured resistance
in the ice tests (RMeasured) minus the resistance measured
in the baseline open water tests (ROpen Water).

(R Total Ice )Mean = (R Measured )Mean

(R

Open Water

(RTotalIce)CorrectMean = (RTotalIce)MeasuredMean* ho
hm

(2b)

where (RTotal Ice) Correct Mean is the corrected total ice


resistance, (RTotal Ice) Measured Mean is the measured total
ice resistance, ho is the nominal ice thickness (40 mm),
and the hm is the measured ice thickness at a distance D
(D is the distance in the tank where hm .is measured,
Derradji-Aouat and van Thiel, 2004).
Note that only the results of tests in
continuous ice (Run # 1, # 2 and # 3) are subjected to
ice thickness corrections. In broken ice test results
(Run # 4, # 5 and # 6), no corrections were necessary.
This stems from the fact that, in broken ice tests, the
original ice thickness profiles are not maintained.
The time histories measured in the creeping
speed test runs are also not subjected to corrections for
ice thickness variation. The length of each creeping
speed test run is small (only one ship length 3.8 m),
the variation of ice thickness over this small length can
be ignored.

) (2a)

Table 2: Example for calculations of random uncertainties in mean tow forces


(Ice Sheet # 2, Run # 6, Model speed = 0.2 m/s)

Note: Shaded segments are outliers, as per the Chauvenet criterion.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

NON-HOMOGENEOUS ICE PROPERTIES


Mean flexural strength profiles along the
length of the ice tank were given by Derradji-Aouat
and van Thiel (2004). In each ice sheet, two flexural
strength profiles along the SQP and NQP are measured
every 15 m.
The flexural strength profiles are obtained
using in-situ cantilever beam tests. The beam
dimensions have the proportions of 1:2:5 (thickness,
hf,: width, w: length, L). The flexural strength f is
calculated as:

f =

6 PL
2
wh f

(3a)
where P is the point load.
The uncertainty in the flexural strength is Uf:

U f = U + U + U
2
P

2
L

2
W

+ 2U

2
hf

(3b)

where UL, UW, and Uhf are the uncertainties in the


measured dimensions (L, W and hf). Up is the
uncertainty in the measured point load.
The uncertainties in the flexural strength
profiles are calculated using Eq. 3b. Derradji-Aouat
(2002) reported that any data correction for ice
thickness includes, implicitly, the correction for the
flexural strength of the ice. This is due to the fact that
ice thickness is a fundamental measurement while the
flexural strength is a calculated material property
(flexural strength is calculated from measurements of
applied point load and dimensions of the ice cantilever
beam). Since this work deals with EUA of actual
fundamental measurements, it is recognized that if
corrections were to be made for both ice thickness and
flexural strength, double correction (double counting)
would take place, and the final uncertainty values
would be overestimated. The same argument is valid
for corrections for the comprehensive strength of ice
(the latter is calculated from applied axial load and
measurements of actual dimensions of the ice sample).
Measured ice density profiles along the length
of the ice tank were also given by Derradji-Aouat and
van Thiel (2004). The density of ice, i, is given as:

i = w

M
V

(4a)

where w is the density of water. M is the mass of the


ice sample. The volume, V, is calculated from the
sample dimensions (length, L, width, W, and thickness,
H): The uncertainty in the ice density is:

U i = U 2H + U 2L + U 2W + U 2M

(4b)

The variation of density along the centre line


of the tank was between 4.58% to 8.60%. To a large
extent, this range is a reflection of the uniformity of
non-bubbly ice. From the ice tank operational point of
view, in non-bubbly ice sheets, density value could not
be controlled but its uniformity is reasonably assured.
In bubbly ice, however, the inverse is true, the target
density values can be achieved in pre-set locations in
the tank but its spatial uniformity is compromised.
CALCULATIONS RANDOM UNCERTAINTIES
In the following example, the discussion will
be focused on the mean tow force history (Figures 1a
and 1b). Table 2 shows the segments for the mean tow
force history obtained in ice sheet # 2 for Run #6.
Essentially, the tow force history is divided into 10
segments. Mean tow force (TFMean) is obtained for each
segment.
The mean of the 10 means (Mean_TFMean) and
the standard deviation of the 10 means (STD_TFMean)
were calculated (as shown in Table 2). Random
uncertainties in the tow forces U(TFMean) are calculated
in three steps:
Step # 1: In Table 2, after the calculations of the mean
of means and standard deviation of means, the
Chauvenets criterion is applied to identify outliers
(outliers are discarded data points). The Chauvenet
number for mean tow forces is (Chauv #)Mean:

(Chauv #)Mean =

TFMean

Mean_TFMean)

(5a)

STD_TFMean

For 10 to 15 segments, the Chauv # should


not exceed 1.96 to 2.13. In Table 2, data points with
Chauv # greater than 1.96 were disregarded (shaded
cells). A new mean of means and a new standard
deviation of means are calculated from the remaining
data points (remaining segments).

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

18

Run #1: Mean Tow Force in Level Ice

15

50

Mean Tow Force (N)

Mean Tow Force (N)

60

40
30
20
10

Run #4: Mean Tow Force in 9/10th Ice


Concentration

12
9
6
3

PICE-1

PICE-2

PICE-3

PICE-4

PICE-1

Ice Sheet

35

18

Run #2: Mean Tow Force in Presawn Ice

Mean Tow Force (N)

Mean Tow Force (N)

15

25
20
15
10

PICE-4

Run #5: Mean Tow Force in 8/10th Ice


Concentration

12
9
6
3

PICE-1

PICE-2

Ice Sheet

PICE-3

PICE-1

PICE-4

15

Mean Tow Force (N)

30

20

10

PICE-2

PICE-3

PICE-4

Ice Sheet

Run #3: Mean Tow Force in Unsupported Ice

40
Mean Tow Force (N)

PICE-3

Ice Sheet

30

50

PICE-2

Run #6: Mean Tow Force in 6/10th Ice


Concentration

10

PICE-1

PICE-2

PICE-3

Ice Sheet

PICE-1

PICE-4

PICE-2

PICE-3

PICE-4

Ice Sheet

Figure 4a Measured mean tow forces presented as function of the ice sheet number.

Step # 2:After calculating the new mean of the means


and the new standard deviation of the means (from the
remaining segments - data points), random uncertainty
in the mean tow force is:

U(TFMean ) = t*(STD_TFMean )

(5b)

where t 2, and N is the number of the remaining data


points (valid segments).

Step # 3: Random uncertainties (Eq. 5b), are expressed


in terms of uncertainty percentage (UP):

UP(TF

Mean

U(TF Mean )
) =
*100

Mean_TF Mean

(5c)

Note that the above three steps are also used


to calculate random uncertainties in maximum tow
forces. This is achieved by substituting the subscript
mean by the subscript max in Eqs. 5a, 5b and 5c.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

9%

Run #1: Uncertainty in Level Ice

3%

2%

1%

PICE-1

PICE-2

Ice Sheet

PICE-3

PICE-1

30%

Run #2: Uncertainty in Presawn Ice

PICE-2

PICE-3

Ice Sheet

PICE-4

Run #5: Uncertainty in 8/10th Ice Concentration

25%

6%
5%
4%
3%
2%
1%
0%

20%
15%
10%
5%
0%

PICE-1

6%

3%

PICE-4

U Mean Tow Force (%)

U Mean Tow Force (%)

7%

6%

0%

0%

PICE-2

PICE-3

Ice Sheet

PICE-4

PICE-1

14%

Run #3: Uncertainty in Unsupported Ice

PICE-2

PICE-3

Ice Sheet

PICE-4

Run #6: Uncertainty in 6/10th Ice Concentration

12%

5%
U Mean Tow Force (%)

U Mean Tow Force (%)

Run #4: Uncertainty in 9/10th Ice Concentration

4%
U Mean Tow Force (%)

U Mean Tow Force (%)

5%

4%
3%
2%
1%

10%
8%
6%
4%
2%

0%

0%

PICE-1

PICE-2

PICE-3

PICE-4

Ice Sheet

PICE-1

PICE-2

PICE-3

Ice Sheet

PICE-4

Figure 4b: Uncertainties in mean tow forces presented as function of the ice sheet number.

It is important to note that the above


procedure (segmentation of measured time history,
check for the steady state requirement, correction for
ice thickness, the use of the three calculation steps) is
valid for calculating random uncertainties in all other
measured ship motion parameters (such as pitch,
heave, yaw and sway).
RANDOM EU IN MEAN TOW FORCES
Figures 4a and 4b show the measured mean
tow forces and their random uncertainties (calculated
using the above steps). The figures show that:

In level (continuous, unbroken) ice test runs


(Run # 1, # 2 and # 3), calculated random uncertainties
in mean tow forces are less than 6%. In fact, all
uncertainties were below 4%, except for two data
points (ice sheet #1, Run #2 and ice sheet # 4, Run #3),
where uncertainties values were 5.84% and 4.08%,
respectively.
In broken ice test runs (Run # 4, # 5 and # 6),
all random uncertainties were below 18%, except for
test run # 5 in ice sheet # 1, where the uncertainty
value was 25.94 %. It should be emphasized that in
broken ice tests, no corrections for ice thickness
profiles were made.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

EFFECTS OF DATA REDUCTION EQUATIONS

CONCLUSIONS AND RECOMMENDATIONS

Equation 2b was proposed to correct for the


effects of ice thickness variations on the values of
random uncertainties. It should be recognized that the
corrected resistance curves are not direct laboratory
measurements, but they are calculated from the
analytical Eq. 2b. The process of using analytical
equations to correct measured parameters is called
Application of Data Reduction Equations, DRE.
In EUA, there are additional random
uncertainties involved in the application of the DRE.
The uncertainty involved in using Eq. 2b is:

In continuous ice test runs, the uncertainty


range of 3% to 10% was obtained (in all three phases).
This is consistent with the range of uncertainties
obtained in Phases I and II test programs. This is also
consistent with the previously reported studies (in the
literature) using different ship models, in different ice
tanks, in different countries over a time span of 10 to
12 years (Appendix B).
In broken ice, the uncertainties ranged from to
3% to 26%. This is also consistent with the calculated
range in Phases I and II test programs. In pack ice
conditions, large uncertainties are possible (and
sometimes expected) in randomly broken ice.

(6)

In the above equation, (UR/R) is the total


uncertainty in resistance, R. Both (UR0/R0) and (Uh/h0)
are the relative uncertainty in the measured ice
resistance (as shown in the example in Table 2) and the
relative uncertainty in the measured ice thickness,
respectively. Note that, in Eq. 6, the value of (Uh/h0) is
an additional relative uncertainty, which is induced by
the use and application of the DRE. The total
uncertainty is the geometric sum of both relative
uncertainties (UR0/R0) and (Uh/h0).
COMPARISON OF UNCERTAINTY VALUES IN
CONTINUOUS ICE AND IN PACK ICE TESTS.
In continuous ice (including presawn ice
sheets), random uncertainties were mainly under 6%.
However, in broken ice tests, uncertainties of up to
18% were obtained (except in 2 cases). The value of
18% is higher than in continuous ice (6%). The
difference between the two uncertainties is attributed to
several factors (the details were given by DerradjiAouat, 2002).
COMPARISON OF UNCERTAINTY VALUES IN
MEAN AND IN MAXIMUM TOW FORCES
Figure 5 shows comparisons between random
uncertainties in mean tow forces and those in
maximum tow forces. In general, uncertainties in
maximum tow forces are higher than those in mean
tow forces (ration of 2 to 4). The details were given by
Derradji-Aouat and van Thiel (2004).
Examples for omparisons among the results of
all three phases of testing are given in Figures 6a and
6b.

Comparison between uncertainty in


maximum and mean tow force

30%

Uncertainty in Max Tow Force (%)

2
UR 0 2
UR
Uh

=
+

R
h0
R0

25%

20%

Run #1

Run #2

Run #3

Run #4

Run #5

Run #6

15%

10%

5%

0%
0%

5%

10%

15%

20%

Uncertainty in Mean Tow Force (%)

25%

30%

Figure 5: Comparison of uncertainties in measured


mean tow force with those in maximum tow force.
REFERENCES
ASME PTC 19.1-1998, Test Uncertainty. Supplement
to Performance Test Code, Instruments and
Apparatus.
Coleman H.W. and Steele W.G., Experimentation and
Uncertainty Analysis for Engineers, 2nd edition, John
Wiley & Sons publications, New York, 1998.
Derradji-Aouat A., Moores C. and Stuckless S., Terry
Fox Resistance Tests. The ITTC Experimental
Uncertainty Analysis Initiative, IMD report # TR2002-01.
Derradji-Aouat A., Experimental uncertainty analysis
for ice tank ship resistance experiments, IMD/NRC
report # TR-2002-04.
Derradji-Aouat A. and Coff J., Terry Fox
Resistance Tests Phase II. The ITTC Experimental
Uncertainty Analysis, IMD report # TR-2003-07.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

40
Mean Tow Force (N)

60

Speed = 0.1 m/s : Compare P1 & P3 Mean Tow


Force

Phase 1: Tow Post

30

Phase 3: PMM
20
10

40

Phase 1: Tow Post


Phase 3: PMM

30
20
10

0
Run #1

50

Run #2

Run #3

Run #4

Run #5

Run #6

Run #1

75

Speed = 0.2 m/s: Compare P1 & P3


Mean Tow Force

40
Phase 1: Tow Post

30

Phase 3: PMM
20
10

Mean Tow Force (N)

Mean Tow Force (N)

Speed = 0.4 m/s: Compare P1 & P3


Mean Tow Force

50
Mean Tow Force (N)

50

Run #2

Run #3

Run #4

Run #5

Run #6

Speed = 0.6 m/s: compare P1 & P3


Mean Tow Force

60
Phase 1: Tow Post

45

Phase 3: PMM
30
15
0

0
Run #1

Run #2

Run #3

Run #4

Run #5

Run #6

Run #1

Run #2

Run #3

Run #4

Run #5

Run #6

Figure 6a: Comparison of test results: Measured mean tow forces using the tow post (Phase I) and measured mean
tow forces using the PMM (Phase III).

Derradji-Aouat A. Phase II Experimental uncertainty


analysis for ice tank ship resistance experiments,
IMD/NRC report # TR-2003-09.

APPENDIX A:

Derradji-Aouat A and van Thiel A., Terry Fox


Resistance Tests Phase III (PMM Testing). The
ITTC experimental uncertainty analysis initiative,
IOT/NRC report # TR-2004-05.

In phase III test program, the main


components of the test set up are: The Terry Fox ship
model, the Planar Motion Mechanism (PMM), data
acquisition system (DAS) and video cameras.
This phase of testing required 4 different ice
sheets. All ice sheets were made up of non-bubbly ice,
they all have the same target thickness (40 mm) and the
same target flexural strength (35 kPa). Preparing the
ice tank, seeding and growing the ice sheets, surveying
for the mechanical properties of ice (flexural and
comprehensive strengths, elastic modulus and density)
and ice thickness profiles were performed as per the
IOT internal standards and work procedures, which are
similar to the ITTC recommended procedures.

Derradji-Aouat A.. A Method for Calculations of


Uncertainty in Ice Tank Ship Resistance Testing,
Proceedings of the 19th International Symposium on
Sea Ice, 2004, PP. 196-206, Mombetsu, Japan.
GUM., General Uncertainty Measurements, 2002
(http://www.gum.dk/)
ISO., Guide to the Expression of Uncertainty in
Measurements, 1995, .ISBN 92-67-10188

TEST PROGRAM AND TEST RESULTS

Two types of experiments were performed:


Experiments in ice and experiments in open water.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

35%
30%

20%

PICE-1: Compare P1 & P3 Uncertainty


in Mean Tow Force at 0.1m/s

PICE-3: Compare P1 & P3 Uncertainty


in Mean Tow Force at 0.4m/s

U(%) Mean Tow Force

U(%) Mean Tow Force

16%
25%

Phase 1: Tow Post


Phase 3: PMM

20%
15%
10%

Phase 1: Tow Post

12%

Phase 3: PMM

8%

4%
5%
0%

0%

Run #1

25%

Run #2

Run #3

Run #4

Run #5

Run #1

Run #6

14%

PICE-2: Compare P1 & P3 Uncertainty


in Mean Tow Force at 0.2m/s

Run #2

Run #3

Run #4

Run #5

Run #6

PICE-4: Compare P1 & P3 Uncertainty


in Mean Tow Force at 0.6m/s

12%

U(%) Mean Tow Force

U(%) Mean Tow Force

20%
10%

15%

Phase 1: Tow Post


Phase 3: PMM

10%

Phase 1: Tow Post


Phase 3: PMM

8%
6%
4%

5%
2%
0%

0%

Run #1

Run #2

Run #3

Run #4

Run #5

Run #6

Run #1

Run #2

Run #3

Run #4

Run #5

Run #6

Figure 6b: Comparison of uncertainties in mean tow force using the tow post (Phase I) and the PMM (Phase III).
The experiments in ice involved:
1.a:
Experiments in level ice sheets (unbroken).
1.b:
Experiments in pre-sawn ice sheets.
1.c:
Experiments in pack ice (broken)
The experiments in open water involved:
2.a:
Standard resistance experiments in open water
(as per the ITTC procedure)
2.b:
Baseline experiments in open water (constant
speed through the length of the tank)

Figure A.1b: Terry Fox model on the shop floor


(model is in its storage wooden cradle). Length at
water level is 3.74 m, Maximum beam at water level is
0.79 m, Model scale is 1/20.8
Figure A.1a: Terry Fox model on the calibration frame
in the fabrication shop:

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

EXPERIMENTS IN ICE
Ship model speeds of 0.1 m/s, 0.2 m/s, 0.4
m/s, and 0.6 m/s were selected. Each ice sheet was
tested for only one speed. For example, ice sheet #1
was tested for speed of 0.1 m/s, ice sheet # 2 was tested
for speed 0.2 m/s. In each ice sheet, six (6) different
test runs were performed. The first three runs were
conducted in level and pre- sawn ice sheets, while the
last three runs were conducted in pack broken ice. A
total of 24 resistance test runs in ice were performed.
A schematics for the six (6) ice test runs is
shown in Figure A.2. The first three test runs are:

Run # 1: This run was performed in level ice,


along the centerline of the tank (the central
channel, CC). The carriage speed was kept
constant along the entire useable length of the ice
tank (~ 65 m). Note that after the completion of
this run, an open water channel along the
centerline of the tank is created.
Run # 2: The model was moved to the South
Quarter Point (SQP) of the tank. The south half of
the ice sheet was constrained (using pegs), and the
ice was pre-sawn along the SQP straight path.
Resistance test run was performed in the pre-sawn
ice at constant speed (same speed as Run # 1).
Run # 3: The model was moved to the North
Quarter Point (NQP) of the tank. The north half of
the ice sheet was neither pre-sawn nor constrained.
Resistance test run was performed in the ice sheet
at constant speed (same speed as Run # 1).

Runs # 4, # 5, and # 6 were performed in pack


(broken) ice. After the completion of the first 3 runs,
the ice sheet was broken (manually) into small blocks
(about 300 mm X 300 mm). Then, the blocks were
redistributed in the tank, manually, to achieve the
desired pack ice concentration. Three (3) different ice
concentrations were targeted. These are:

12 m
NQP

CC

SQP

76 m

Free Boundary
Tests Run # 3

Test Run # 1

Constrained
Boundary
Pre-Sawn Ice
Test Run # 2

Figure A.2: A schematic for Run # 1, # 2 and # 3

Run # 4: Test run in 9/10ths ice concentration, tow


along the NQP at a constant speed.
Run # 5: Test run in 8/10ths ice concentration, tow
along the CC at a constant speed.
Run # 6: Test run in 6/10ths ice concentration, tow
along the SQP at a constant speed.

These ice concentrations were chosen to reflect actual


pack ice environment. Ice concentration less than about
6/10 yields a behaviour equivalent to that of baseline
open water tests. Figure A.3 is

Figure A.3: Typical test run in broken ice (Phase III)

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Tow Force versus Velocity:


Phase 1, 2 & 3 Open Water Baseline Resistance Tests

Level Ice (Run #1)


Pre-sawn Ice (Run #2)
Unsupported Ice (Run #3)

50

y = 19.091x + 1.3155x + 0.0388


2

R = 0.9959

40

Tow Force (N)

Tow Force (N)

Phase III: Tow Force vs. Velocity in Continuous Ice

60

5
4
3

30

20

2
10

Velocity (m/s)
0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

Figure A.4a: Baseline open water tests (best fit for all
three phases of testing).

90

8/10th ice concentration (Run #5)

0.3

0.4

0.5

0.6

0.7

Tow Force versus Velocity:


Phases 1, 2 & 3 Open Water Standard Resistance Tests

70

6/10th ice concentration (Run #6)

y = 32.21x2 - 11.354x
R2 = 0.9735

60

12

Tow Force (N)

Tow Force (N)

0.2

80

9/10th ice concentration (Run #4)


16

0.1

Figure A.5a: Measured Tow Force in Continuous Ice.

Phase III: Tow Force vs. Velocity in Broken Ice

20

Velocity (m/s)

50
40
30

20

Velocity (m/s)

0
0

0.1

0.2

0.3

0.4

10
0.5

0.6

0.7

Velocity (m/s)

0
0

Figure A.4b: Standard open water tests (best fit for all
three phases of testing).
EXPERIMENTS IN OPEN WATER
Standard Resistance Experiments in Open Water: A
series of standard resistance tests in open water were
performed in the ice tank (calm water). In all tests,
turbulent stimulation studs were placed on the model
and beach absorbers were used. Ship speeds from 0.1
m/s to 1.8 m/s were covered (increments of 0.2 m/s).
Baseline Resistance Experiments in Open Water: In
this series of tests, the turbulent studs and beach
absorbers were removed. In each test, model was
towed in calm open water at constant velocity along
the entire useable length of the ice tank (~ 65 m).
Velocities of 0.1 m/s, 0.2 m/s, 0.4 m/s and 0.6 m/s
were tested.
TEST RESULTS
Figures A.4a and A.4b show the results from
the standard open water resistance tests and the results
from the baseline open water tests, respectively.

0.2

0.4

0.6

0.8

1.2

1.4

1.6

1.8

Figure A5b: Measured Tow Force in Broken Ice


Figures A.5a and A.5b give the results for the
test runs in ice. Best fit curves through the data are
provided. As expected, the results are typical of
resistance data of a ship such as the Terry Fox in ice.
Detailed presentation of the test program and test
results were given in an internal report by DerradjiAouat and van Thiel (2004). The results for the surveys
for the flexural and compressive strengths, elastic
modulus and ice density were also included in the
report.
APPENDIX B:
BASIC FORMULATION
In a typical experiment, the total uncertainty,
U is the sum of a bias uncertainty, B and a random
uncertainty, P. Bias uncertainties (also called
systematic uncertainties) are due to uncertainty sources
such as load cell calibrations, accuracy of motion
instrumentations and DAS. Random uncertainties (also
called precision or repeatability uncertainties) are a

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

measure of the degree of repeatability in the test results


(i.e. if a test was to be repeated several times, would
the same results be obtained each time?). Examples for
random uncertainty sources are the changing test
environment (such as fluctuations in room temperature
during testing), small misalignments in the initial test
setup, human factors, etc.
Mathematically, the total uncertainty is:

(B

U =

P2

(B.1)

For a single test population (where only one


test is performed, and for that one test, N data readings
are obtained), random uncertainty P from an
uncertainty source X is Px:
(B.2a)
P = t *S
X

The coefficient t is obtained from the


standard table for a normal Gaussian distribution. Its
value depends on the desired level of confidence
(usually, 90% to 95%) and the number of the Degree
of Freedom (DOF) in the sample population. The DOF
= N 1, where N is the numbers of data readings. For
DOF 10, t 2.
In a multi-test population (where the same test
is repeated N times, and each test is represented by
only one data point in the population of N data points),
the random uncertainty from an uncertainty source X
is PNX:

P NX

t * S

NX

(B.2b)

Derradji-Aouat (2002) showed that in a


typical ice tank ship resistance test, the bias uncertainty
component (B) is much smaller than the random one
(P), he reported that, in Phase I ship model tests in ice,
the value of (B) is at least one order of magnitude
smaller than the value of (P). He concluded, therefore,
that; in routine ship resistance ice tank testing, the total
uncertainty (U) can be taken as equal to the random
one. Simply, without a loss of accuracy, the bias
uncertainty component can be neglected. It follows
that:

REVIEW - EUA IN ICE TANK TESTING


Newbury (1992) conducted several standard
ship resistance tests in ice, at the NRC/IMD ice tank,
using a model fo
r R-class icebreaker. He used nine (9) ice
sheets to execute the test program. Of the 9 ice sheets,
five (5) ice sheets had nominal thickness of about 35
mm and flexural strength of about 40 kPa, two (2) ice
sheets had nominal thickness of about 50 mm and
flexural strength of about 40 kPa, and two (2) ice
sheets had nominal thickness of about 22.5 mm and
flexural strengths of about 40 kPa and 20 kPa,
respectively. The tests were conducted for various
model speeds (ranging from 0.15 m/s to 0.9 m/s).
In each test, Newbury (1992) calculated the
mean and standard deviation for ice breaking
resistance. In most tests, his calculations showed that
the ratio of the standard deviation to the mean
resistance was within 5%; expect for 2 cases, where the
ratios were 8.7% and 11.8 %, respectively. If
uncertainties were to be calculated using Eq. B.2,
Newburys (1992) test results show that random
uncertainties in ice breaking resistance would be about
10% to 12%. In the extreme two cases, random
uncertainties will be about 18 % to 22%. Note that
random uncertainty range of 10% to 12% is, more or
less, supported and confirmed by Kitagawa et al. (1991
and 1993).
Kitagawa et al. (1991 and 1993) published ice
breaking resistance data obtained in the ice tank of
National Maritime Research Institute of Japan (NMRI).
Total uncertainties of about 10% were calculated (in
some cases, uncertainties of about 20% were reported).

(B.3)
The above equations are valid for direct
measurements (directly measured variables, such as
load, deformation, motion, pitch, roll, etc.). In most
cases, the measured variables are used to compute
engineering parameters (such as stress, strain,
resistance, etc.) using Data Reduction Equations
(DRE). Additional uncertainties due to the use of DRE
need to be considered (as discussed in the paper).

Figure A6: Actual ice braking operations off the coast


of Newfoundland and Labrador, Canada.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics


St. Johns, Newfoundland and Labrador, CANADA, 8-13 August 2004

Preliminary Modeling of Ship Maneuvering in Ice

Michael Lau and Ahmed Derradji-Aouat


(Institute for Ocean Technology, National Research Council of Canada, Canada)
ABSTRACT
Recent developments of offshore oil and gas
reserves in several countries, together with economic
studies to increase transportation through the Arctic,
has led to a renewed interest in the maneuverability
of vessels in ice. Despite the sizeable volume of
work, there is not yet a universally accepted
analytical method of predicting ship performance in
ice. The Institute for Ocean Technology (IOT) of the
National
Research
Council
of
Canada
(http://www.iot-ito.nrc-cnrc.gc.ca/) has conducted
physical, numerical and mathematical modeling of
ship maneuvering characteristics in ice, as part of a
larger effort to develop reliable modeling techniques
to assist in the design of new ice-worthy vessels and
in the simulation of their navigating characteristics.
In this paper, the preliminary results of a
series of physical and mathematical modeling are
presented. As the yaw moment and turning radius are
the important indicators of the maneuvering
performance, this paper will focus on the interaction
processes and the influence of ship motions on the
yaw moment exerted on the ship hull. The dominant
ice-ship interaction processes are identified. The
results show large influence of ship motions and
interaction geometry on the measured yaw moments.
The geometrical aspect of the interaction processes is
described and its influences on ice loads are
discussed.
Conclusions
are
made
and
recommendations for future works are provided.

Council
of
Canada
(http://www.iot-ito.nrccnrc.gc.ca/) has conducted physical, numerical and
mathematical modeling of ship maneuvering
characteristics in ice, as part of a larger effort to
develop reliable modeling techniques to assist in the
design of new ice-worthy vessels and in the
simulation of their navigating characteristics. The
objective is to develop a physical representation of
the complex interaction processes of a ship
maneuvering in ice and to build a mathematical
model to satisfactorily predict its performance. In
turn, the mathematical model will provide a tool for
ship designers to use as part of the assessment of ship
navigation in ice infested routes. It can also be
incorporated into marine simulators to train mariners,
or into automatic ship control systems for better ship
maneuvering.
In this paper, the preliminary results of a
series of physical and mathematical modeling are
presented. As the yaw moment and turning radius are
the important indicators of the maneuvering
performance, this paper focuses on the interaction
processes and the influence of ship motions on the
yaw moment exerted on the ship hull. The dominant
ice-ship interaction processes are identified. The test
results show large influence of ship motions and
interaction geometry on the measured yaw moments.
The geometrical aspect of the interaction processes is
described and its influences on ice loads are
discussed.
Conclusions
are
made
and
recommendations for future works are provided.

INTRODUCTION

MODEL TESTS

Recent developments of offshore oil and gas


reserves in several countries, together with economic
studies to increase transportation through the Arctic,
has led to a renewed interest in the maneuverability
of vessels in ice. Past experiences with icebreakers
have shown that the maneuverability of a ship can be
improved by modifying specific features of the hull,
the propulsion system and by using maneuvering aids
- such as a thruster, or a bubbler.
Despite the sizeable volume of work, there
is not yet a universally accepted analytical method of
predicting ship performance in ice. The Institute for
Ocean Technology (IOT) of the National Research

Ship maneuvering in various ice conditions


is a complex subject. Our present understanding of
the nature of ship-ice interactions is still limited.
Considering the complexity of the loads imposed by
ice during ship maneuvers, a preliminary series of
ship maneuvering experiments in ice were conducted
to provide insights to assist in the subsequent
numerical and mathematical modeling. These
experiments were carried out with a 1:21.8 scale
model of the Canadian Coast Guards icebreaker,
Terry Fox (IOT Model # 417). The test conditions for
these experiments are given in Table 1. The model
was mounted to the towing carriage through a PMM

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

(Planar Motion Mechanism, Spencer and Harris,


1997) at the models center of gravity (see Figure 1).
The model was towed at a controlled planar motion
through a level ice sheet. In each experiment, tow
forces, turning moments, and ship motions were
measured. The transducers outputs were sampled
digitally at 50 Hz, and filtered at 200 Hz. The model
surface was finished to a friction coefficient of 0.01
with Duponts Imron paint. Two video recordings
were made of each test: one on the starboard side manually controlled to follow the models
maneuvers, and the other looking down ahead of the
model at the port side.
The PMM apparatus consists of two primary
components: a sway sub-carriage that is mounted
beneath the main towing carriage, and a yaw
assembly that is connected to the sway sub-carriage.
The apparatus allows the model to yaw and sway in a
controlled manner, while measuring the sway and
surge forces as well as the yaw moment. The
combination of sway and yaw allows a variety of
maneuvers to be performed. The specifications for
the PMM are given in Table 2.
The test matrix for the experimental
program is summarized in Table 3. For the tests
described in this program, the ice sheets had a target
ice thickness of 40 mm and a target flexural strength
of 35 kPa. The following maneuvers were utilized:
(1) static drift, in which the model was towed along a
straight line at a fixed yaw angle, and (2) pure yaw
through a constant radius maneuver so that the
heading of the model was always tangential to the
path of its center of gravity resulting in zero sway
force and a yaw moment. The static drift was
performed with three drift angles (-20 to 20), and the
constant radius maneuver was conducted with two
turning radii (50 m and 10 m). All tests were
conducted with model velocity ranging from 0.02 m/s
to 0.6 m/s. Concurrent to the testing in ice maneuvers
in open water were also conducted.
The experiments were carried out in CDEG/AD/S ice (Spencer and Timco, 1990). With
inclusions of air bubbles into the growing ice sheet,
the model ice significantly improves scaling of ice
density, elastic and fracture properties. The useable
model ice area is 12 m x 76 m. For each ice sheet,
flexural, compressive and shear strengths were
measured frequently throughout the test period.
Strength versus time curves were created for each ice
sheet and the strength values reported at each test
time were interpolated from these curves. Flexural
strength, f , was measured using in-situ cantilever
beams. The shear strength measurement was
performed immediately after the flexural strength test

to provide index values for comparison with the


measured flexural strengths. The ratio of shear
strength to downward breaking flexural strength
varied from 1.03 to 3.16. The reported ice thickness,
h, is the average thickness of approximately 65
measurements of the ice sheet thickness along the test
path. The IOT standards and work procedures were
followed for producing and characterizing level ice
sheets (IOT Standard GM-4, 1999).
TEST RESULTS AND DISCUSSIONS
Icebreaking pattern and channel width
As the ship advances into an unbroken ice
field, individual cusps or wedges begin to break off
from the level ice at the point of contact at the bow.
These broken cusps and wedges are, then, rotated
downward, pushed farther down the hull, and
eventually cleared away from the hull at the sides.
Once the rest of the level ice sheet contacts the hull,
the same breaking process continues. This sequential
icebreaking creates a channel wide enough for the
passage of the hull. Figure 2 shows an idealization of
the channel created by the hull. The breaking initiated
at the bow creates an initial channel width, Wi,
slightly wider than the ships beam, i.e., the initial
channel width was about one ships beam plus 0.4
times the ice characteristic length, l c (Lau et al,
1999). For a tighter turn, further ice breaking at the
leeward side of the hull may be necessary to create a
final channel width, W f , wide enough for its
passage. The location of ice-ship contacts, and hence
the local icebreaking load, can be estimated by
considering the geometry of the interaction during
turning. The zones for possible contact at different
parts of the ship can be defined by a number of
concentric circles. For a typical ship, the zone of
possible ice contact for the outer side is always larger
than the inner side, and W f is greatly dependent on
the turning radius, as shown in Figure 3. In Figure 3,
the theoretical and measured W f are plotted against
the turning radius showing agreement between theory
and test data. For the 10 m radius turn, the measured
channel width is 4% greater than the theoretical width
and for the 50 m radius turn, 6% smaller.
The effect of ship turning on yawing moment
The results for the turning circle runs are
given in Figure 4. Two turning circle radii of 50 m
and 10 m were tested, each with the velocity ranging

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

from 0.02 m/s to 0.6 m/s. These velocities


corresponded with a yaw rate ranging from 0.02
deg/s to 3.4 deg/s. The data has been corrected to
correspond with 35 kPa ice flexural strength. The
data shows a bilinear relationship between yaw
moment and yaw rate, with a moment offset at 12.7
Nm and 43 Nm for the 50 m and 10 m turning
maneuvers, respectively. The change in slope occurs
at a yaw rate of approximately -0.5 deg/s.
Preliminary analysis was performed to
understand the observed trend. It is believed that the
moment offsets were mainly contributed by velocity
independent ice breaking and submergence
components, and the initial slope was determined by
velocity dependent ice clearing and the open water
components. At this stage, the explanation for the
secondary slope needs further analyses. In the
following analysis, the discussion will be focused on
the initial moment offsets, and the effect of the
interaction geometry on their values.
MATHEMATICAL MODELING
In the following section, a conceptual model
of turning moment imposed on the ship hull during a
steady turn is presented. This model forms the
framework for future mathematical development.
Analogous to ice resistance (Spencer, 1992), the
expression for total yaw moment, Ntot, is divided into
the hydrodynamic, N ow , icebreaking, N br , ice
submergence (buoyancy component),
clearing,

N b , and ice

N cl , components:
N tot = N br + N cl + N b + N ow

Breaking component, Nbr


For the case of a semi-infinite beam on an
elastic foundation end loaded by a concentrated
transverse force, the maximum vertical load per unit
width, PVm , and the associated end deflection, y m ,
are given as follows:

PVm = 0.68 f ( w h 5 / E )1 / 4
ym =
Where

2 PV m
l c w

is the flexural strength of ice,

(3)

is the

specific weight of water, h is the thickness of ice,


is the Youngs modulus of ice and

l c is the

characteristic length of the ice beam.


Assuming an idealization of the ice breaking
force-displacement history, as shown in Figure 5. The
maximum displacement of the ship forward, a ,
before ice failure is related to

ym :

a = y m / tan
where

(4)

is the stem angle.

If the average breaking length, la, is taken as


0.2lc (Lau et al, 1999), then the average vertical force
per unit width, PVa, acting on the ship surface, where
ice breaking occurs, is equal to:

PVa = PVm

(1)

The fundamental reason for this approach is


that different components may not all scale in the
same manner to full-scale. The ice breaking term has
ice strength and thickness as parameters, and takes
into account the effect of channel width and
interaction geometry on the zone of application of the
ice forces. The ice submergence term calculates the
buoyancy forces. These two components are
insensitive to model speed, and hence contribute to
the moment offset at zero ship speed. The ice
clearing and the open water terms include ice added
mass and inertial contribution, and hence are velocity
dependent.
This section presents a simple analysis of
the ice breaking and buoyancy components to
illustrate the importance of ship/ice interaction
geometry during ship maneuvering.

(2)

2f h
a
= 5.7
2 * 0.2l c
E tan

(5)

For the case of ship maneuvering at a


constant yaw rate with a radius, R , as shown in
Figure 6, lets further neglect the frictional
component and assume the energy required for ice
breaking is proportional to the volume of broken ice
created. If the effects of the broken ice pieces sizes
are neglected, then the ice will contact the bow and
the half side of the hull with the three horizontal
loads, F1h , F2 h , F3 h , which can be computed as
follows:

Fh 1 = PVa (l1 l 2 ) tan


Fh 2 = PVa (l 2 l 3 ) tan
Fh 3 = PVa (l 3 l 4 ) tan

Copyright National Academy of Sciences. All rights reserved.

(6)
(7)
(8)

Twenty-Fifth Symposium on Naval Hydrodynamics

Where

l1 , l 2 , l 3 and l 4 are the geometric lengths as

shown in Figure 6, and the angles and are the


angles between the normal of the bow and hull side
surfaces and the vertical line, respectively. Hence, the
yaw moment due to ice breaking from the forward
part is given as follows:

N br = ( Fh1 + Fh 2 )l 5 + Fh3 l 6
where

(9)

l 5 and l 6 are the lengths between the

respective force centers to the ships mass center.


We used a 2-dimensional beam-bending
model, in which the structure was regarded as having
an infinite width. The edge effects should be
considered when calculating ice forces. Modification
to the above formulation was implemented by
considering the ice breaking width adjustments, l1
and l 2 , as shown in Figure 7. By assuming the
following
proportionality
from
geometric
consideration:

l1 l1 l 2
=
l 2 l 2 l 3

(10)

The total width, Wb, of ice broken by the bow is equal


to:

Wb = l1 + l 2 + l1 l 3

(11)

Submergence component, Nb
The buoyancy force on the hull was
calculated by considering the amount of ice covering
the different parts of the hull. For the bow part, as
shown in Figure 8, the vertical components, Fv _ 1
and

Fv _ 2 , of the buoyant forces acting at the

respective side of the bow can be calculated using the


following equations:

l1 l 2
( w i )hS
l1 l 3
l l
= 2 3 ( w i )hS
l1 l3

Fv _ 1 =
Fv _ 2

(12)

where

is the specific weight of ice and

S is the

horizontal projection of the bow surface.


Ignoring the ice/hull
friction, the
corresponding horizontal forces, Fh _ 1 and Fh _ 2 , on
the respective side of the bow due to buoyancy are
given as follows:

l1 l 2
( w i )hS tan( )
l1 l 3
l l
= 2 3 ( w i )hS tan( )
l1 l 3

Fh _ 1 =

(14)

Fh _ 2

(15)

And the yaw moment,

N b , due to buoyancy forces

from the bow is equal to:

N b = ( Fh _ 1 + Fh _ 2 )l 5

(16)

The lengths l1, l2, l3 and l5 are given in Figure 6.


Similarly, the buoyant forces on other parts of the
wetted surface of the hull can also be calculated.
Comparison with model test data
According to the present model, the
components, N br and N b , are independent of yaw
rate, but greatly influenced by the turning radius, R ,
as shown in Figure 9. As the ship maneuvers in
tighter turns, it needs to break more ice at the inner
side resulting in increasing yaw moment. As shown
in Figure 9, the model predicts a yaw moment of 30.7
Nm and 6.5 Nm for the 10 m and 50 m radii,
respectively. In comparison with the measured values
of 43 Nm and 12.7 Nm, the model under-predicted
the moment offset by 29% and 49% for the 10 m and
50 m tests, respectively.
In the present analysis, the friction force and
the in-planed ice compression were neglected in
order to make the problem simpler. This tends to
underestimate ice load where a steep slope is present,
i.e. at the side hull. When calculating the buoyancy
force, some assumptions were made for the broken
ice motions. All these simplifications may introduce
uncertainties and errors to the predictions.
FURTHER WORK

(13)

At the present, a multifaceted approach


would give a more fruitful result. This includes

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

deriving the form of prediction equation through


analytical means, and performing extensive
numerical simulations for a range of hull forms and
maneuvers. Fitting data to the analytical model will
provide empirical coefficients to a set of semiempirical equations.
The analysis presented in this paper has
been significantly simplified. Efforts are underway to
conduct further physical and numerical modeling and
to refine the mathematical models by considering the
possibility of different failure modes occurring
simultaneously along the hull, ice friction, and a
refined interaction geometry and ice breaking and
clearing processes.

in the development of the mathematical model. Dr.


F.M. Williams provided valuable discussions through
out the project. Their support is gratefully
acknowledged.

CONCLUSIONS

Lau, M., Molgaard, J., and Williams, F.M., An


analysis of ice breaking pattern and ice piece size
around sloping structures, 18th international
conference on offshore mechanics and arctic
engineering, OMAE, 1999.

In this paper, the preliminary results from a


multi-faceted study of ship maneuvering test series
were presented. A simple analysis illustrated the
importance of interaction geometry on the interaction
processes and the resulting yaw moment. Despite the
simplicity of the problem treatment, the analysis gave
a favorable prediction. Future work will include a
refinement of the problem treatment, as well as an
extensive series of numerical and physical
experiments with the aim of developing a
mathematical model to successfully predict a ships
maneuvering performance in various ice conditions.
ACKNOWLEDGEMENTS
The investigations presented in this paper
were partially funded by the Atlantic Innovation
Fund through the Marine Institute, Newfoundland.
J.C. Liu provided assistance during model testing and

REFERENCES
Intera Technology, Inc., 1986, DECICE Theoretical
Manual, Lakewood, Colorado.
IOT Standard GM-4, Environmental Modeling
Ice, Version 2, Institute for Ocean Technology, St.
Johns, Newfoundland, 1999.

Spencer, D.S. and Harris, C., The development and


commissioning of a large amplitude planar motion
mechanism for maneuvering of ships in ice and open
water, Contract Report CR-1997-5, NRC/IOT, St.
Johns, Newfoundland, 1997.
Spencer, D.S. and Timco, G.W., CD model ice A
process to produce correct density (CD) model ice,
Proc.10th International IAHR Symp. on Ice, Vol. 2,
Espoo, Finland, 1990, pp. 745-755.
Spencer, D., A standard method for the conduct and
analysis of ice resistance model tests, 23rd ATTC,
New Orleans, 1992, pp. 301-307.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Table 1: Model Hydrostatics


IOT Model #417, scale 1/21.8, without appendages
Displacement (kg)
665.6
Waterline Length (m)
3.739
Waterline Beam at Mid-Ship (m)
0.789
Draft at Mid-Ship (m)
0.368
Center of Buoyancy Forward of Mid-Ship
-0.07
(m)
Center of Aft Body Buoyancy Forward of
0.594
Mid-Ship (m)
Stem Angle ()
23.27
Waterline Entrance Angle ()
32.15

Wf

Wi

Table 2: Specifications of the PMM


Max Sway Amplitude (m) 4.0
Max Yaw Amplitude () 175
Max Sway Velocity (m/s) 0.70
60.0
Max Yaw Rate (/s)
2200
Max Sway Force (N)
Max Yaw Moment (N-m) 3000

Figure 2: The influence of turning motion on channel


width, showing the ice breaking at the bow and hull.

Channel Width, W f (m)

1.3
Theory
Measurement

1.2

Table 3: Matrix of the test program


Turning Radius, R (m)

Drift Angle, ()

-2 to 2

Model Speed,
Yaw Rate,

V (m/s)

(deg/s)

1.1

50

10
0

0.9

0.02~0.6

0.02~ 0.34

Ice Thickness (mm)

40

Ice Strength (kPa)

35

20
40
60
Turning Radius, R (m)

80

Figure 3: Theoretical and measured channel


width as a function of the turning circle.

Yaw Moment, N (Nm)

120
100
80
60
N o,10 = 43 Nm

40
20

N o,50 = 12.7 Nm

0
0

Figure 1: Terry Fox model attached to the


PMM.

-1

R=10 m
R=50 m
Linear (R=50 m)
Linear (R=10m, Primary Slope)
Linear (R=10m, Secondary Slope)

-2

Yaw Rate, r (deg/s)

-3

Figure 4: Moment versus yaw rate for the Terry


Fox model turning in ice with 10 m and 50 m radii.

Copyright National Academy of Sciences. All rights reserved.

-4

Twenty-Fifth Symposium on Naval Hydrodynamics

V
PVm
Force

l1 l2

V
l2 l3

la
Figure 5: The idealized force displacement
history when ship advancing.

V
Figure 8: Bow geometry, showing amount
of ice sliding on bow surface.

Yaw Moment, Nbr+Nb (Nm)

60

CG

Theory
Measurement

50

Lb

40

F1h

l5
l6

l1 l 2
F2h

F3h

l 4 l1 l3 l
2

30
20

l2 l3

l l

10
0
0

20
40
60
80
Turning Radius, R (m)

100

Figure 9: Predicted moment offset as a function of


turning radii.

O
Figure 6: Geometry of a ship maneuvers at a
constant yaw rate.

l1

Edge effect
Crack

l1 l2

l2

l 2 l3
Edge effect
Figure 7: Edge effect on ice breaking pattern at
the bow.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics


St. Johns, Canada, 8-13 August 2004

Experimental and Numerical Study of Semi-displacement


Mono-hull and Catamaran in calm water and incident waves
C. Lugni1, A. Colagrossi1, M. Landrini1, O.M. Faltinsen2
(1 INSEAN, The Italian Ship Model Basin, Roma - Italy, 2 Centre for Ships and
Ocean Structures, NTNU, Trondheim - Norway)
ABSTRACT
There is a broad variety of high-speed vessels. In the
present work, the attention is focused on semi displacement mono-hulls and catamarans both in calm water
and in incident head sea waves. Experimental and numerical studies have been performed to investigate the
main features of the flow pattern near the bow, along the
vessel and downstream the transom.
In the steady model tests, the Froude number has
been varied in a wide range and the interaction between
the two catamaran demi-hulls was studied by comparing the related flow field with the one generated by a
mono-hull identical to a catamaran demi-hull. The experiments data have been compared with results of linear 3D and nonlinear 2D+t computations. In the monohull case, for ship-length Froude number smaller than
0.6 the former capture the phasing of the wave pattern
and give more reliable quantitative information. At higher speeds, the latter agree better with the measurements. This is due to two factors: the transverse waves
become less relevant and the importance of nonlinearities increases. In the catamaran case, the same trend
is observed but the 2D+t theory gives globally the best
agreement at smaller speeds than for the mono-hull. For
both the models, the linear code is not able to predict
the stern flow correctly at ship-length Froude equal or
greater than about 0.5. This is because, from this value
on, experimentally the transom stern stays dry during
the vessel motion. The 2D+t method was used to analyze the mechanisms driving the transom flow field and
the interaction between the demi-hulls.
In the unsteady experiments, different wave am-

plitudes and vessel speeds have been considered and the


incoming wave frequency was varied in the heave and
pitch resonance range. This study highlighted the effect
of nonlinearities on the mono-hull and catamaran flow
patterns and showed the relevance of the interaction between steady and unsteady flows on the wave induced
body motions. Such investigation was performed by
combining in a synergic manner the experimental data
with 3D linear computations. For the considered speeds
and within the resonance range, the measured mean trim
and sinkage of the catamaran are mainly governed by
the steady flow.
INTRODUCTION
The design features of high-speed vessels in use can be
very different. The vessel weight can be supported by
submerged hulls, hydrofoils, air cushion, or a combination of these effects. Cavitation and ventilation on foils,
struts and propulsors limit the speed. Mono-hulls and
catamarans, often equipped with foils, trim tabs and/or
interceptors to minimize wave induced motions, represent nowadays the most popular concepts. Catamaran
designs include the wave-piercing and semi SWATH
(Small Water-plane Area Twin Hull) style hulls. A monohull with the same displacement as a catamaran is characterized by a lower wave induced vertical acceleration
since its larger length is beneficial from this point of
view. The beam-to-draught ratio B/D of high-speed
mono-hulls may vary from around 5 to values larger
than 7. Large B/D values result in more limited accelerations in heave and pitch motions, Faltinsen (1990).
However the roll motions of monohulls need special attention. A SWATH vessel has higher heave and pitch

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

natural periods and generally lower vertical excitation


loads than a similarly sized catamaran. On the other
hand, if control surfaces are not introduced, a threshold Froude number exists beyond which the SWATH is
dynamically unstable in the vertical plane. When operating in head-sea conditions, its seakeeping behaviour is
better than the one of a corresponding catamaran. However, if the sea state, speed and heading cause resonant
vertical motions, the SWATH may not have a good seakeeping behaviour. Surface Effect Ships (SES) use an
air cushion mechanism to obtain the desired cruise velocity and performances. The excess pressure in the air
cushion between the two SES hulls lifts the vessel and
carries about 80% of its weight. On the other hand, it
reduces the metacentric height and causes wave generation and additional wave resistance. However, the total
calm water resistance is smaller than the one of a catamaran of similar dimensions. A drawback with an SES
is that it can suffer more speed loss in waves than for
instance a catamaran. Further, the skirts in the bow of
an SES are easily wom out. Trimarans and pentamarans
with a large center hull combined with smaller outrigger
hulls represent interesting new concepts.
In the case of mono-hulls and catamarans, comprehensive experimental studies have been performed
by Molland et al. (1995) in calm water conditions. The
model tests showed that the forward speed effect on
trim and sinkage starts to be significant at a ship-length
Froude around 0.35. This has important consequences
on the vessel performances since trim and sinkage are
directly connected with the resistance and matter also
for the wetdeck slamming on multi-hull vehicles. Doctors (2003) conducted calm water tests on five different
transom-stern mono-hull models to investigate the transition from wet to dry transom. A dry transom is generally beneficial and affects the trim and rise as well as
the damping of vertical ship motions. The author identified the transom-draft Froude number as the parameter
governing such transition. Keuning (1988) performed
high-speed mono-hull experiments in calm water and
head sea waves and analyzed the wave elevation near
the vessel and the variation of the hydrodynamic coefficients along the ship hull. Forced heave and pitch
model tests on a high-speed catamaran were presented
in Ohkusu and Faltinsen (1991). A weak interaction between the demi-hulls was suggested by the fairly good
agreement of the measured hydrodynamic forces with
numerical 2D+t results based on the assumption of hy-

drodynamically independent demi-hulls.


The numerical studies of wave resistance and
wave induced ship motions on mono-hulls and catamarans are mainly based on the assumptions of incompressible fluid in irrotational motion. The solution of
the more general unsteady Navier-Stokes Equations is
still in its infancy and the use of Reynolds Averaged
Navier-Stokes equations would lead anyway to uncertainties in the turbulence modeling. Linear wave resistance analyses based on the Michells thin-ship theory
(Michell 1898) and accounting for the transom effect
have been presented by different researchers (i.e. Molland et al. 1994, and Doctors and Day 1997). The developed methods have been applied to high-speed monoand multi-hull vessels. In the latter case, the diffraction of the waves generated by one demi-hull due to the
presence of the other demi-hull is not accounted for, i.e.
the waves generated by the separated hulls are simply
superimposed. This type of analysis is very efficient
and important at a pre-design stage. However, the role
played by the nonlinearities becomes more and more
relevant as the Froude number increases. The same is
true for the interaction between steady and unsteady
flows when the seakeeping is of concern. In principle
three-dimensional effects should be accounted for. Several numerical codes exist handling them properly for
the linear seakeeping case. However, the solution of
the fully nonlinear three-dimensional problem is challenging from the numerical and CPU time requirements
points of views. In terms of nonlinear effects, simplifications are often made, for instance, by including
nonlinearities associated with hydrostatic and FroudeKriloff loads. In terms of three-dimensional effects,
rather accurate approximate solutions may be obtained
by using physical considerations. This is based on that
divergent waves dominate for a semi displacement vessel in deep water. The main 3D effects can therefore
be adequately captured by using a 2D+t (also referred
to as 2.5D) theory. In this case the longitudinal flow
variations are assumed smaller than the transverse ones,
leading to a sequence of 2D problems in the ship crossplanes. The 3D information travels one way only, that
is the solution in each cross-plane is influenced just by
the flow upstream. Faltinsen (2001) pointed out the relevance of 3D flow effects in the close vicinity of both
the bow and the transom stern. These aspects represent
a limitation for the 2D+t theory since this one assumes
that both velocity potential and free-surface elevation

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

are zero at the bow and it is not aware of what happens


downstream. The latter implies a predicted pressure at
the transom stern different from the atmospheric value.
These considerations are also relevant in unsteady flow
conditions. Fontaine et al. (2000) accounted for the
bow wave elevation by combining a 3D bow model with
the 2.5D theory. A domain decomposition method (see
i.e. Greco et al. 2002) can be introduced to couple
a global 2D+t modeling with a local 3D analysis for
the description of the transom-stern flow. The same
strategy can also be applied to handle other phenomena characterized by three-dimensional flow features,
i.e. the water on deck and the wetdeck slamming. If
the water-hull interaction is not characterized by particularly small angles between the impacting free surface
and the hull, slamming loads can be modeled as an integrated part of the analysis. Otherwise local hydroelasticity will matter. In this case it is not practical to
treat the slamming phenomenon within the global analysis due to the very different time scales involved. Obviously, the global elastic effect of the slamming must
be included in the global analysis to describe properly
the occurrence of whipping. Springing, i.e. steady-state
wave induced global resonant hydroelastic vibrations,
may be a relevant fatigue issue for larger high-speed
vessels. Since the natural frequencies are high, head sea
are of major concern in this context. Springing may be
excited by linear and nonlinear wave effects. The linear wave excitation corresponds to small incident wave
lengths relative to the ship length. The spatial oscillations of head-sea waves cause strong variations of the
flow in the longitudinal ship direction. As a result, in
this case a 2D+t theory is not suitable.
Still many mechanisms are not fully understood
and quantified. For instance the relevance of nonlinear
effects including the interactions of the steady and unsteady flows. The study of such aspects is more challenging for the catamarans where the interaction between the demi-hulls represents an additional factor entering the problem. Present research work aims to a
more clear picture of the flow features and of the seakeeping properties of these vehicles. Consequently a
dedicated experimental investigation of the effect of the
vessels forward speed has been carried out both in calm
water and in incident head sea waves by using a monohull and a catamaran model. Several Froude numbers,
wave amplitudes and wave frequencies have been selected. The mono-hull geometry is used to analyze the

interaction between the demi-hulls of the catamaran.


Therefore it has been shaped identically to a demi-hull,
implying a much smaller beam-to-draught ratio than the
one characterizing the usual mono-hull high speed vessels. Detailed measurements of the steady and unsteady
wave-field features have been made both along the hull
and downstream the transom stern. Also the flow area
between the two demi-hulls of the catamaran was studied. The experiments have been compared with nonlinear 2D+t and linear 3D Rankine Panel Method computations. The results confirmed the relevance of a nonlinear flow description and the validity of a parabolized
approach in case of sufficiently large Froude numbers.
In the next two sections the mathematical models used
for the numerical computations are described. Section
3 deals with the experimental set-up. A numerical and
experimental study for monohulls and catamarans in
steady conditions is then reported in Section 4. Finally
in section 5 the unsteady motion of the catamaran in
waves is considered.
1

2D+t THEORY

The 2D+t theory leads to a sequence of 2D problems to


be solved in the transverse cross-sectional planes of the
vessel. The 3D effects are partially accounted for since
the generic cross-section is influenced by the flow in the
upstream cross-sections of the hull. The nonlinearities
of the problem are retained. For vertical ship motions
and steady symmetric flows (straight coarse), this approach is suitable at sufficiently large Froude numbers,
let us say larger than 0.4, so that the ship transverse
wave system is not dominating. For horizontal wave induced motions and steady antisymmetric flows (maneuvering), the transverse waves are less important also at
small Froude numbers, i.e. the 2D+t theory can be used
for Froude number smaller than 0.4. Faltinsen and Zhao
(1991) used the 2.5D method to study the ship motions
of high speed mono-hulls. Nonlinear steady and linear unsteady analyses were considered and the interaction between steady and unsteady flows was accounted
for. Maruo and Song (1994) retained also the nonlinearities in the unsteady problem but assumed linear incident waves. In our case, the 2D+t theory is applied to
investigate the steady flow patterns for mono-hulls and
catamarans. In the latter case the interaction between
the demi-hulls is accounted for. No correction of the
local flows at the bow and at the transom stern is intro-

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

duced. Damping foils, trim tab and interceptors are not


modeled. In the following the solution method is briefly
outlined under the assumptions of inviscid steady flow
induced by a ship with constant forward speed and fixed
trim and sinkage.

ze z y
ey

Free Surface(x)
Lagrangian markers

exx
Hull Section (x)

dx/dt = u , p =const

Hull Section (x + U t)

2D

Figure 1: Qualitative sketch of the 2D+t approximation for the steady three-dimensional flow around a ship
with constant forward speed U . Left: 3D ship problem.
Right: equivalent unsteady 2D problem (2D+t).
Let us consider a ship moving with constant velocity U (see left sketch in Figure 1). We assume a
beam-to-draft ratio B/D O(1) and both B and D
individually much smaller than the ship length L, say
= B/L, D/L  1. We also assume the Froude num
ber F r = U/ gL = O(1). In a ship-fixed frame
of reference, the hull geometry is given implicitly as
H(x, y, z) = 0 and the a priori unknown free surface
can be represented as W(x, y, z) = z (x, y) = 0.
Assuming that the fluid is in irrotational motion, the
flow field is described by the Laplace equation combined with the kinematic hull (H) boundary condition
H = 0

(1)

and the kinematic and dynamic free surface (W) boundary conditions
W = 0

and

p=0

(2)

respectively. The statement of the problem is completed


by the upstream radiation condition
U x

as x .

(3)

It is convenient to formulate the problem in terms of the


perturbation potential , linked to by = U x +
. The longitudinal gradient /x can be neglected
with respect to those in the transverse plane. Therefore, the problem for reduces to a sequence of twodimensional problems in the ship cross planes. The related boundary conditions imply the generic cross plane
2D is influenced by the upstream solution and unaware
of the following cross sections. The problem sequence
can be solved once given the conditions at the bow and

the far-field behavior of the solution at each hull cross


section. In a fixed frame of reference with respect to
the unperturbed fluid the problem can be re-written as a
two-dimensional time-dependent problem. This is fully
equivalent to the unsteady problem of the free-surface
flow generated by a body deforming in time (see right
sketch in Figure 1). Consistently, the resulting approximation is called here 2D+t. The 2D unsteady problem is solved through the Mixed Eulerian Lagrangian
method (Longuett-Higgins and Cokelet 1976, Faltinsen
1977, and originally suggested by Ogilvie 1967), that
is the problem is split in a kinetic and a time evolution
step. The kinetic problem for is solved by means of
the Greens second identity used as integral representation of the velocity potential. Applying the latter at the
domain boundary leads to an integral equation for the
unknown boundary data /n and , on the free surface and body boundary, respectively. The continuity
of the velocity potential is enforced at the intersection
points between the body and the free surface. A Boundary Element Method (BEM) with linear shape functions
for the geometry and boundary data is then introduced.
The relevant integrals are computed analytically, and
after some manipulations the discretized integral equations lead to a system of linear algebraic equations for
the unknowns at the collocation points. The system influence matrices are only dependent on the geometry of
the problem. In the time evolution step, the free-surface
boundary conditions, expressed in a Lagrangian form,
and the body velocity are integrated in time to provide
the new boundary configuration and related data for the
next time instant. The time stepping is performed by a
fourth-order Runge-Kutta method. The discretization of
the free surface is controlled through numerical regridding and the grid refinement is adapted to the evolution
of the solution. If the angle between the body and the
free surface becomes too small, the jet-like flow created
is partially cut to avoid numerical errors (cf. Zhao and
Faltinsen 1993). Unphysical reflection of the outgoing
waves is prevented by using a damping layer technique
(for short waves) and a panel stretching (for longer wave
components) toward the edges of the computational domain. Invariance of the solution under mesh refinement
and size changes of the computational domain has been
widely checked. Since a BEM is used, bow wave postbreaking phenomena cannot be studied. The breaking
is limited by cutting off the jet flow in the plunging bow
waves. This is not believed to be an important error

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

source.
2

LINEAR 3D METHOD

In our analysis of the steady and unsteady mono-hull


and catamaran flows, we used a linear 3D method as additional numerical instrument. In the following its features are briefly outlined, for more details see i.e. Nakos
(1990).
We consider the free-surface flow generated by
a ship advancing at constant forward speed U in regular incoming waves. The problem is solved by using the potential flow theory and neglecting the nonlinearities connected with the wave-body interactions.
In case of vessels with transom stern, the latter is considered always wet. The total velocity potential of the
fluid is decomposed as the sum of the steady and unsteady wave fields, the latter consisting of the incident,
the diffraction and the radiation waves. Then the problem is split into eight sub-problems: one steady and
seven unsteady. Assuming small amplitude of the incident waves and ship motions, we linearize the free surface conditions for the unsteady problems around the
steady free surface. Then a further linearization is performed. Two alternatives are considered: the DoubleModel (DM) and the Neuman-Kelvin (NK) linearization (see i.e. Nakos 1990). These are, respectively, a
low-speed and a slender-body approximation. In this
way, the free-surface boundary conditions can be transfered on the undisturbed free surface and the steady and
unsteady problems may be solved separately. Once this
is accomplished, the pressure can be evaluated from the
Bernoullis equation. Its integration over the ship surface furnishes the added mass and damping coefficients,
the restoring forces and the wave exciting forces. The
restoring terms are due to both the hydrostatic pressure
and the steady hydrodynamic pressure. The Response
Amplitude Operator (RAO) is obtained by coupling the
fluid dynamic problem with the body motion through
the hydrodynamic loads. Finally, the steady wave pattern and the radiation and diffraction waves are calculated from the related kinematic free-surface conditions.
Here both the steady and unsteady potentials are
represented in terms of source distributions on the body
H and the free surface W. The problem is then solved
numerically by using a lower order Rankine Panel Method (RPM, Hess and Smith 1996). This leads to a

system of algebraic equations for the unknown source


strengths at the collocation points. The collocation points on H are placed at the centers of the corresponding
panels. The same is made for those along the free surface but additionally they are rigidly shifted one panel
downstream in order to enforce numerically the radiation condition (see i.e. Bertram 1990). In practice,
this numerical radiation condition is valid only for =
U e /g > 0.25, when waves do not propagate upstream
the ship. Here e is the encounter circular frequency.
An important task is represented by the evaluation of
the mj terms in the body boundary conditions for the
radiation problems. They represent the interaction with
the steady flow and their estimate can lead to relevant
numerical errors. Therefore the mj terms are often neglected. Here, in the DM case, they are estimated by
using an extrapolation procedure for the velocity gradient on the body boundary. In the NK case the effect
of the local steady flow is not incorporated in the mj
terms and the evaluation of the non-zero mj terms is
straightforward.
3

EXPERIMENTAL SET-UP

A dedicated and comprehensive experimental investigation has been performed to analyze the steady and
unsteady behaviour of semi-displacement mono-hulls
and catamarans. A catamaran model was built consistently with the geometric ratios normally used for
semi-displacement catamarans. The main characteristics are reported in table 1. The same parameters have
been considered for the mono-hull geometry coinciding
therefore with a catamaran demi-hull. This leads to a
shape finer than those of the existing semi-displacement
mono-hulls. The experimental activity has been carried out at the INSEAN basin No. 2: 220 m long, 9 m
large and 3.6 m deep. During the tests, each model was
towed by the carriage through a constant force mechanism. Trim and sinkage were free while the center of
rotation was fixed to the center of gravity of the vessel.
The model tests reproduced two main conditions: forward motion in calm water and in head sea waves. In the
former case, several Froude numbers have been inves
tigated for the mono-hull geometry: F r = U/ gL =
0.3 0.8 with a step F r = 0.1. A smaller number of
speeds was considered for the catamaran, corresponding to F r = 0.3, 0.4, 0.5, 0.6. A capacitance wave
probe system was used to measure the wave field near

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

as those used to measure the external wave pattern of


the catamaran have been considered for the mono-hull.

L
LCG
KG
r55
T
BWL
2p

= 6.25
= 25 m
= 1.7 m
= 3 m
= 0.26 L
= 1.75 m
= 2 m
= 5 m
= 40.48 m3

Figure 2: Body plan of the demi-hull and hydrostatic


properties. 2p indicates the distance between the centrelines of the catamaran demi-hulls, cf. Figure 3.
the hull. This was composed by an array of 40 transducers placed transversally to the basin. The distance
between two consecutive wave probes was 4 cm. To
refine the grid of measure each run was repeated shifting the sensor array of 2 cm transversally to the vessel
axis. With this set-up, the wave field has been recorded
within a lateral distance of 0.5L from the model hull. A
very detailed picture of the steady wave pattern has been
achieved, also including the area between the two demihulls in the catamaran case. Figu-re 3 gives a sketch of

Figure 3: Sketch: top view of the wave probes (indicated by the dots) used along the external profile of a
catamaran demi-hull and along the vessel central axis.

the wave probe arrangement along the external profile


of a catamaran demi-hull and along the vessel central
axis. The external probes were placed to follow the ship
profile at a distance of 3 cm from the hull. The sensors
were fixed to the carriage and therefore moved with the
vessel forward speed. The same wave probe locations

In the tests in waves, experimental transfer functions in heave and pitch were estimated both by a transient test technique and regular waves of different steepnesses. More in detail, Response Amplitude Operators
(RAO) have been determined preliminarily by a transient test technique. In this way the frequency range
characterizing the resonance area of heave and pitch
motions has been identified. Then, for this range of frequencies, tests have been carried out in regular head sea
waves with different steepnesses and considering several Froude numbers. Also in this case we measured the
wave profile along the hull and the centreplane of the
catamaran.
Our studies are of fundamental nature and particular care has been taken in performing a dedicated
error analysis; for instance each test condition was repeated between 5 and 10 times to ensure repeatability.
The error analysis did not investigate the error bias but
just the precision error. Results of such study will be
discussed in the following sections for some variables
of interest. Due to space limits the unsteady investigation will be presented for the catamaran only.
4

DISCUSSION: STEADY CASE

Systematic calm water experiments have been carried


out for both the mono-hull and catamaran models. Such
study aimed to a better understanding of the steady wave
pattern features and to verify the validity and limits of
the numerical methods presented in sections 1 and 2.
The NK linearization is used for the linear 3D (3D RPM)
results, that is the basis steady flow is assumed given
just by the uniform flow due to the ship speed. The results of the repeatability analysis for the model tests are
reported by presenting the experiments as mean measured values and error bars. The latter are given as
, with the standard deviation. The wave profiles
along the mono-hull are presented in figure 4 for two
selected Froude numbers, respectively, F r = 0.5 (top)
and F r = 0.7 (bottom). Numerically the wave elevation was calculated following the ship profile at a distance of 3 cm from the hull. This was made to be consistent with the measurements (see section 3). Globally
the agreement between the experiments and the numerical results is satisfactory. Locally large discrepancies

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

/L*102
0.8
0
-0.8
-0.4

1.58

-0.2

0.91

0.71

0.4 x/L
0.53 Frx

0.2

0.6
/L*102

1.6

-1.6
-0.4

-0.2

2.21

1.28

0.2

0.99

0.84

0.4 x/L
0.74 Frx

Figure 4: Wave profile along the mono-hull for


F r = 0.5 (top) and F r = 0.7 (bottom). Experiments
(symbols: mean value and error bar) and numerical results obtained by the 3D RPM (dashed line) and 2D+t

(solid line) methods. F rx = U/ gx, with x the longitudinal distance from the bow.

can be detected in the bow area between the model tests


and the 2D+t theory at the smaller Froude number. The
former predict a negative wave elevation, the latter gives
a quite large free-surface rise along the hull. This disagreement can be partially explained by ventilation of
the sensors in the bow area. Such phenomenon implies
an underestimated wave elevation since the probes stay
more dry than they should be. This is indirectly confirmed by a larger experimental error bar near the bow.
The two results fit quite well at the larger speed. In
this case, the 2D+t model is able to capture the bow
splash phenomenon. The stern wave disturbance is predicted correctly by the nonlinear method for both the
speeds. Also at this Froude number the experimental repeatability is less satisfactory in the vicinity of
the bow, the same phenomena described for the smaller
speed are believed to be the main reason for it. The
3D RPM results agree better with the experiments at
the smaller speed, when the nonlinearities are less important and the three-dimensional effects are stronger.
At the larger F r the agreement is still satisfactory but a
smoother bow rise-up and a deeper trough near the stern
are predicted as a consequence of the linear formula-

tion. The steady wave patterns at F r = 0.5, 0.6, 0.7


and 0.8 are reported in figures 5 and 6 where the experiments are compared with the 2D+t and 3D RPM results, respectively. For the smallest ship-length Froude
number shown, the linear 3D computations capture correctly the phasing of the wave pattern and give reliable quantitative information in the bow area and along
the ship hull. The stern and downstream flows are not
properly described since the numerical transom is assumed wet while at this speed the experimental transom was already dry. Such aspect will be discussed
later. Concerning the 2D+t results, the agreement with
the experiments is just fair along the hull in terms of
the phasing of the wave system generated. The predicted amplitudes are closer to the measurements in the
bow area and tend to underestimate them largely going
downstream. A global extension of the lobes (contour
lines with constant elevation) is observed and it is not
shown by the measurements. The comparison slightly
improves behind the stern. This suggests the relevance
of nonlinearities in the wake and the need of handling
the dry stern condition. In this context, the 2D+t theory
assumes that the flow leaves tangentially the transom
stern in the downstream direction. This approximation
appears suitable to capture the transom flow behavior.
Ohkusu and Faltinsen (1990) showed that theoretically
the 2D+t theory should not be applied for ship-length
Froude smaller than 0.4. A reason is that the method
neglects the transverse wave system generated by the
ship while this is important at such speeds. As F r increases, the transverse waves become longer and the related importance reduces with respect to the divergent
waves. Therefore a 2D+t formulation can handle satisfactorily the features of the wave pattern. According
to our results, reliable results can be obtained for F r at
least larger than 0.5. Faltinsen (2000) defined a local

Froude number F rx = U/ gx, with x the longitudinal distance from the ship bow. He used such a local
number to estimate the goodness of the 2D+t theory for
the ship bow waves prediction. A similar local Froude
number should be introduced for the ship waves generated at the transom. In this case x is the longitudinal distance from the ship stern. For a given F r, the
2D+t results can be considered suitable within the region where F rx is sufficiently large, say greater than
0.4. Obviously such region enlarges as the ship speed
increases. The local bow Froude number is reported in
the figures.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Figure 5: Contour lines of the mono-hull steady wave


pattern. From top to bottom: F r = 0.5, 0.6, 0.7 and
0.8. In each plot the experiments (bottom) are com
pared with the 2D+t results (top). F rx = U/ gx, with
x the longitudinal distance from the bow.

Figure 6: Contour lines of the mono-hull steady wave


pattern. From to bottom: F r = 0.5, 0.6, 0.7 and 0.8. In
each plot the experiments (bottom) are compared with

the 3D RPM results (top). F rx = U/ gx, with x the


longitudinal distance from the bow.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Globally, for F r larger than 0.5 the 2D+t theory gives


better results than the 3D RPM code. This is due to two
factors: the transverse waves become progressively less
relevant and the importance of nonlinearities increases.
The linear results underestimate the peaks and troughs
along the hull and in the wake. However the dimension
of the lobes of the global wave system is reproduced
quite nicely while the 2D+t results show a wider extension of the lobes even at the largest speed.
The wave profiles along the catamaran are presented in figure 7 for the two largest speeds tested, that
Fr=0.5

Fr=0.6

1.2

1.2
0

0
-1.2

External

External

-1.2

/L*102
1.8

2.8

-1.8

-2.8

Centre-line

Centre-line

-0.4
1.6

-0.2
0.9

0
0.7

0.2
0.6

0.4 x/L
0.53 Frx

-0.4
1.9

-0.2
1.1

0
0.85

0.2
0.72

0.4 x/L
0.63 Frx

Figure 7: Catamaran: external (top) and centre-line


(bottom) wave profile (cf. figure 3) for F r = 0.5
(left) and F r = 0.6 (right). Experiments (symbols:
mean value and error bar) and numerical results obtained by 3D RPM (dashed line) and 2D+t (solid line).

F rx = U/ gx, with x the longitudinal distance from


the bow.
are F r = 0.5 (left plots) and F r = 0.6 (right plots).
The top and bottom plots refer, respectively, to the external and centre-line wave profiles, as shown in figure
3. The agreement among the results is generally fairly
good. Globally the 2D+t theory compares better with
the model tests, although local discrepancies can be detected, in particular in the external aft hull. Here the
experiments show a large error bar for F r = 0.5 and a
deeper trough than the 2D+t results for F r = 0.6. At
the latter Froude, in the stern area they are closer to the
linear 3D calculations. An error bar not negligible is
also noticed at the aft profile along the centre-line for
the smaller speed. The 3D RPM code predicts a lower
external bow splash than the experimental and 2D+t results. Also the free-surface peak at the centre-line is
underestimated. Differently larger values for the external and inner troughs are predicted. The 3D RPM wave
profiles seem to have a phase shift with respect to the
other results at the smaller speed. This suggests a role
of the nonlinearities both externally to the catamaran

Figure 8: Contour lines of the catamaran steady wave


pattern. Top: F r = 0.5. Bottom: F r = 0.6. In each
plot the experiments (bottom) are compared with the

2D+t results (top). F rx = U/ gx, with x the longitudinal distance from the bow.

and between the demi-hulls. The steady wave patterns


for F r = 0.5 and 0.6 are reported in figures 8 and 9
where the experiments are compared, respectively, with
the 2D+t and 3D RPM predictions. The linear 3D code
is more able to capture the global picture of the wave
field along the vessel, showing that the transverse waves
play still a role. However, for both the speeds, it does
not quantify correctly the free-surface disturbance due
to the ship forward motion. This implies that nonlinear
effects matter. The 2D+t theory shows a wide extension of the lobes near the bow, similarly to the monohull case, and it has a relevant phase difference with
respect to the experiments at the lower Froude. Despite this, even at F r = 0.5 the peaks and troughs are
closer to the measured values than the results obtained
with a linear method. Figure 10 shows the same results for F r = 0.7 (top) and 0.8 (bottom), as obtained
numerically with the 2D+t and 3D RPM codes. At

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

these larger speeds the nonlinearities are quite important, therefore the linear results underestimate substantially the wave pattern. On the other hand, it is interesting to note that the two results agree quite well in
terms of phasing. This confirms the unimportance of
the transverse waves at such Froude numbers. Top

Figure 10: Contour lines of the catamaran steady wave


pattern. Top: F r = 0.7. Bottom: F r = 0.8. In each
plot the 3D RPM (bottom) are compared with the 2D+t

results (top). F rx = U/ gx, with x the longitudinal


distance from the bow.
Figure 9: Contour lines of the catamaran steady wave
pattern. Top: F r = 0.5. Bottom: F r = 0.6. In each
plot the experiments (bottom) are compared with the 3D

RPM results (top). F rx = U/ gx, with x the longitudinal distance from the bow.
picture of figure 11 shows the experimental wave field
behind the transom for the mono-hull at F r = 0.5. The
dry transom causes the formation of a hollow just behind the vessel. The water reaches a minimum value
and then rises to form a rooster tail developing into a
divergent breaking wave system. The mono-hull transom flow features have been thoroughly investigated by
Landrini et al. (2001) due to the practical relevance
of the resulting breaking phenomena. These lead to
vortical structures responsible of the visible signature
left downstream by the ship. In their study, the authors
used the 2D+t theory. The transom was enforced to be

dry. A nonlinear BEM solver, as used in the present


study, was adopted to simulate the flow evolution until the incipient wave breaking and a Smoothed Particle Hydrodynamics (SPH) method was initialized by
the BEM to handle the post-breaking evolution of the
wave system. In the bottom of the figure 11, the results obtained by using just the BEM 2D+t method for
F r = 0.5 are given. As we can see, the 2D+t formulation is able to reproduce the flow scenario behind the
transom: hull hollow, rooster tail and incipient breaking
divergent wave system. Nevertheless, since the plunging jet is cut to avoid the occurrence of impact on the
underlying water, the energy of the wave system is focused close to the crest of the divergent wave. Differently, in the physical phenomenon the breaking causes
a spatial spread of the wave energy. Figure 12 shows
the longitudinal wave cut along the centre-line of the

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

/L*102

3
2
1
0

Fr=0.5
Fr=0.6
Fr=0.7
Fr=0.8

-1
-2
-3

0.1

0.2

0.3

0.4

0.5

0.6

0.7

x/L

Figure 12: Mono-hull: cut of the steady wave pattern


along the centre-line of the transom stern. x is the longitudinal distance from the transom.

0.1

/L*102

-0.38

y/L

-0.19

-1
0.19

0.19

Fr=0.5
Fr=0.6
0.38 Fr=0.7
Fr=0.8

-2
-3

-0.38

-0.1

-0.96

-0.19

Fr x

0.2
1.12

0.4
0.79

x/L

-0.19

Figure 11: Mono-hull: transom stern wave field at


F r = 0.5. Experimental picture (top) and contour lines
of the steady wave pattern predicted by 2D+t theory

(bottom). F rx = U/ gx, with x the longitudinal distance from the transom.

mono-hull transom as obtained by the 2D+t theory. The


results show that the rooster tail height is not affected by
the Froude number while both its horizontal width and
the extension of the hollow increase with the speed. The
hollow extension can be measured as the longitudinal
distance between the transom position and the location
where the free surface becomes zero. The wave elevation downstream a catamaran demi-hull transom (see
figure 13) shows a quite different behaviour. Except for
the smaller speeds, showing an increase of the hollow
extension with the Froude number, the hollow width
is not particularly affected by the speed. The rooster
tail height is lower than the corresponding value for the
mono-hull. It shows a non monotonic but rather limited variation with the Froude number. This demi-hull
wave behaviour behind the transom is due to the presence of the other demi-hull. The arrangement of the
two demi-hulls causes three rooster tails downstream
the catamaran, respectively, in correspondence of the
demi-hull transom sterns and of the catamaran centreline (cf. 2D+t contour plots in figures 8 and 10).
It is important to understand if the main responsible

0.1
0.38

0.2

0.3

0.4

0.5

0.6

0.7

x/L

Figure 13: Catamaran: cut of the steady wave pattern


along the centre-line of the demi-hull transom stern. x
is the longitudinal distance from the transom.

mechanisms are connected with the hydrodynamic interaction between the demi-hulls or if they are related
to the demi-hulls interference only. The latter means the
diffraction caused by one demi-hull on the waves generated by the other demi-hull is negligible and the catamaran wave pattern is just given by the sum of the wave
fields produced by each demi-hull as if the other was not
there. To this purpose figure 14 gives the longitudinal
wave cut along the centre-line of the catamaran for different Froude numbers. The 3D RPM and 2D+t calculations are presented together with the experiments (for
the tested speeds). In the plots, the curves with circles
give the 2D+t results obtained as the superimposition
of two mono-hulls solutions, that is the interaction between the demi-hulls is not accounted for but just their
interference. From the results, the interference is not the
governing mechanism. The interaction between the two
demi-hulls plays a fundamental role. This interaction
is mainly nonlinear as evidenced both by difficulties of
the linear solution in capturing the first peak and by the
phase shifting existing between the linear and nonlinear
results accounting for the demi-hull interaction.
As previously discussed, the 2D+t model enforces a dry transom stern condition independently from
the forward speed. In reality the flow at the transom
stern can be quite complicated at sufficiently small Frou-

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Fr
03
04
05
06
07
08

3.6

/L*10

1.8
0

F rT
176
209
246
291
332
382

F rT trans
182
243
304
365
426
487

Transom
wet
wet
dry
dry
dry
dry

Table 1: Monohull: Dry and wet transom stern conditions as a function of different Froude numbers.

-1.8

-0.4
1.6

0
0.7

0.4
0.53

0.8
0.44

1.2
0.38

x/L
Frx

2.8

/L*102

T /L
00290
00364
00413
00424
00444
00438

Fr
03
04
05
06

T /L
00290
00360
00525
00598

F rT
176
211
218
245

F rT trans
182
243
304
365

Transom
wet
wet
dry
dry

Table 2: Catamaran: Dry and wet transom stern conditions as a function of different Froude numbers.

-2.8

-0.4
1.9

0
0.85

0.4
0.63

0.8
0.53

1.2
0.46

x/L
Frx

-0.4
2.2

0
0.99

0.4
0.74

0.8
0.61

1.2
0.54

x/L
Frx

-0.4
2.53

0
1.13

0.4
0.84

0.8
0.7

1.2
0.6

x/L
Frx

/L*102

2.8

-2.8

5.6

/L*10

2.8
0

-2.8
-5.6

Figure 14: Catamaran: cut of the steady wave pattern along the catamaran center-line. Experimental data
(square symbols: mean value and error bar), 3D RPM
code (dashed lines) and 2D+t theory (solid lines). The
curves with circles give the 2D+t results obtained as the
superimposition of two mono-hulls solutions, that is the
interaction between the demi-hulls is not accounted for.
From top to bottom: F r = 0.5, 0.6, 0.7, 0.8.

de numbers, involving partial or full ventilation phenomena. Several flow regimes can be distinguished and it
is difficult to identify the critical Froude for the transition from one regime to another. In this context, analytical (Vanden-Broeck and Tuck 1977), numerical (Scor-

pio and Beck 1997) and experimental (Doctors 2003


and Maki et al. 2004) investigations have been carried out. Vanden-Broeck and Tuck (1977) and Scorpio

and Beck (1997) identified F rT = U/ gT = 2.5-2.6


as the critical Froude number based on the draft (T ) for
the occurrence of ventilation behind a 2D semi-infinite
body with a flat bottom. The experiments by Doctors
(2003) on transom-stern ships confirmed this critical
value.
Table 1 gives the F rT corresponding to the
speeds tested in our mono-hull experiments. Due to
the hull geometry (see figure 2) two different Froude
numbers, F rT and F rT trans , are reported based, respectively, on the real draft at the transom (including
the trim and sinkage effects) and on the transom draft in
still water conditions. The last column in the table gives
the transom stern conditions at the different speeds. The
critical F rT value is about 2.46. The value has been
identified by using top-view pictures from the experiments, as reported in figure 15 for the different tested
speeds. At F r = 0.3 (F rT = 1.76) a partial ventilation
of the flow appears. This becomes a full ventilation at
F r = 0.4 (F rT = 2.09). Increasing the model speed
(F r 0.5, i.e. F rT 2.46) the dead-water region behind the transom disappears and the transom stays completely dry.
In table 2 the same parameters discussed before
are reported for the catamaran hull. The corresponding
experimental top-view pictures are given in figure 16.
From these, the transition to dry transom condition occurs at F r = 0.5 (F rT = 2.18). The catamaran configuration presents a larger value of the trim and sinkage
(cf. tables 3 and 4) than the mono-hull. This implies
smaller F rT with respect to the mono-hull.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Fr
03
04
05
06
07
08

Velocity (m/s)
U

1880
0001
2510
0001
3133
0001
3754
0.0001
4384
0.0001
5008
0003

Sinkage (mm)
s

-6189
0110
-13974
0057
-19637
0057
-16684
0029
-14054
0021
-12536
0413

Trim (degree)

-0010
0001
0543
0093
1177
0004
1424
0003
1348
0001
1328
0022

Table 3: Monohull: mean value for carriage speed (U ),


sinkage(s) and trim () and related standard deviation
().
Fr
03
04
05

Velocity (m/s)
U

1883
00005
2509
00003
3125
0006

Sinkage (mm)
s

-8792
0321
-21385
0072
-27062
0223

Trim (degree)

0025
0008
0424
0009
2652
0021

Table 4: Catamaran: mean value for carriage speed (U ),


sinkage(s) and trim () and related standard deviation
().

Figure 15: Pictures of the monohull transom stern


field. From top to bottom: F r = 0.3, 0.5, 0.7 (left column), F r = 0.4, 0.6, 0.8 (right column).

Figure 16: Pictures of the Catamaran transom stern


field. From top to bottom: F r = 0.3, 0.4, 0.5, 0.6.

DISCUSSION: UNSTEADY CASE

A second experimental activity was dedicated to the


study of the unsteady behaviour of the mono-hull and

catamaran models. Preliminarily, the heave and pitch


frequency resonance ranges for F r = 0.3, 0.4 and 0.5
have been identified through the transient test technique
(Colagrossi et al. 2001). Then, tests in regular incoming waves within such range and with different wave
amplitudes have been performed. Also in this case a
careful error analysis has been realized. In the following the results will be discussed for the catamaran.
The RAO experimental data are presented in figure 17 together with the predictions by the 3D linear
RPM code. The standard deviation () connected with
the transient test technique is also given in the plots
showing a good reliability of the experiments. For the
numerics both the NK and DM approximations (see section 2) are considered. The NK approach neglects the
interaction with the local steady flow. The DM accounts
for it but not in a consistent way for a high-speed ship.
The numerical results overestimate the pitch motion.
For all investigated speeds, the DM linearization shows
the best agreement with the experiments. This is consistent with the conclusions in Bertram (1999) documenting an important role of the interaction with the steady
flow in the wave induced body motions even at Froude
around 0.2, for a S-175 ship. A strong amplification
of the motions is generally observed near the resonance
due to a small damping level. This suggests the need
of proper active control systems and foils. However it
should also be noted that the values of the wavelengthto-ship length ratio giving resonance in heave and pitch
in head sea increase with the Froude number. This im-

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

1.5

8
6

0.5

/L

/L

/L

3
15
2.5
12

L/ A

1.5

0.5
0

/L

3.2
16

L/ A

1.6

12

8
0.8

/L

/L

Figure 17: Catamaran: heave (, left) and pitch (,


right) Response Amplitude Operators. From top to bottom: F r = 0.3, 0.4, 0.5. A and k = 2/ are the
regular incoming wave amplitude and wavenumber, respectively. is the incoming wavelength.

TAV (degree)

Calm Water
kA=0.0125
kA=0.01875
kA=0.025

Fn = 0.3 ----> = 0.55


Fn = 0.4 ----> = 0.51
Fn = 0.5 ----> = 0.47

-8

-16

-24

-24
Fn = 0.3 ----> = 0.6
Fn = 0.4 ----> = 0.54
Fn = 0.5 ----> = 0.49

0.3
0.60

Fr = 0.5

-8

-16

Fn = 0.3 ----> = 0.55


Fn = 0.4 ----> = 0.51
Fn = 0.5 ----> = 0.47

SAV (mm)

0.4
0.54

Fr
0.5
0.49 (Hz)

Fn = 0.3 ----> = 0.6


Fn = 0.4 ----> = 0.54
Fn = 0.5 ----> = 0.49

0.3
0.55

0.4
0.51

Fr
0.5
0.47 (Hz)

Figure 19: Catamaran: mean trim and sinkage in calm


water and in regular waves. Regular waves have been
generated with frequency (Hz) close to the heave (,
left) and pitch (, right) resonance frequencies.

2.5
2
/A

/A

2.4

TAV (degree)

/A

SAV (mm)

/A

10

L/ A

2.5

plies that the excitation loads along the ship become


stronger in phase as F r increases. The consequence is
larger excitation loads. The experiments evidence clear
nonlinear effects. The regular wave results do not show
a convergence to the transient test results as the wave
amplitude reduces. One possible error source is a variation of the wave amplitude along the track of the model.
This aspect has not been investigated. At high Froude
numbers the RAO for the pitch motion shows a double
peak behaviour, typical for the multi-hull vessels. So it
proves that it is important to account for the hull interaction in the ship motions calculations. The RAO related to the heave motion is practically unchanged with
respect to the mono-hull configuration (see figure 18).
From the experiments, the mean trim and sinkage are
not influenced substantially by the incident wave steepness, even at a wave frequency equal to the heave and
pitch resonance frequency, as shown in the left and right
plots of figure 19, respectively. It implies that they are

12
Trans. test tech.
(Trans. Test)
reg. wave (kA = 0.0125)
reg. wave (kA = 0.0187)
reg. wave (kA = 0.025)
reg. wave (kA = 0.05)
Num. results (NK base flow)
Num. results (DM base flow)

1.5
1
Catamaran

0.5

Monohull
0

/L

Figure 18: Heave () Response Amplitude Operator for


the mono-hull (dashed) and catamaran (solid) vessels
at F r = 0.5. A and are the regular incoming wave
amplitude and wavelength, respectively.

dominated by the steady flow. These results are relevant


for instance for the wetdeck slamming which is sensitive to the trim angle (Ge 2002). In the wetdeck slamming predictions it also matters an accurate estimate of
the relative motions in the impact area. The presented
results suggest that the theoretical methods to evaluate
wave induced motions have to be improved for a better
prediction of for instance wetdeck slamming.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

CONCLUSIONS
An experimental investigation has been carried out to
analyze the flow field around semi-displacement monohulls and catamarans both in calm water and in incident head sea waves. The mono-hull model has been
shaped identically to a catamaran demi-hull to investigate the interaction between the demi-hulls of the catamaran and the related influence on the rooster tail developing from the transom stern. The chosen mono-hull
geometry implies a much smaller beam-to-draught ratio than the one characterizing the usual mono-hull high
speed vessels. A dedicated error analysis of the tests has
been performed confirming a general reliability of the
measurements. The physical investigation was focused
on the wave-field features at the bow, along the hull
and downstream the transom stern. In the steady experiments, very detailed measurements of the wave pattern were performed for both models, including the inner region between the two catamaran demi-hulls. The
influence of the Froude number has been analyzed by
varying such parameter in a wide range. The experimental data were compared with the results by a linear
3D RPM code and a nonlinear 2D+t method. For the
mono-hull, the 3D RPM simulations are able to capture the wave pattern along the hull for Froude numbers
smaller than 0.6. They do not succeeded in handling the
stern and wake flows for F r 0.5. This is because a
wet transom is assumed while at those speeds both the
experimental mono-hull and catamaran transoms were
dry. At F r 0.6 the nonlinearities become relevant
and the linear method gives only a qualitative information. The opposite trend is shown by the 2D+t results.
These are not satisfactory at the smaller Froude due to
the relevance of the transverse wave pattern. For F r
greater than 0.6 they fit well with the experiments. The
stern and wake flows are properly described since a dry
transom condition is enforced. For the catamaran, the
same trend is observed but the 2D+t theory gives globally the best agreement at smaller speeds than for the
mono-hull. For both geometries this model predicts a
wide extension of the lobes near the bow region. This is
not observed from the measurements. The 2D+t model
has been used to investigate the physical mechanisms
causing the flow features downstream the transom, and
to quantify the related influence of the interaction between the catamaran demi-hulls. In the unsteady experiments, both a transient test technique and regular in-

coming waves have been used to simulate seaway conditions. The former method was applied to identify the
heave and pitch resonance frequency regions. Then regular incoming waves were generated within such areas
of interest. This means we studied a frequency range
where nonlinearities matter for the wave-body interactions. During the tests the Froude number and the amplitude of the incoming waves have been varied and the
importance of nonlinear effects brought into the problem has been deduced. The interaction between the
steady and unsteady wave fields was studied by combining in a synergic way the information from the experiments with the results from a linear unsteady 3D code.
The results confirmed that the interaction plays a relevant role and should be properly accounted for. For the
considered speeds and within the heave and pitch resonance ranges, the experiments showed that the mean
trim and sinkage of the catamaran are mainly governed
by the steady flow. This is relevant for instance for the
wetdeck slamming.
ACKNOWLEDGEMENTS
Present research activity is partially supported by the
Centre for Ships and Ocean Structures, NTNU, Trondheim, within the Green Water Events and Related Structural Loads project, and partially done within the framework of the Programma di Ricerca sulla Sicurezza
funded by Ministero Infrastrutture e Trasporti.
REFERENCES
Bertram, V. Fulfilling Open-Boundary and Radiation
Condition in Free-Surface Problems Using Rankine
Sources. Ship Technology Research 37(2), 1990.
Bertram, V. Numerical innvestigation of steady flow
effects in three-dimensional seakeeping computations.
Proceedings 22th Symposium on Naval Hydrodynamics. Washington D.C., USA, 1999.
Colagrossi, A., C. Lugni, M. Landrini, and
G. Graziani. Numerical and experimental transient
tests for ship seakeeping. Int. Journal Off. and Ocean
Struct., Volume 11, pp. 6773, 2001.
Doctors, L. Hydrodynamics of the flow behind a transom stern. Proceedings of 29th Israel Conference on
Mechanical Engineering, pp. 111. Haifa, Israel, 2003.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Doctors, L. and A. Day. Resistance prediction for


transom-stern vessels. Proceedings of 5th International
Conferance for Fast Sea Transportation (FAST97),
Volume 2, pp. 743750. Sydney, Australia, 1997.

Maki, K., A. Troesch, and R. Beck. Qualitative investigation of transom stern flow ventilation. 19th International Workshop on Water Waves and Floating Bodies.
Cortona, Italy, 2004.

Faltinsen, O. Numerical solutions of transient nonlinear free-surface motion outside or inside moving bodies. Proc. 2nd Int. Conf. Num. Ship Hydr., 1977.

Maruo, H. and W. Song. Nonlinear analysis of bow


wave breaking and deck wetness of a high speed ship
by the parabolic approximation. Proc. 20th Symp. on
Naval Hydrod. National Academy Press, 1994.

Faltinsen, O. Steady and vertical dynamic behaviour of


prismatic planing hulls. Proceedings HADMAR 2001.
Varna, Bulgaria, 2001.
Faltinsen, O. and R. Zhao. Numerical predictions of
ship motions at high forward speed. Phil. Trans. R. Soc.
Lond. A Vol. 334, pp. 241252, 1991.
Faltinsen, O. M. Sea loads on ships and offshore
structures. Cambridge, England: Cambridge University
Press, 1990.
Faltinsen, O. M. Sea loads on High Speed Marine Vehicles. Trondheim, Norway: NTNU, 2000.
Fontaine, E., O. Faltinsen, and R. Cointe. New insight into the generation of ship bow waves. J. Fluid
Mech. Vol. 421, pp. 1538, 2000.
Ge, C. Global Hydroelastic Responce of Catamarans
due to Wetdeck Slamming. Ph. D. thesis, Dept. Marine
Structures, NTNU, Trondheim, Norway, 2002.
Greco, M., O. M. Faltinsen, and M. Landrini. Water Shipping on a Vessel in Head Waves. Proceedings
24th Symposium on Naval Hydrodynamics, technical
session: Slamming, Green Water and Capsizing, pp. 1
14. Fukuoka, Japan, 2002.
Hess, J. and A. Smith. Calculation of nonlifting
potential flow about arbitrary bodies. Prog. Aero.
Sci. Vol. 8, pp. 1138, 1996.
Keuning, J. Distribution of added mass and damping
along the length of a ship model at high forward speed.
Technical report, Report No. 817-P, Delft University of
Technology: Ship Hydrodynamics Laboratory, 1988.
Landrini, M., A. Colagrossi, and M. Tulin. Breaking
bow and transom waves: numerical simulations. 16th
International Workshop on Water Waves and Floating
Bodies. Hiroshima, Japan, 2001.
Longuett-Higgins, M. S. and E. D. Cokelet. The deformation of steep surface waves on water. I A numerical method of computation. Proc. R. Soc. London A,
Volume 350, pp. 126, 1976.

Michell, J. The wave resistance of a ship. Philosophical Magazine Vol. 45, pp. 106123, 1898.
Molland, A., J. Wellicome, and P. Couser. Theoretical
prediction of the wave resistance of slender hull forms
in catamaran configurations. Technical report, Report
No. 72, Southampton University: Ship Science, 1994,
September).
Molland, A., J. Wellicome, and P. Couser. Resistance
experiments on a systematic sense of high speed displcement hull forms: Variation of length-displacement
ratio and breadth-draught ratio. The Royal Institution
of Naval Architects, 1995.
Nakos, D. Ship Wave Patterns and Motions by a
Three Dimensional Rankine Panel Method. Ph. D. thesis, MIT, Cambridge, 1990.
Ogilvie, T. F. Nonlinear high-Froude-number free surface problems. J. of Engineering Mathematics 1(3), pp.
215235, 1967.
Ohkusu, M. and O. Faltinsen. Prediction of radiation
forces on a catamaran at high froude number. Proc.
of 18th Symp. on Naval Hydrod., pp. 520. National
Academy Press, Washington D.C., USA, 1990.
Ohkusu, M. and O. Faltinsen. Prediction of radiation
forces on a catamaran at high froude number. Proc.
of 18th Symp. on Naval Hydrod., pp. 520. National
Academy Press, Washington D.C., USA, 1991.
Scorpio, S. and R. Beck. Two-dimensional invscid
transom stern flow. 12th International Workshop on
Water Waves and Floating Bodies. France, 1997.
Vanden-Broeck, J. and E. Tuck. Computation of nearbow or stern flows, using series expansion in Froude
number. 2nd International Conf in Num. Ship Hydro.
Berkeley, California, 1977.
Zhao, R. and O. M. Faltinsen. Water entry of twodimensional bodies. J. Fluid Mech. Vol. 243, pp. 593
612, 1993.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Volker Bertram
ENSIETA, France
Ship seakeeping, particularly for high
Froude numbers as treated by the authors, poses
still many challenges and requires considerable
skill in determining the appropriate simplification
in the model. The authors have chosen too possible
models in their analyses, namely a linear 3-d
Rankine panel method without transom stern model
and a 2.5 D method which models (at least in good
approximation) also the highly nonlinear wave
deformation.
I find the validation against detailed and
repeated experiments most interesting. We need
more such validation, preferably with more
research groups and different methods involved.
The Gothenburg 2000 workshop, Larsson et al.
(2003), showed the benefits of such comparisons
for steady flows, with a recommendation to move
towards unsteady flows. Benchmark tests in Japan,
described in English in Bertram and Iwashita
(1996), for ship seakeeping at high Froude numbers
yielded also interesting insight, but almost a decade
later numerical methods have progressed
considerably.
Could
hull
geometry
and
experimental results be made available in electronic
form for everybody?
For the steady case, I am puzzled by the
good agreement between computation and
experiments in Fig.4, though. Does the 2.5 D
method consider trim and sinkage? Often the nonlinear aspect is limited to a treatment of the nonlinear free-surface deformation. The 3-D RSM has
definitely no trim and sinkage considered.
However, at Froude numbers of 0.5 and 0.7 there is
quite good agreement for the wave profile
measured free in sink and trim and I would expect
at least a noticeable sinkage. What were measured
trim and sinkage for the mono-hull in the steady
experiments? Were the plotted results corrected for
trim and sinkage or can you explain the good
agreement despite different treatment of trim and
sinkage?
Steady trim and sinkage could be
predicted e.g. by a nonlinear wave resistance code
and then be considered in any seakeeping method
as average position. As a first insight, one could
take the measured values and repeat the numerical
seakeeping simulations to see the influence. It
would be interesting to see how a simple estimate
of trim and sinkage influences the results with
standard linear seakeeping codes for cases as the
one considered here with high Froude numbers.

Dry transoms require special treatment in


potential flow methods. While probably all
commercial wave-resistance codes claim to be able
to treat dry transoms, this is probably too optimistic
and based on success for moderate transom sterns.
Fast hull shapes with maximum cross section area
at the transom pose still considerable problems for
codes based on Rankine panels, while field
methods handle these flows apparently well.
Nevertheless, I encourage the authors to develop
their method further to include some transom stern
model, because this may change relative
performance. Again, having the test cases public
domain may stimulate other research groups to try
their codes and collectively we may form a picture
of which approaches perform well.
In the seakeeping results for the
catamaran, I am surprised that the double-body
linearization performs better than the NeumannKelvin linearization, even though I found similar
effects for the S-175. The S-175 was at half the
Froude number, thus slow in comparison, and the
geometry featured a much blunter waterline
entrance. Further numerical studies may shed light
on what the driving effect is in the case: the freesurface elevation or the flow field at the bow.
It is always easy to specify further work or
to criticize, and always difficult to do better. This is
particularly the case for such a rich work as
presented by Claudio Lugni and his co-authors. The
work contributes to our global understanding and is
in the best of tradition of the involved two research
institutes. I look forward to seeing the continuation
of this research. Congratulations!
REFERENCES
Bertram, V. and Iwashita, H., Comparative
Evaluation of Various Methods to Predict
Seakeeping Qualities of Fast Ships, Schiff+Hafen
Vol. 48, No.6, June 1996, pp.54-58.
Larsson, L., Stern, F., and Bertram, V.,
Benchmarking of CFD for Ship Flows The
Gothenburg 2000 Workshop, J. Ship Research Vol.
47, No. 1, March 2003, pp.63-81.

AUTHORS REPLY
We appreciated the precious comments
and suggestions of Prof. Bertram and we thank him
for that.
[1] All the experimental and numerical
results, as well as the geometry of the model are

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

available in electronic form. We are arranging to


include them in the INSEAN web site
(www.insean.it). In the meanwhile you can ask the
data via e-mail to c.lugni@insean.it.
[2] The FNL 2D+t results presented in the
paper include both the trim and sinkage measured
in the experiments. Differently the linear 3D RPM
computations do not take into account the real trim
and sinkage. In our opinion, the quite good
agreement among the mathematical and numerical
models considered and the experiments for the
wave profile along the hull is due mainly to the
geometry of the model. In fact, this one is pretty
much wall sided. Differently the flow behind the
transom stern is strongly dependent on the dynamic
position of the hull as could be evinced from the
fig. 5, 6 (for the monohull) and fig. 8-10 (for the
catamaran). There the 2D+t results accounting for
the trim and sinkage reproduce better the
experimental data than the 3D RPM ones. Further,
we performed new computations with the steady
3D-RPM numerical model accounting for the
dynamic position of the hull and the real flow
transom condition. The new results, recently
presented at the HIPER conference in Rome (Lugni
et al, 2004) show that the agreement between linear
3D-RPM and experiments is signifcantly increased
accounting for the trim, sinkage and transom
conditions.
[3] Following the suggestion of Prof.
Bertram we performed new computations with
linear 3D RPM frequency domain code accounting
for the steady trim and sinkage measured during the
experiments, and the real transom flow condition.
The results, included in fig.A (catamaran: Fn=0.5),
show as the dynamic position of the hull is quite
important also in the seakeeping calculations,
highlighting the influence of the steady
characteristics on the unsteady simulations. As
suggested by Prof. Bertram it would be very
interesting to study the influence of the FNL steady
flow on the linear seakeeping prediction (Iwashita
et al. , 2000), and in this sense we are continuing
our numerical activity.

REFERENCES
Lugni C., Colagrossi A., Colicchio G., Faltinsen
O.M., (2004), Numerical and Experimental
Investigations on semi-displacement Mono and
Multi-hull., Proc. of HIPER Conf., Rome, 2004.
Iwashita, Nechita, Colagrossi, Landrini, Bertram, A
critical assessment of potential flow models for
ship seakeeping, Osaka Colloqium (2000)

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Makoto Ohkusu
Japan Agency for
Technolgy, Japan

Marine-Earth

Science

and

My sincere congratulation to Dr. Claudio


Lugni and his coauthors for such an interesting paper
on hydrodynamics of semi-displacement high-speed
vessels. My discussions and comments are as
follows:
[1] Considering the flow model assumed in
2D+t approach it is not difficult to understand that a
local Froude number based on the distance from the
bow is an appropriate index for the local accuracy of
the prediction. It is, however, not so obvious why a
local Froude number based on the distance from the
stern that the authors suggested is a parameter of
accuracy with respect to the prediction of the ship
wave generation from the transom stern. I appreciate
more explanation.
[2] Apparently the authors concluded with
Fig.14 that the hydrodynamic interaction between
demi-hulls of a catamaran is nonlinear. Indeed the
wave elevation between demi-hulls computed by
nonlinear 2D+t agrees better with the experiment
than by linear 3D. However the agreement looks
similar regardless of with or without interaction
accounted. I understand with Fig.14 that the
nonlinear effect is strong for the wave elevation yet it
will be too early to conclude the interaction is
nonlinear.
[3] Discrepancy in RAO computed with two
different base steady flow of NK and DM as seen in
Fig.17 is extremely large for pitch mode in particular.
With such slender hull-form and at high Froude
number the discrepancy is apparently larger than
expected and moreover it is difficult to understand
why DM gives better result. There must be another
reason to account for this discrepancy.
[4] It appears that the authors take that the
result in Ohkusu and Faltinsen (1991) is based on the
assumption of no hydrodynamical interaction of
demi-hulls. It accounts for the interaction, however,
correctly in terms of linear 2D+t approach: 2D
governing equation is solved with two sections. The
interaction effect is studied with the computation by
different distances between demi-hulls and concluded
not to be ignored at high Froude number of 0.5 for
damping of heave motion.
AUTHORS REPLY
Thank you very much to Prof Ohkusu for
your interesting discussion and very useful
observations.

[1] As outlined by Prof. Ohkusu, the local


Froude number based on the distance from the bow,
is generally used to predict the accuracy of the 2D+t
method. This one, in fact, neglects completely the
transverse wave system. In this sense in the figures
4-10 this definition for the local Fr number has been
used.
Just behind the transom stern (fig. 11), in the
region close to the centerline of each hull the
characteristics of the free surface flow (hull hollow
and rooster tail) are strongly influenced by the shape
of the transom stern. In a way we can say that the
flow at the transom including the hullow is a
significant wave generator. This is, of course,
influenced by the flow upstream. However, for this
reason we considered the local Fn based on the
distance from the transom stern.
[2] We agree with the analysis of Prof.
Ohkusu. The nonlinearity is particularly evident on
the free surface elevation, and causes a large
discrepancy of the linear model with respect the
nonlinear one. But it is also evident a better
prediction of the first peak through the 2D+t theory
by taking into account for the interaction between the
hulls with respect to the model accounting for the
interference. Moreover this latter does not seem to
reproduce well the flow downstream, causing a large
shift of the wave elevation with respect the FNL
2D+t model applied to the catamaran configuration.
[3] Concerning the numerical and
experimantal comparison for the RAO, we performed
the new calculations, relativley to the catamaran at
Fn=0.5, with the linear frequency domain code
considering both the steady trim and sinkage and
adding to the hull a false body reproducing the shape
of the steady hull hollow as abtained by 2D+t
simulations. We performed also a convergence
analisys and the results are shown in the figure A.
The numerical prediction seems to be
reasonably grid independent. The effect of the trim
and sinkage, and even more, the effect of the hull
hollow at the transom are large enough, as evident
from the better agreement between numerics and
experiments with respect to ones presented in the
paper. The false body technique, in fact, contributes
to increase the hull lift damping (Faltinsen, 2004),
characterizing the transom flow of a high speed
vessel in waves.
The DM linearization seems still better than
the NK. So the new results confirm the influence of
the steady flow for the unsteady calculations. To this
purpose we observed a larger value of the
contribution due to the double model flow on the
restoring forces of the ship with respect that one
calculated with the NK linearization.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Figure A. Heave (Top) and Pitch RAOs for


the catamaran at Fn=0.5 considered in the paper.
Convergence analisys: Np represents the number of
panels used for the numerical simulations.
Faltinsen (2004), Hydrodynamics
high-speed marine vehicles, Book in press.

of

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Jinzhu Xia
Australian Maritime College, Australia
The disagreement in modern prediction as
shown in Figure 17 may be due to nonlinearity with
respect to ambient wave height rather than steady and
unsteady flow interaction as suggested by Bertram
(1999) for the S175 container ship. In fact, for the
S175, both model tests (ODea et al, 1992, ONR) and
theoretical prediction (Xia, Wand and Jensen, 1998,
in Marine Structures) have confirmed that the
disagreement is very likely due to nonlinearity. I
would like to see some further careful study for the
unsteady case presented in this paper.
As for the results of trim and sinkage in
Tables 3 and 4, it would be interesting to see a
comparison
with
numerical
predictions.
Congratulations for a very valuable contribution!
AUTHORS REPLY
Thank you for your observation that has
been a source for more refined study about the linear
potential seakeeping model.
There are many effects to be considered as
possible error source, for instance:
a) nonlinear effects,
b) local steady and unsteady flow interaction,
c) transom flow,
d) sinkage and trim,
e) hull interaction,
What we have focused on are the points b) and e) in
the paper. Our comments related to the other error
sources are as follow.
Generally speaking, we agree with Dr. Xia
about the role of the nonlinearities. Nevertheless, to
predict how the nonlinearities matter, during our
experiments we considered several regular wave
systems with different steepness.
Further, we
performed also the transient seakeeping test, with a
small steepness of the wave packet. The comparison
with the numerical results, (see fig. 17), has shown
that the nonlinear effects matter but
are not
dominant.
On the other hand, the new numerical study
we recently performed (see fig. A) has shown that a
correct modeling of the transom stern flow in steady
condition, is very important for a more reliable
prediction of the ship motions at high Fn. This latter,
in fact, causes a conspicuous damping due to the
effect of the hull lift (see Faltinsen 2004).
Faltinsen (2004), Hydrodynamics of high-speed
marine vehicles, Book in press.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Twenty-Fifth Symposium on Naval Hydrodynamics


St Johns, Newfoundland and Labrador, Canada, August 813, 2004

Environmental Wave Generation


of High-Speed Marine Vessels
Lawrence J. Doctors (The University of New South Wales, Australia)
Gregory Zilman (Tel-Aviv University, Israel)
Abstract
The work described in this paper covers the
wavemaking characteristics of three types of advanced high-speed marine vessels: catamarans, trimarans, and air-cushion-assisted catamarans.
Experiments on the wavemaking of a model
catamaran are used to demonstrate that the classic
inviscid theory without surface tension accurately
predicts the influence of demihull spacing and water
depth. This is true except in the vicinity of the critical speed, for a waterway of restricted width. In this
case, both the theory and the experiments become
unreliable. Furthermore, it is demonstrated that
the inclusion of surface tension eliminates the unrealistic high-frequency waves that would otherwise
be predicted by the theory at low speeds. The inclusion of viscosity and surface elasticity have marginal
but favorable effects in this respect.
A second set of experiments on a model trimaran are employed to show that the effect of the
stagger of the sidehulls (their longitudinal position)
can be properly predicted by the theory. The influence of the sidehulls on the wavemaking is significant at particular Froude numbers.
Finally, a purely theoretical exercise is undertaken in order to demonstrate that substantial reductions in the wave generation of a catamaran can
be achieved if the vessel is off-loaded by an air cushion supporting approximately 60% of the weight.

1 Introduction
1.1 Previous Work
The matter of wave generation of vessels is of
great importance when considering the operation

in rivers, because of the concern of damage to the


shores. There is now a considerable body of literature on this subject that has been produced over
the last 13 years. An early example was the work
of Doctors, Renilson, Parker, and Hornsby (1991),
in which a total of ten candidate vessels, both catamarans and trimarans, was studied for operation on
the Parramatta River, leading into Sydney Harbor
in Australia. It was demonstrated experimentally
that increasing the spacing between the two demihulls of a catamaran would reduce the height of the
generated waves. This outcome correlated well with
theoretical predictions of the wave resistance, using
the traditional thin-ship theory based on Michell
(1898) and modified for the case of a restricted waterway such as a canal by Sretensky (1936).
It is generally assumed that a smaller wave
height is a desirable requirement for the operation
of river vessels; this is based on the notion that the
waves can cause damage to the banks. However, it
is clear that this may not be true. For example,
a larger-height wave is more likely to break earlier
on approaching the river bank. The consequent remaining smaller-height wave could do less damage.
A second point is that the period of the waves
is a factor. It is commonly understood that a larger
period is more deleterious, as far as erosion of the
river banks is concerned. To this end, Doctors,
Phillips, and Day (2001) examined different multihull configurations. They showed that it was difficult to achieve much improvement in this regard.
In order to apply the abovementioned thinship theory, one must have an adequate model for
the hollow in the water behind the transom, a feature that is characteristic of most high-speed vessels. The earliest example of this was that published
by Molland, Wellicome, and Couser (1994). There,
the hollow was assumed to be a smooth extension
of the hull, with a length equal to 3.0 times the

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Figure 1: Definition of the Problem


(a) Physical Layout

Figure 1: Definition of the Problem


(b) Estimation of Hollow Length

transom beam. Similarly, good results for the wave


resistance were also achieved by Doctors and Day
(1997), using a more sophisticated idealized firehose model that permitted the length to increase
with the speed of the vessel.

minimum) wave elevation, along a longitudinal cut,


decays with respect to the lateral offset from the
track of the vessel.

The purpose of this effort was to avoid a considerably more elaborate approach, such as that of
Doctors and Day (2000a), in which a near-field analysis was developed in order to ensure that the pressure acting on the surface of the hollow is zero (atmospheric). This method is further complicated by
the fact that one must iterate the hollow shape to
achieve this aim. The results of this more sophisticated and intellectually satisfying theory were no
better than the results based on an a priori estimate, described above.
Another effort in this area was that of Doctors
and Zilman (2004). They showed that a transomstern model based on the classic backward-facingstep problem, such as the one analyzed by Hall
(2001), worked remarkably well. This latest model
assumed the factor 3.340 for the length of the hollow (slightly larger than the factor noted above).
Additionally, a positive or negative convergence of
the stern of the vessel, either in plan or in profile,
was accounted for by considering the shape of the
hollow and matching it to the stern. For example,
a positively converging stern would create a shorter
hollow.
The matter of the rate of decay of the wave
pattern has been the subject of much debate over
the last decade. In particular, one is interested in
the rate at which the envelope of the maximum (or

In the literature, it is frequently (and wrongly)


quoted that the exponent N in the formula:
/1

A(y/y1 )N

(1)

is 1/3. Here, is the wave elevation as a function


of the lateral offset y and 1 is the value at the first
offset y1 . The values of A and N are obtained from a
best-fit analysis. The value of 1/3 probably results
from a misunderstanding of the work of Wehausen
and Laitone (1960, p. 487, Equation (13.42b)) and
of Stoker (1966, p. 242, Equation (8.2.40)). Those
equations are only applicable for the variation along
the Kelvin angle in the linearized far-field wave system generated by a point source traveling in deep
water.
Doctors and Day (2001) demonstrated theoretically that, in a realistic situation, the exponent
could typically vary between 1.06 and 0.20, depending upon the speed of the vessel. At high values of the Froude number, it can be shown that the
exponent approaches the ideal value 1/2.
In an attempt to reduce the wave generation
of marine passenger vessels, many researchers have
studied less-conventional configurations.
Thus,
Doctors and Day (2000b) performed calculations on
an air-cushion vehicle (ACV) in which the planform
was subdivided into different arrangements of subcushions, each with its own pressure. Considerable
reductions in wave generation can theoretically be
achieved in this manner. Day and Doctors (2001)

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Figure 2: Numerical Aspects


(a) Fitting Mesh to the Vessel

Figure 2: Numerical Aspects


(b) Dispersion Relationship

evaluated the performance of different craft and


demonstrated that the ACV and the surface-effect
ship (SES) performed better than a traditional vessel in this regard.

of the hull design of a potential vessel. Examples


of such efforts are those of Degiuli, Werner, and
Doliner (2003), Heimann, and Harries (2003), and
Koushan (2003).

Doctors (2003b) considered a series of SES designs, based on a total displacement of 60 t, in which
different proportions of the total weight were carried by the sidehulls and the cushion. In the case
of deep water, a reduction of 54% is possible at a
speed of 40 knots, when the cushion supports 60%
of the weight. Under the same conditions, in water
of depth 7.5 m, the expected reduction is 59%.

Raven (2000) has presented ship-wave calculations based on nonlinear inviscid theory using the
panel method. Brizzolara and Bruzzone (2003) and
Janson, Leer-Andersen, and Larsson (2003) have
implemented a novel combined method. The near
field is computed on the basis of a fully nonlinear
method, while the far field is computed using the
linear method. This technique should prove to be
efficient and is justified because the waves in the far
field have a lower amplitude.

An example of another unusual configuration


is arrow trimarans which were analyzed by Day,
Clelland, and Nixon (2003). These researchers
showed theoretically and experimentally that an appropriate choice of the arrangement of the subhulls
could result in a lower wave generation.
Doctors (2003a) carried out an exhaustive experimental investigation on a model catamaran, in
which the demihull spacing, the water depth in the
ocean basin, and the model speed were all varied.
Corresponding longitudinal wave profiles were measured and compared with the predictions of inviscid theory for approximately 156 combinations of
the physical setup. Generally, excellent correlation between the experiments and the predictions
for the wave profiles was achieved. Similarly, comparisons of root-mean-square wave elevation were
equally promising.
Clearly, an ultimate goal of this research, to
which this paper contributes, is the optimization

For these nonlinear methods, there is the serious matter of whether one can achieve both a numerically converged solution for each test geometric
configuration and also a sufficient number of iterations to obtain an optimized result in a reasonable timeframe.
The idea of predicting the wave system in a totally realistic waterway (one with an arbitrary cross
section) is attractive. Work in this area has been reported by Henn, Sharma, and Jiang (2001) and by
Hong and Mutsuda (2002). Doctors and Renilson
(1993) also studied the impact of a non-rectangular
waterway in the towing tank by the use of flat banks
with a constant slope. It was shown that the formula for the wave resistance applicable to a rectangular channel could still be used with reasonable accuracy by employing in this formula the same water
depth together with an effective width that provides
the same cross-sectional area of the channel.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Figure 3: Subject Vessels


(a) Modified Series 64 Demihull

An interesting recent development has been


the introduction of some real-fluid phenomena into
the classic inviscid linearized theory. Thus, Tuck,
Scullen, and Lazauskas (2000 and 2002) showed that
the assumption of low viscosity allowed one to estimate the rate of viscous decay of a wave system
(once the wave system had been created on an inviscid basis).
Doctors (2003a) followed this approach extensively by comparing the predictions of the wave system with experimental results for a small model
catamaran tested in a towing tank. This work was
extended by Doctors and Zilman (2004), who also
incorporated the effect of surface tension at the free
surface. It was shown that both viscosity and surface tension tend to reduce the peaks in the wave
system, although the effects are relatively small,
even for the 1.5 m-long model at low speeds, for
distances up to 10.5 m behind the stern.

1.2 Refinements to the Theory

In the current investigation, the following theoretical considerations will now be taken into account: (a) the length of the hollow that is generated behind the transom stern of the vessel, (b) the
surface tension of the water, (c) the viscosity of the
water, and (d) the presence of surfactants (modeled
by the inclusion of surface elasticity). The last three
aspects are likely to be more relevant at model scale
and might cast some doubt on extrapolation of tests
for wave generation conducted in an ocean basin at
relatively low speeds.

Figure 3: Subject Vessels


(b) AMECRC Model 9 Trimaran

1.3 Advanced Marine Vehicles


In the later sections of the paper there are
some computer predictions of the wave system generated by different types of advanced marine vehicles. These types will include catamarans, trimarans, hovercraft, and surface-effect ships. In this
way, it will be possible to assess the relative performance of these vehicles with respect to their environmental impact.

2 Theory
2.1 Definition of the Problem
Figure 1 provides a definition of the physical
problem, indicating the principal dimensions of the
vessel and its hollow. In the case of a catamaran,
model tests can be effected on just one demihull, as
illustrated in Figure 1(a). The vessel length L, the
effective spacing between the demihulls s, and the
effective width of the channel w are indicated.
The hollow behind the transom is estimated on
the basis of the flow behind a backward-facing step,
described in some detail by Hall (2001). This is indicated in Figure 1(b). The shape of the hollow is
known to be roughly elliptic; this knowledge can be
employed to fit the geometry of the hollow so that
the effective surface for the hull-hollow combination
is smooth and does not suffer any discontinuity in
slope. In fact, the computer program estimates both
the profile and the plan of the hollow, with the maximum of the two values of the length assumed.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Item
Displacement mass
Waterline length
Waterline beam
Draft
Waterplane-area coef.
Maximum section coef.
Block coefficient
Prismatic coefficient
Slenderness coefficient

Symbol

L
B
T
CW P
CM
CB
CP
L/1/3

Value
3.745 kg
1.500 m
0.1113 m
0.04623 m
0.7866
0.7292
0.4860
0.6665
9.654

Table 1: Particulars of Series 64 Hull

Item
Overall length
Effective width
Water depth

Value
60.0 m
7.100 m
0.300 m
1.500 m

Table 2: Particulars of Towing Tank

in which kx and ky are the wave numbers in the


longitudinal and transverse directions. Thus, we
can write the transformed potential as
=

2.2 Potential-Flow Theory


The equation of continuity for the perturbation potential ,

Symbol
l
w
d

Q
exp(k|z z 0 |)
4k
Q

exp(k|z + z 0 + 2d|)
4k
+A(k, ) cosh[k(z + d)] ,

(7)

(2)

where k is the circular wave number and is the


wave angle.

must be satisfied throughout the domain of the water, together with the kinematic free-surface condition,

Finally, we have the downstream potential due


to a point source:

xx + yy + zz

z + U x

0,

0,

(3)

and the dynamic free-surface condition,


g U x (xx + yy )

0.

(4)

Here, U is the speed of the vessel, g is the acceleration due to gravity, is the density of the water, and
is the surface tension. This equation was used, for
example, by Webster (1966). On the bottom of the
channel, we have:
z

0 on z = d ,

(5)

in which d is the depth of the water.


The solution for the potential can be obtained
in the standard way by assuming the form,
Q
Q

+,
(6)
4r 4r0
where r and r0 are the radial distances,
p
(x x0 )2 + (y y 0 )2 + (z z 0 )2
r =
p
0
(x x0 )2 + (y y 0 )2 + (z + z 0 + 2d)2 .
r =

Next we utilize the double-Fourier-transform,


=

1
2
1
2

Z
Z
(x, y) exp[i(kx x + ky y)] dxdy

Z
Z

x , ky ) exp[i(kx x + ky y)] dkx dky ,


(k

2Q

Z
0

2
X
0
3
dky
k0 k +
k
U 2
j=1

cosh[k(z 0 + d)]
cosh2 (kd)
sin[kx (x x0 )] cos[ky (y y 0 )]

. df
.
(8)
(1)j kx
dk

cosh[k(z + d)]

This equation makes use of the gradient of the


dispersion relationship (without viscosity or surface
elasticity), given below by Equation (25). The relationship between the two forms of the dispersion
function is just
f ( = 0, = 0)

(U )2 f .

(9)

2.3 Downstream Wave Elevation without


Viscosity or Surface Elasticity
The complete vessel can then be represented
as a sum of such source points, from which we can
write the downstream wave elevation as

2
. df
4i X X 0
j

(1) kx k
x =
w i=0 j=1
dk
exp(ikx x) cos(ky y) (U iV) , (10)

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Item
Demihull spacing

Symbol
s

First wave-probe offset


Wave-probe spacing
Last wave-probe offset
Wave-cut length

Value
0.300 m
0.400 m
0.500 m
1.000 m
0.500 m
3.000 m
10.500 m

y1
y
y2
Lcut

Table 3: Experiments in Towing Tank

Item
Displacement mass
Waterline length
Waterline beam
Draft
Waterplane-area coef.
Maximum section coef.
Block coefficient
Prismatic coefficient
Slenderness coefficient

Symbol

L
B
T
CW P
CM
CB
CP
L/1/3

Value
12.804 kg
1.596 m
0.2002 m
0.08047 m
0.7958
0.7996
0.4990
0.6241
6.817

Table 4: Particulars of AMECRC Model 9


where

1
2

for i = 0
for i 1

(11)

In this equation, A0 is the area over which the cushion pressure p(x, y) is defined.
2.5 Damping due to Viscosity and Surface
Elasticity

kx

q
k 2 ky 2

(12)

ky

2i/w ,

(13)

and
k0

g/U 2 .

(14)

The finite-depth wave functions are:


U
V

= [P + + exp(2kd)P ]/[1 + exp(2kd)]


(15)
= [Q+ + exp(2kd)Q ]/[1 + exp(2kd)] ,

which can be expressed in terms of the standard


deep-water wave functions:
Z
P + iQ =
b(x, z)
S0

exp[i(kx x + ky y) kz) dxdz . (16)


The vessel is defined by means of the local beam
b(x, z) over the area of its centerplane/s S0 .
2.4 Contribution from Air Cushion

The Fourier transform of the modified combined free-surface condition (the dispersion relationship) was published for the case of deep water, by
Zilman and Miloh (2001, p. 161, Equation (61)),
and for the case of water of finite depth, by Hansen
and Ahmad (1971, p. 30, Equation (82)). We shall
base our work here on the latter source, even though
the difference in the numerical results would be
slight. This is because only relatively short waves
are damped and these would be classified as deepwater waves:
f

= [ 2 (gk + k 3 ) tanh(kd)]( 2 mk 2 )
k 3 (gk + k 3 ) + 4i 3 k 2
+ 42 2 k 3 [m tanh(kd) k]
= 0.

We now consider the two finite-depth wave


functions in Equation (15), which, for an SES, can
be calculated by adding the wave functions from the
pressure distribution and the two demihulls.
The results from Doctors (1992), for example,
give the following contributions from the cushion:
ZZ
U + iV =
[p(x, y)/g]
A0

exp[i(kx x + ky y)] dxdy .

We now extend the above analysis by including


the effect of the molecular viscosity of the water
and the elasticity of the surfactant . Both of these
are considered to be small in the current work.

(17)

(18)
(19)

Equation (19) is equivalent to that given by Wehausen and Laitone (1960, p. 632, Equation (24.3)),
except for the additional terms involving the viscosity and the surface elasticity. These terms introduce
an imaginary component to the wave number of the
free-surface waves and this produces a spatial damping factor.
The additional symbols in Equation (18) represent the circular frequency of the wave:

Copyright National Academy of Sciences. All rights reserved.

U kx

(20)

Twenty-Fifth Symposium on Naval Hydrodynamics

Figure 4: Effect of , and on Wave


Profiles (a) d/L = 0.2 and F = 0.2028

Figure 4: Effect of , and on Wave


Profiles (b) d/L = 0.2 and F = 0.6024

Figure 4: Effect of , and on Wave


Profiles (c) d/L = 1.0 and F = 0.2026

Figure 4: Effect of , and on Wave


Profiles (d) d/L = 1.0 and F = 0.6024

and the complex wave number in the vertical direction for the stream function which describes the
(small) rotational component of the flow:
p
m =
(21)
k 2 i/ .

For this purpose, we can write


f

= k 3 (gk + k 3 ) + 4i 3 k 2
+ 42 2 k 3 [m tanh(kd) k]

(24)

and
We can now adapt the analysis of Hansen and
Ahmad (1971, p. 22, Equation (46)), in which we
consider a first-order correction to the value of the
wave number. The complex wave number is:
k

k + i ,

df
dk

= [2 2 /k (g + 3 k 2 ) tanh(kd)
(gk + k 3 )d sech2 (kd)]
( 2 mk 2 ) .

(25)

(22)

where k is the standard (inviscid) solution of Equation (19). We substitute this definition into Equation (19) and apply just one iteration of Newtons
method. Hence:

df
= Im f /
.
(23)
dk

The viscous damping factor that is to be included for each component of the wave spectrum is
then
V

= exp[||(|x | cos + |y | sin )] . (26)

The offsets x and y are the distances from the

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Figure 5: Effect of , and on RMS Wave


Elevation (a) s/L = 0.2000 and d/L = 1.0

Figure 5: Effect of , and on RMS Wave


Elevation (b) s/L = 0.2667 and d/L = 1.0

3.2 Solution of the Dispersion Relationship

The dispersion relationship, Equation (19),


can be solved in a straightforward manner by using
an initial estimate for the wave number k, namely
the value for deep-water and without surface tension, viscosity, or surface elasticity. The precise solution can be found by means of a standard NewtonRaphson iteration, using the gradient from Equation (25).

Figure 5: Effect of , and on RMS Wave


Elevation (c) s/L = 0.3333 and d/L = 1.0

Some understanding of the nature of the function to be solved can be derived from an examination of Figure 2(b).

3.3 Recursion Formulas


source point (approximated by the center of buoyancy) to the field point.
A very efficient recursion algorithm is employed in the computer program for evaluating the
wave elevation of the free surface, expressed by
Equation (10).

3 Numerical Aspects
3.1 Discretization of the Hull
Figure 2(a) shows how the centerplane paneling is used to represent the hull and the hollow in
the water behind the transom stern. The panels
are, in effect, tent functions, which overlap. They
have been employed previously by Doctors and Day
(1997), for example. Typically, one requires only
about 60 tents in the longitudinal direction and 20
tents in the vertical direction in order to obtain converged results.

The recursion algorithm takes advantage of the


fact that the vast majority of the calculation is common for all the points on the free surface and that
the intermediate results can be stored for repetitive
use.
As can be seen, the dependence on position
(x, y) is confined to two simple factors, which differ by just one constant complex multiplication for
field points equally spaced either longitudinally or
transversely.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Figure 6: Effect of Demihull Separation on


RMS Wave Elevation (a) d/L = 0.2

3.4 Upper Limit of the Wave Number


Summations
The summations in Equation (10) theoretically involve an infinite number of terms. Some
experimental calculations have demonstrated that
one can typically choose a cutoff value for the maximum value of the transverse number given by ky =
3000/L, where L is the length of the vessel. In this
way, the resulting accuracy in the wave elevation
for the case of inviscid flow is three decimal places
or better. When viscosity is included, lower cutoff
values of ky may be chosen.

4 Catamarans
4.1 Experiments
The subject vessel is depicted in Figure 3(a).
It is a slightly modified version of the original Series 64 hull, which was described by Yeh (1965). Its
hydrostatic particulars are listed in Table 1.
The dimensions of the towing tank are presented in Table 2 and the experimental arrangements are shown in Table 3.
4.2 Surface Tension, Viscosity, and Surface
Elasticity
Figure 4 shows a comparison of the experiment
and theory for the wave profile along one longitudinal cut for which y/L = 1.333. It is noted, for

Figure 6: Effect of Demihull Separation on


RMS Wave Elevation (b) d/L = 1.0

the low Froude number of 0.2028 in Figure 4(a),


that the inclusion of the effect of surface tension
is to reduce the amplitude of the waves (a desirable
effect).
It is observed that the most sophisticated theories, which also include the molecular viscosity
(not the much larger eddy viscosity) and the surface elasticity , marginally reduces the magnitude
of the waves, providing a slightly better correlation
between theory and experiment.
Figure 4(b) shows the equivalent results, but
for the much higher Froude number of 0.6024. In
this case, the influences of surface tension, molecular viscosity and surface elasticity are quite negligible.
Figure 4(c) and Figure 4(d) present similar calculations for the deep-water case, for which
d/L = 1.
The root-mean-square wave elevation RMS
over all five longitudinal wave cuts is plotted in Figure 5 for the deepest condition of d/L = 1.0. The
three different demihull spacings s are depicted in
the three parts of this figure. The low-speed range of
the data only is shown. This allows one to see more
clearly that the influence of surface tension, viscosity, and surface elasticity is to reduce the numerical
value of the predictions. However, the effects are
quite unimportant with regard to the root-meansquare wave elevation.
It is encouraging to note that the theoretical
oscillations in the curve are verified by the theory.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Item
Overall length
Width
Water depth

Symbol
l
w
d

Value
35.0 m
12.0 m
0.450 m
0.900 m

Table 5: Particulars of Ocean Basin

However, the phasing of these oscillations is imperfect. It is thought that this difference may be related to the precise geometry of the transom hollow,
which affects the equivalent wavemaking length of
the vessel.
4.3 Demihull Spacing
Finally, Figure 6 presents the effect of demihull spacing s on the root-mean-square wave elevation for the two different water depths. It is seen
that increasing the spacing leads to a lower wave
generation.
The experimental phenomena are predicted
accurately by the current enhanced theory, with the
exception already noted by Doctors (2003a). This
is the problem of vessel speeds corresponding to a
depth Froude number of unity, where the theory behaves poorly, if the width of the waterway w is also
finite.

5 Trimarans
5.1 Experiments
The subject vessel is depicted in Figure 3(b).
The centerhull is Model 9 of the AMECRC series,
which was described by Bojovic and Goetz (1996).
Its hydrostatic particulars are listed in Table 4. The
sidehulls were constructed as geosyms of the centerhull, with the linear scale factor 0.4590.
The dimensions of the ocean basin are presented in Table 5 and the experimental arrangements are shown in Table 6.
The experimental setup permitted the stagger
(the longitudinal position of the transom of the sidehulls with respect to the centerhull) to be adjusted.
Three values of the stagger were selected.

Item
Sidehull stagger

Symbol
r

First wave-probe offset


Wave-probe spacing
Last wave-probe offset
Wave-cut length

y1
y
y2
Lcut

Value
0.640 m
0.160 m
0.320 m
1.000 m
0.500 m
4.500 m
8.000 m

Table 6: Experiments in Ocean Basin

5.2 Correlation of Results


Figure 7 depicts the experimental and theoretical results for the forwardmost stagger of the
sidehulls (r/L = 0.2) at the two abovementioned
water depths. The first set of experimental data
(indicated by the circles) shows that the theory and
the experiments follow the same trends. However,
the numerical agreement is certainly not as pleasing
as that for the case of the catamaran investigation
already discussed above.
It is thought that the explanation for the
poorer correlation in the case of the trimaran is that
the tests were conducted in the ocean basin, which
is much shorter than the towing tank used for the
catamaran tests. It is believed that the wave pattern may not have reached a steady-state situation.
As a result, one could reasonably argue that each
experimental data point is likely to: (a) correspond
to an earlier steady-state Froude number and (b) be
of lower magnitude. This thinking is in accordance
with the theoretical work of Doctors and Sharma
(1972), who demonstrated the identical trends for
the wave resistance of an accelerating ACV.
For the purpose of the current effort, we have
used a factor of 0.95 on the abscissa (the Froude
number) and a factor of 1.10 on the ordinate (the
RMS wave elevation) for the re-plotting of the experiments (indicated by the squares) in the two
parts of Figure 7. As a consequence, it is seen that
the correlation of experiment and theory is much
improved, lending credibility to this line of reasoning.
5.3 Sidehull Stagger
All three sidehull staggers for the experiments
are considered in Figure 8. The strong effects due

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Item
Displacement mass
Waterline length
Waterline beam
Draft
Sidehull spacing
Waterplane-area coef.
Maximum section coef.
Block coefficient
Prismatic coefficient
Slenderness coefficient

Symbol

L
B1
T
s
CW P
CM
CB
CP
L/1/3

Value
30 t
30 m
1.000 m
1.500 m
10 m
0.8333
0.8000
0.6667
0.8333
9.655

Vessel
Type

Sidehull
Beam
B1
(m)

Catamaran
SES
SES
SES
SES
Hovercraft

1.000
0.800
0.600
0.400
0.200
0.000

Cushion
Depression

(m)
0.000
0.040
0.080
0.120
0.160
0.200

Total
Displacement

(t)
60
60
60
60
60
60

Table 7: Particulars of Parent Sidehull

Table 8: Particulars of the Six Vessels

to the stagger of the sidehulls is evident. That is,


the interleaving nature of the curves with respect
to the Froude number is evident. Thus, one can
clearly observe that each sidehull stagger has its own
optimal speed of operation.

A family of surface-effect ships was created by


considering six different values of the vessel sidehull beam. The consequent loss in displacement was
compensated for by means of the cushion pressure.
This is detailed in Table 8. The new sidehulls were
created from the parent sidehull by means of a simple affine transformation, as described by Doctors
(1995).

Once again, the unrealistically high values of


the predictions for the root-mean-square wave elevation near the critical speed are evident. Of course,
it is also true that the experimental data is very suspect in this region, because the ocean basin would
certainly not have been sufficiently long in order to
achieve steady-state data for the wave elevation.
Comparison of the two parts of this figure
demonstrate the influence of the water depth, which
are predicted well by the linearized theory.

6 Surface-Effect Ships
6.1 Parent Sidehull

The chosen parent sidehull can be seen in Figure 9(a). For the sake of simplicity, this was based
on the classic Wigley (1934) hull. Specifically, the
bow half of the sidehull possesses parabolic waterlines and its sections are also parabolic, but with
some wall-sided portion below the waterline. The
stern half of the sidehull possesses parallel waterlines and its sections blend in with the sections of
the bow.
The stern, having a transom, is typical of the
SES, because this permits the traditional waterjet
installation. The geometric data of the parent sidehull is given in Table 7.

6.2 Wave-Decay Curves


Different measures of the wave system were
considered in this work. In this section, we shall
concentrate on the wave range range . This is simply the difference between the highest and the lowest points in the wave cut.
Figure 10 shows the transverse decay of these
wave measures for the pure catamaran. Predictions
are shown for speeds of 20 knots and 40 knots. A
regression curve of the type given by Equation (1)
has also been fitted; it is seen that the fit is excellent
for these cases.
6.3 Wave-Decay Coefficients
Now we present in Figure 11 the two decay
coefficients in Equation (1). The fitted wave height
1 for the six vessels is plotted as a function of the
speed in Figure 11(a). This is a condition of the
vessel traveling in water of depth 7.5 m. It is most
interesting to observe that the optimal craft appears
to be the SES in which the sidehull beam B1 is
0.4 m. At the high end of the speed range, the
reduction in the wave height can be more than 50%
in comparison with the catamaran.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Figure 7: Correlation of RMS Wave


Elevation (a) d/L = 0.2813 and r/L = 0.2

Figure 7: Correlation of RMS Wave


Elevation (b) d/L = 0.5625 and r/L = 0.2

Figure 8: Effect of Sidehull Stagger on RMS


Wave Elevation (a) d/L = 0.2813

Figure 8: Effect of Sidehull Stagger on RMS


Wave Elevation (b) d/L = 0.5625

Care must be taken when drawing conclusions,


because the exponent N also varies from one craft to
the other, as seen in Figure 11(b). It must be borne
in mind that we desire the largest negative value of
this exponent. At high speeds, on the other hand,
it can be seen that all the vessels possess the same
value of the decay exponent, namely 1/2.

7 Concluding Remarks

The case of infinite water depth is presented


in Figure 12. Two points can be readily seen in
Figure 12(a). The first point is the now lesser magnitude of the wave generation which occurs at the
hump speed of around 21 knots. Secondly, the relative gain enjoyed by the SES (with a sidehull beam
of 0.4 m) is now perhaps less dramatic at the highest speeds. The approach of the decay exponent N
to 1/2 at high speeds in Figure 12(b) is not as
pronounced in this case of infinite-depth water.

The work presented in this paper has shown


that:
1. The general ability of classic linearized inviscid
wave theory to predict the wavemaking characteristics of catamarans and trimarans traveling
in water of different depths. The influences of
the demihull spacing for the catamaran and the
sidehull stagger for the trimaran are also properly predicted.
2. The effect of introducing surface tension into
the calculations is to reduce some of the peaks
in the wave profiles giving better correlation
with the experiments. However, the impact

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Figure 9: Subject Surface-Effect Ship


(a) Layout of the Vessel

of this improvement is quite marginal with regard to overall wave measures, such as the rootmean-square wave elevation.
3. The effect of including molecular viscosity and
surface elasticity into the calculations is miniscule, even at the very small model scale and at
the lowest Froude numbers considered in this
work. Nevertheless, the very small influence is
favorable as it further reduces unrealistic peaks
in the predictions of the wave profiles.
4. The strong theoretical oscillations in the curves
of root-mean-square wave elevation at low
speeds have been verified by extensive experiments documented here.
5. An air-cushion assisted catamaran generates a
much milder wave system than a traditional
catamaran. The calculations have shown that
a craft in which 40% of the weight is carried by
the displacement and 60% is carried by the air
cushion produces a configuration that is nearly
optimal over the entire speed range. The reduction in wave generation can be around 54%
in the case of deep water and up to 59% in the
case of a water depth of 7.5 m for the 60 t vessel
operating at 40 knots.
Research work planned for the future includes
the following:
1. Repeating the experiments on the model trimaran in a longer test facility, such as a towing
tank, with the aim of achieving steady-state
wave data.

Figure 9: Subject Surface-Effect Ship


(b) Modified Wigley SES Sidehull

2. Conducting experiments, both in deep water


and in water of finite depth, on a model SES
in order to verify the theoretical advantages of
this vessel with respect to wavemaking.
3. Enhancing the transom-stern model with the
intention of improving the estimation of the
length of the hollow, by means of an investigation using computational fluid dynamics
(CFD). This could be achieved by running a
number of different transom-stern configurations in order to determine a better estimate
of the hollow length.

8 Acknowledgments
The model tests were performed in the Towing Tank and the Ocean Basin at the Australian
Maritime College (AMC) by Mr Olav Opheim and
Mr Stephen Helmstedt, students at UNSW, under the supervision of Mr Gregor Macfarlane and
Mr Richard Young of the AMC.
Dr Stephen Hall of UNSW provided much useful theoretical advice regarding the shape of the
transom-stern hollow.
The authors would like to acknowledge the assistance of the Australian Research Council (ARC)
Discovery-Project Grant Scheme (via Grant Number DP0209656).
The in-kind support of this work by The University of New South Wales and Tel-Aviv University
is also greatly appreciated.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Figure 10: Curves of Wave Decay for


Catamaran (a) Speed of 20 Knots

9 References
Bojovic, P. and Goetz, G.: Geometry of AMECRC Systematic Series, Australian Maritime
Engineering Cooperative Research Centre, Report AMECRC IR 96/6, 53+v pp (1996)
Brizzolara, S. and Bruzzone, D.: Near and
Distant Waves of Fast Ships in Unlimited
and Limited Depths, Proc. Seventh International Conference on Fast Sea Transportation
(FAST 03), Ischia, Italy, Vol. 3, pp H.1H.12
(October 2003)
Day, A., Clelland, D., Nixon, E.: Experimental and Numerical Investigation of Arrow Trimarans, Proc. Seventh International Conference on Fast Sea Transportation (FAST 03),
Ischia, Italy, Vol. 3, pp D2.23D2.30 (October
2003)
Day, A.H. and Doctors, L.J.: Concept Evaluation for High-Speed Low-Wash Vessels,
Proc. Sixth International Conference on Fast
Sea Transportation (FAST 01), Royal Institution of Naval Architects, Southampton, England, Vol. 1, pp 121133 (September 2001)
Degiuli, N., Werner, A., and Doliner, Z.:
Determination of Optimum Position of Outriggers of Trimaran Regarding Minimum Wave
Pattern Resistance, Proc. Seventh International Conference on Fast Sea Transportation
(FAST 03), Ischia, Italy, Vol. 3, pp D2.15
D2.22 (October 2003)
Doctors, L.J.: The Use of Pressure Distributions to Model the Hydrodynamics of AirCushion Vehicles and Surface-Effect Ships,

Figure 10: Curves of Wave Decay for


Catamaran (b) Speed of 40 Knots

Proc. High-Performance Marine Vehicle Conference and Exhibit (HPMV 92), American Society of Naval Engineers, Washington,
pp SES56SES72 (June 1992)
Doctors, L.J.: A Versatile Hull-Generator Program, Proc. Twenty-First Century Shipping
Symposium, University of New South Wales,
Sydney, New South Wales, pp 140158, Discussion: 158159 (November 1995)
Doctors, L.J.: The Influence of Viscosity on the
Wavemaking of a Model Catamaran, Proc.
Eighteenth International Workshop on Water
Waves and Floating Bodies (18 IWWWFB),
Le Croisic, France, pp 12-1124 (April 2003)
Doctors, L.J.: A Comparison of the Environmental Wave Generation of Hovercraft and
other High-Speed Vessels, Proc. Fifth AirCushion Technology Conference and Exhibition, Hovercraft Museum, Chark Lane, Leeon-the-Solent, England, 13 pp (October 2003)
Doctors, L.J. and Day, A.H.: Resistance
Prediction for Transom-Stern Vessels, Proc.
Fourth International Conference on Fast Sea
Transportation (FAST 97), Sydney, Australia, Vol. 2, pp 743750 (July 1997)
Doctors, L.J. and Day, A.H.: Steady-State
Hydrodynamics of High-Speed Vessels with a
Transom Stern, Proc. Twenty-Third Symposium on Naval Hydrodynamics, Val de Reuil,
France, pp 12-112-14, Discussion: p 12-15
(September 2000)
Doctors, L.J. and Day, A.H.: Wave-Free
River-Based Air-Cushion Vehicles, Proc. In-

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Figure 11: Decay Coefficients in Finite Depth


(a) Magnitude of Range

Figure 11: Decay Coefficients in Finite Depth


(b) Decay Exponent

Figure 12: Decay Coefficients in Deep Water


(a) Magnitude of Range

Figure 12: Decay Coefficients in Deep Water


(b) Decay Exponent

ternational Conference on Hydrodynamics of


High-Speed Craft: Wake Wash and Motions
Control, Royal Institution of Naval Architects,
London, England, pp 12.112.9 (November
2000)

Influence of Demihull Separation and River


Banks on the Resistance of a Catamaran,
Proc. Second International Conference on
Fast Sea Transportation (FAST 93), Yokohama, Japan, Vol. 2, pp 12311244 (December
1993)

Doctors, L.J. and Day, A.H.: The Generation and Decay of Waves behind High-Speed
Vessels, Proc. Sixteenth International Workshop on Water Waves and Floating Bodies
(16 IWWWFB), Hiroshima, Japan, pp 3336
(April 2001)
Doctors, L.J., Phillips, S.J., and Day, A.H.:
Focussing the Wave-Wake System of a HighSpeed Marine Ferry, Proc. Sixth International Conference on Fast Sea Transportation
(FAST 01), Royal Institution of Naval Architects, Southampton, England, Vol. 1, pp 97
106 (September 2001)
Doctors, L.J. and Renilson, M.R.:

The

Doctors, L.J, Renilson, M.R., Parker, G.,


and Hornsby, N.: Waves and Wave Resistance of a High-Speed River Catamaran,
Proc. First International Conference on Fast
Sea Transportation (FAST 91), Norwegian
Institute of Technology, Trondheim, Norway,
Vol. 1, pp 3552 (June 1991)
Doctors, L.J. and Sharma, S.D.: The Wave
Resistance of an Air-Cushion Vehicle in Steady
and Accelerated Motion, J. Ship Research,
Vol. 16, No. 4, pp 248260 (December 1972)
Doctors, L.J. and Zilman, G.: The Influence of Surface Tension and Viscosity on the

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Wavemaking of a Model Catamaran, Proc.


Nineteenth International Workshop on Water
Waves and Floating Bodies (19 IWWWFB),
Cortona, Italy, pp 11.111.4 (March 2004)
Hall, S.D.: An Investigation of the Turbulent
Backward Facing Step Flow with the Addition of a Charged Particle Phase and Electrostatic Forces, University of New South Wales,
School of Mechanical and Manufacturing Engineering, Doctoral thesis, 303+xv pp (February
2001)
Hansen, R.S. and Ahmad, J.: Waves at Interfaces, Progress in Surface and Membrane Science, Academic Press, New York, Vol. 4, pp 1
56 (1971)
Heimann, J. and Harries, S.: Optimization of
the Wave-Making Characteristics of Fast Ferries, Proc. Seventh International Conference
on Fast Sea Transportation (FAST 03), Ischia, Italy, Vol. 3, pp D1.37D1.44 (October
2003)
Henn, R., Sharma, S.D., and Jiang, T.: Influence of Canal Topography on Ship Waves in
Shallow Water, Proc. Sixteenth International
Workshop on Water Waves and Floating Bodies (16 IWWWFB), Hiroshima, Japan, pp 49
52 (April 2001)
Hong, C.B. and Mutsuda, H.: Numerical
Study on the Effect of Waterway Shape on
Ship Wave Characteristics, J. Society of
Naval Architects of Japan, Vol. 192, pp 8187
(December 2002)
Janson, C.-E., Leer-Andersen, M., and Larsson, L.: Calculation of Deep-Water Wash
Waves Using a Combined Rankine/Kelvin
Source Method, J. Ship Research, Vol. 47,
No. 4, pp 313326 (December 2003)
Koushan, K.: Automatic Hull Form Optimisation towards Lower Resistance and Wash
Using Artificial Intelligence, Proc. Seventh
International Conference on Fast Sea Transportation (FAST 03), Ischia, Italy, Vol. 3,
pp H.41H.50 (October 2003)
Michell, J.H.: The Wave Resistance of a Ship,
Philosophical Magazine, London, Series 5,
Vol. 45, pp 106123 (1898)
Molland, A.F., Wellicome, J.F., and
Couser, P.R.: Theoretical Prediction of
the Wave Resistance of Slender Hull Forms

in Catamaran Configurations, University of


Southampton, Department of Ship Science,
Report 72, 24+i pp (March 1994)
Raven, H.C.: Numerical Wash Prediction using a Free-Surface Panel Code, Proc. International Conference on Hydrodynamics of
High-Speed Craft: Wake Wash and Motions
Control, Royal Institution of Naval Architects,
London, England, pp 10.110.12 (November
2000)
Sretensky, L.N.: On the Wave-Making Resistance of a Ship Moving along in a Canal,
Philosophical Magazine, Series 7, Supplement,
Vol. 22, No. 150, pp 10051013 (November
1936)
Stoker, J.J.: Water Waves, Interscience Publishers, Inc., New York, 567+xxviii pp (1966)
Tuck, E.O., Scullen, D.C., and Lazauskas,
L.: Ship-Wave Patterns in the Spirit of
Michell, Proc. IUTAM Symposium on FreeSurface Flows, Birmingham, England, 8 pp
(July 2000)
Tuck, E.O., Scullen, D.C., and Lazauskas,
L.: Wave Patterns and Minimum Wave
Resistance for High-Speed Vessels, Proc.
Twenty-Fourth Symposium on Naval Hydrodynamics, Fukuoka, Japan, Vol. 2, pp 150165
(July 2002)
Webster, W.C.: The Effect of Surface Tension
on Ship Wave Resistance, University of California, College of Engineering, Berkeley, Report NA-66-6, 114+iii pp (July 1966)
Wehausen, J.V. and Laitone, E.V.: Surface
Waves, Encyclopedia of Physics: Fluid Dynamics III, Ed. by S. Fl
ugge, Springer-Verlag,
Berlin, Vol. 9, pp 445814 (1960)
Wigley, W.C.S.: A Comparison of Experiment and Calculated Wave-Profiles and WaveResistances for a Form Having Parabolic Waterlines, Proc. Royal Society of London, Series A, Vol. 144, No. 851, pp 144159 + 4 plates
(March 1934)
Yeh, H.Y.H.: Series 64 Resistance Experiments
on High-Speed Displacement Forms, Marine
Technology, Vol. 2, No. 3, pp 248272 (July
1965)
Zilman, G. and Miloh, T.: Kelvin and V-Like
Wakes Affected by Surfactants, J. Ship Research, Vol. 45, No. 2, pp 150163 (June 2001)

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION

relatively small number of longitudinal cuts, and may


not be highly reliable.

Dr A.H. Day
Universities of Glasgow and Strathclyde, Scotland
I would like to congratulate the authors on
an interesting study. I was particularly interested to
see the effect of the inclusion of surface elasticity on
the mathematical model.
With regard to the inclusion of molecular
viscosity in the calculation of ship waves, I have
carried out similar calculations to the authors, and
found similar results that is, that the near field
waves are essentially unaffected when the physically
correct value for the molecular viscosity is employed,
both in deep water and in finite depth.
However, following Tuck et. al. (1), and
Doctors (2), it was found that the use of values of the
viscosity around three orders of magnitude higher
had a significant impact in some cases. The divergent
wave system, and the transverse waves in the
relatively near field (within around 1km of a vessel of
around 100m waterline length), were still essentially
unaffected, but in the extreme far field (of the order
of 5km downstream of the vessel) the inclusion of the
higher values of viscosity led to the removal of short
wavelength ripples observed on longitudinal cuts
calculated away from the centreline of the vessel).
These ripples in some cases could add
substantially to an estimation of the height of these
waves albeit in a region in which the waves are
quite small compared to the leading divergent waves.
I would thus like to ask if the authors have found any
similar effects using their more sophisticated theory,
and ask for their comments on the appropriate values
of the viscosity and surface tension to optimise
correlation with physical measurements.
On the subject of wave decay, I was
interested to see the further confirmation of the
variation of decay rate with speed and vessel
geometry. I would be interested to know if the
authors have carried out any systematic comparisons
of catamarans and monohulls in this regard.
In some calculations I have carried out
recently I compared the waves generated by a
monohull and catamaran of the same total
displacement, and with similar underwater hull form.
I found that whilst the maximal height wave
generated by the catamaran along a given cut
(relatively close to the vessel track) was generally
smaller than that of the monohull, the rate of decay of
the wave height with distance from the vessel track
could be considerably less particularly in the high
sub-critical speed range. However the calculation of
the decay exponent in these cases was based on a

REFERENCES
1.

TUCK, E.O. AND LAZAUSKAS, L. ShipWave Patterns in the Spirit of Michell


IUTAM Symposium on Free-Surface Flows,
Birmingham, England, July 2000.

2.

DOCTORS, L.J. The Influence of Viscosity


on the Wavemaking of a Model Catamaran.
Proc. Eighteenth International Workshop on
Water Waves and Floating Bodies
(18 IWWWFB), Le Croisic, France, April
2003.

AUTHORS REPLY
The authors would like to thank Dr Day for
his contribution, by way of questions and comments,
to the discussion of their paper.
It is interesting to note Dr. Day's
confirmation that the use of the molecular viscosity
in the calculations of the wave elevation has a very
minor impact on the result. It is necessary to consider
very great distances downstream of the vessel before
any appreciable damping of the wave pattern can be
discerned.
The authors agree that one must choose a
turbulent viscosity many orders of magnitude greater
than the molecular viscosity in order to provide any
reasonable damping of the waves. Unfortunately, it is
difficult to physically justify such large values of the
turbulent viscosity. It is also unclear how it
influences the wave elevation along longitudinal cuts,
which are outside the viscous wake generated by the
vessel. In the case of the catamaran tests discussed in
the paper, the wave probes were stationary and
transversely located in order to avoid interference
with the passage of the model and its wake.
Therefore, we chose only to use the molecular
viscosity in our calculations.
In a similar manner, realistic values of a
surfactant film elasticity have a minor influence on
the wave system as long as relatively short distances
aft of the ship are considered. On the other hand,
variations of the surface tension itself provides a
slightly greater modification, at least at the model
scales studied here. Practically, its influence on the
final results is also insignificant.
Dr. Day also makes the point that ripples
(that would be damped by appropriate choices of the
turbulent viscosity or surface film elasticity) can
affect the estimation of the wave elevation at great
distances downstream. We cannot suggest what are
suitable values of the turbulent viscosity or of the

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

surface film elasticity since we have not made such


comparisons with experiments.
We agree with his remarks about catamarans
and monohulls. That is, catamarans may generate
(and often generate) a wave system near the vessel
with magnitude less than that of the equivalent
monohull possessing the same total displacement and
other appropriate geometric characteristics. However,
the rate of lateral decay of the wave system is often
greater for the monohull. Thus, the comparison of
these two types of vessel with respect to wave
generation, at a large distance off the track of the
vessel, is not entirely obvious.
Interested readers may wish to consult the
additional reference, here, which clarifies the
transverse rate of decay of the wave system for both
monohulls and catamarans.
REFERENCE
Doctors, L.J. and Day, A.H.: "The Generation and
Decay of Waves behind High-Speed Vessels", Proc.
Sixteenth International Workshop on Water Waves
and Floating Bodies (16 IWWWFB), Hiroshima,
Japan, pp 33-36 (April 2001).

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics


St. Johns, Newfoundland and Labrador, CANADA, 8-13 August 2004

An effective scaling device for model testing of air cushion


vehicles in a laboratory
Krish P Thiagarajan
(School of Oil & Gas Engineering, The University of Western Australia)

NOMENCLATURE

ABSTRACT
Air cushions are used to support a number of
floating platforms, e.g. surface effect ships, and
hovercrafts. These platforms suffer from what is termed
the cobblestone effect arising from a resonant
condition due to restoring force provided by the air
cushion. Model testing of air cushion vehicles have
suffered from a potential problem of non-scaling of
atmospheric pressure. Thus although the dynamic
pressure may be properly scaled, a model air cushion is
at a much higher pressure than the corresponding
prototype, resulting in a stiffer air cushion.
A
consequence of the increase in stiffness is the natural
frequency of the model is much higher, hence the
cobblestone effect may not be seen altogether in a
laboratory.
In order to scale the air cushion pressure one
of the following may be deployed:
1. Using a low pressure laboratory facility.
2. Using a reservoir to increase air volume.
3. Deployment of air stiffness correction devices
The first two methods need specialized laboratory or
equipment.
Earlier work from the Netherlands
employed a membrane diaphragm correction device
that allowed the pressure to be scaled properly.
Although some satisfactory results were obtained, there
were difficulties in implementing the device. In the
present paper, we propose a lenticular balloon (i.e. a
balloon with a large diameter ring inserted into it) as an
effective correction device.
The heave natural frequency is theoretically
formulated for the case of a rectangular air cushion
vehicle fitted with a correction device. The pressurevolume relationship of a lenticular balloon is developed
using linear elastic theory. A set of experiments was
conducted for lenticular balloons to evaluate their
volume pressure relationship which are then
compared to the theoretical formulation.

Ab
b
D
E
hb
KB
L
M
Ma
P0
Pa
R
Rr
t
t
V

cross sectional area of air cushion


structure width
structure draft
Youngs modulus of balloon material
initial height of air cushion
balloon correction factor for air stiffness
length of the structure
mass of the structure
heave added mass of the structure
air cushion excess pressure in static condition
atmospheric pressure
radius of curvature of diaphragm surface
Radius of prestressing ring inside balloon
time
balloon thickness
enclosed volume within the diaphragm and ring
surface
VB volume of balloon correction device
constant in adiabatic gas law
ratio of specific heats for air
azimuthal angle of ring when the diaphragm
stretches
model- prototype scaling ratio
ratio of incremental pressure change to excess
pressure
3 heave motion of the structure
a initial density of air in the cushion
instantaneous density of air in the cushion
T tangential stress on balloon
Poissons ratio for diaphragm material
n heave natural frequency of the structure
INTRODUCTION
Air cushions are used to support a number of
floating platforms, e.g. surface effect ships, hovercrafts
and recently, large volume concrete structures during

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

tow-out. In the latter case, the purpose is to provide


sufficient clearance from the seabed in case of shallow
water harbor conditions. For all the cases, vertical
motions of the platform in the presence of air cushions
is critical.
SES and hovercraft suffer from what is termed
the cobblestone effect i.e. accelerations arising from
a heave/ pitch resonant condition due to restoring force
provided by the air cushion.
A comprehensive
discussion of this effect may be found in earlier
publications, for example Kaplan, et al. (1981), Nakos,
et al. (1991), or Steen et al. (1993).
Model testing of air cushion vehicles have
suffered from a potential problem of non-scaling of
atmospheric pressure (e.g. Faltinsen 1991, p. 91-92)
While it is common to use Froude scaling to build a
geosim of a prototype, accurate dynamic simulation in
the laboratory is hampered because atmospheric
pressure is the same at both model and full scale. Thus
although the dynamic pressure may be properly scaled,
a model air cushion is at a much higher pressure than
the corresponding prototype, resulting in a stiffer air
cushion. A consequence of the increase in stiffness is
the natural frequency of the model is much higher,
hence the cobblestone effect may not be seen altogether
in a laboratory.
Commonly known means to properly scale the
air cushion pressure (Chakrabarti, 1994; Sec. 9.9) are:
1. reduction in ambient pressure, such as in a low
pressure laboratory facility.
2. increase in cushion volume by connecting to a
reservoir.
Both these methods suffer from practical difficulties
and require large infrastructure. A simpler method is
adopt a correction device that will allow the pressure to
be scaled properly. One such device is a membrane
diaphragm originally proposed by Kapsenberg and
Blume (1995) and researched by Moulijn (2000). The
latter has presented an elaborate set of model tests
using such devices. Although some satisfactory results
were obtained, there were quite a few difficulties in
implementing the diaphragm correctly. Moulijn (2000)
has also presented a numerical method of evaluating the
performance of the diaphragm.
In an earlier study on air lifted concrete
gravity structures conducted at the University of
Western Australia, a novel method of scaling the air
cushion using a pre-stressed balloon idea was
proposed1. This so-called lenticular balloon proved
1

The original credit for the lenticular balloon


idea goes to two industry participants in the research,
Mr. Chris Candy and Mr. Alan Hill, both from
Maunsell Australia at the time of the study.

quite effective in softening the air spring during the


experiments. Some discussion on this idea may be
found in Hill and Sow, (1997).
The purpose of this paper is to evaluate the
performance of the lenticular balloon device, both
theoretically and experimentally. We start with by
formulating the free heave problem, and explaining
how a correction device can enable proper scaling.
Then we proceed to apply linear elastic theory to the
case of an elastic membrane and then extending the
argument for a lenticular balloon, arriving at a
relationship between volume and pressure. A set of
experiments are conducted for different diameter
balloons to evaluate their volume pressure
relationship which are then compared to the theoretical
formulation.
THEORETICAL BACKGROUND
We follow the notation and analysis approach
suggested by Faltinsen (1990, p. 91-92). We consider
the free heave motion of an open bottom rectangular
structure of length L, width b, and draft D supported by
an air cushion of area Ab (Figure 1). We ignore the
buoyancy effect due to the side walls, and effectively
treat the structure as having area Ab as well. The air
cushion has an excess pressure P0 over atmospheric
pressure. The mass of the structure is supported by the
air cushion, such that

M=

Ab P0
g

(1)

Our aim is to find the heave natural period of the


structure, due to the presence of the air cushion.
The air volume follows the adiabatic
relationship:

( P0 + Pa ) = a ,

(2)

where a is the density of air, is the specific heat


ratio of air, and is a constant. For small heave
motions, the density of air changes according to
changes in pressure

= a 1 +

1
1 + Pa

(3)

P0

where the instantaneous pressure in the cushion is

P = P0 + Pa + p and = p

P0

. The volume of

air within the cushion without wave effects is given by

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

As an extension to the discussion, we can


consider addition of a correction device, whose volume
will be designed in such a way that the natural
frequency scales correctly. If the volume of this
correction device is sinusoidal, then

Ab

hb

VB

= KB
,
t
t

(4)

If no leakage of air occurs, then the mass of air in the


cushion is a constant, which leads to the following
linearized continuity equation (see Faltinsen, 1990)

hb

1+

Pa
P0

3
+
= 0.
t
t

(5)

23
= Ab P0 .
t 2

(6)

If a sinusoidal heave motion is assumed

3 (t ) = {30 e

in t

},

W (t ) = b ( hb + 3 (t ) ) dx +VB (t ) . (12)
L

The equation of continuity for the air mass then


becomes

P
1+ a
P0

The heave equation of motion (Eq. 6) can be solved


along with continuity (Eq. 5) to obtain the linearized
heave natural frequency

n =

P
g
1+ a ,
Cm hb
P0

(9)

where Cm = 1 + Ca, is the heave inertia coefficient of


the structure.
This equation scales improperly because of the
presence of the atmospheric pressure in the equation.
Specifically, if the model full scale ratio is , then the
above equation gives

n ,model
,
n,prototype

instead of

Ab g

n2 =

(7)

(8)

n 0 + Abn30 + VB 0n = 0 ,

(13)

which can be solved along with the heave equation of


motion to obtain the equation of heave natural
frequency as

then this results in a sinusoidal variation of pressure,

(t ) = {0 ei t } .

(11)

where KB is a correction factor, closely related to the


reciprocal of the spring constant of the correction
device. The volume of air at any instant of time is now

The heave equation of motion for the structure


is given by:

(M + M a )

(10)

In the simplest possible case, we can imagine a linear


change in volume for a corresponding change in
pressure, i.e.

Figure 1. Definition diagram

W = b ( hb + 3 (t ) ) dx .

VB = VB 0 eint .

Cm

W
1+ Pa

(14)

+ KB
P0

If VB(t=0) is zero, then as a linear approximation


W=Abhb, and the above equation then simplifies to the
Eq. (9) for KB = 0. By enforcing a relationship
between the natural frequencies of the model and the
prototype, we obtain the following relationship for the
correction factor

KB =

Ab hb

P0 Pa ( 1)
. (15)
( P0 + Pa )( P0 + Pa )

All the parameters in the above equation are at model


scale.
LENTICULAR BALLOON DEVICE
A lenticular balloon is a regular balloon with a
circular ring inserted in it, such that the inflated balloon

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

has a shape of a spherical lens. As common experience


would indicate, the balloon is pre-stretched and hence
the expansion is much more uniform with increase in
pressure. The lenticular balloon device has similarities
to the diaphragm idea proposed by Moulijn (2000). In
fact the behavior of the balloon is similar to two
diaphragms located back-to-back.
We develop the theory of the lenticular
balloon based on linear elastic theory. We analyze onehalf of the balloon and derive a relationship between
the volume and pressure. It is interesting to note that
Moulijn obtained the corresponding result numerically.
Here we propose a closed-form solution.
Consider a ring of radius Rr supporting a
circular diaphragm (Figure 2). As the diaphragm
expands in volume, the pressure applied changes the
tension in the diaphragm, which in turn is balanced by
the reaction force on the ring. A simple force balance
gives the following equation relating pressure with the
tangential stress at the membrane.
(16)
P0 R sin = 2t T
where t is the membrane thickness. The above equation
may be perturbed to identify a relationship between the
rate of change of pressure and corresponding rates of
change of stress and thickness.

p =

2t
2
d T + T dt
Rr
Rr

(17)

Using linear elastic theory, the rate of stress may be


related to the rate of strain. Thus

d T =

E
d
1

(18)

The normal rate of strain is the change in thickness,


which is given by

dt
=
2d T
t
E

(19)

The rate of planar strain is directly proportional to the


rate of change of area of the membrane, thus

d =

1 dA
2 A

(20)

The membrane area can be directly related to the radius


of curvature of the membrane,

d =

dR (1 cos )
R 2 cos

(21)

The volume enclosed by the diaphragm may be found


in a straightforward manner to be

2
1
V = R 3 (1 cos ) RR 2 R cos
3
3

whereupon, the rate of change of volume may be linked


directly to the rate of change of curvature.

dV = 2 R 2 (1 cos )

(22)

RR2
dR .
cos

(23)

It is noted that for a lenticular balloon the rate of


change of volume is twice that of the above equation.
The above gives a complete set of equations
that can be solved to get the relationship between the
rate of change of volume and pressure.
Upon
simplification, we find the correction factor due the
diaphragm

Rr
R

KB =

dV
P0
dP

= 2 R 2 cos

RR2

(1 cos )

P0 Rr R (1 )

( 2

T0

E t

(24)
As before the value for the lenticular balloon is twice
the above value.

y
x
Figure 2. Schematic of circular diaphragm

Moulijn (2000) performed a number of


experiments with a rubber diaphragm. The diaphragm
was rectangular with dimensions 0.745 m x 0.845 m,
had a thickness of 0.35 mm. The Youngs modulus
used was 4.64 x 106 Pa, and a Poissons ratio value of
0.49. The experimental values were found to be in good
agreement with numerical results for circular

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

diaphragm of diameter 0.775 m. Figure 3 shows the


experimental results of Moulijn (2000) and the
theoretical result from Eq. (24) for ring diameter of
0.775 m. As can be seen the agreement is reasonably
good.

8.E-05

Theory
Moulijn (2000)

4.E-05
2.E-05

Figure 4. Comparison of circumferential arc with actual


inflation of a lenticular balloon.
0

200

P0

400

600

Figure 3. Comparison of theory with experimental


results of Moulijn (2000).
VALIDATION WITH EXPERIMENTS
Common party balloons that are expandable to
2 ft in diameter were used. Different balloons were
trialed and the properties were found to vary slightly.
The balloons were pre-stressed by inserting steel rings
into the balloons, stretching the balloons to form a flat
circular shape. The diameter of the ring used was
0.23m. An air pump was used to inflate the balloon
and a pressure transducer was used to measure the
pressure of the air inside the balloon. The amplified
pressure transducer data was logged using data
acquisition software.
The circumference of the balloon was
measured in different directions using a tape measure
and averaged. The inflation of the balloon was not
uniform, and varied due to spatial variation in
thickness, and due to small changes in the geometry of
the metallic ring. Figure 4 shows a photograph of an
inflated balloon and an averaged circumferential arc
superposed on it. The radius of this arc was used in
volumetric calculations.
Initial calculations using the Youngs modulus
as stated by Moulijn (2000) gave inconsistent results.
Hence a proper measurement of the Youngs modulus
was initiated. Rectangular strips of the balloon material
were stretched using an Instron 8501 Mechanical
Tensile Testing Machine at a constant rate and then
slowly released. This procedure was repeated twice,
once at a rate of 10mm/minute and once at a rate of
100mm/minute. Using these results, an estimate of the

engineering Youngs Modulus was calculated. Figure 5


shows the volume- pressure relationship for the
following parameters.
Ring radius, Rr = 0.115 m
Thickness, t = 0.28 mm
E = 8 x 105 Pa
= 0.49
The theoretical predictions show reasonable agreement
with the experimental measurements.

0.E+00

Volume (m )

dV/dP

6.E-05

5.E-03
5.E-03
4.E-03
4.E-03
3.E-03
3.E-03
2.E-03
2.E-03
1.E-03
5.E-04
0.E+00

Theory
Exp

200

400

600

800

1000

Pressure (Pa)

Figure 5. Volume vs. pressure for a lenticular balloon.

EXAMPLE CALCULATION
We consider a rectangular box supported by
an air cushion. Particulars are given in Table 1 below.
For this calculation we are interested in a scale factor of
100. The scaling correction factor (KB) is shown in
Figure 6. It is seen that five balloons are required to
achieve the correct scaling for the model.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

CONCLUSIONS
The paper has presented and discussed the
theoretical formulation of a circular diaphragm type
device for scaling air cushions in a laboratory. A
development of the diaphragm idea is a lenticular
balloon described in this paper. The balloon is easier
to deploy than the diaphragm, as it requires lesser deck
space on the model. Further, the number of balloons
and the prestressed diameter may be easily modified to
obtain the required scale ratio.
It was noted during our experiments that the
performance of the balloon is very sensitive to its
thickness and Youngs modulus. Thus time and effort
needs to be allocated to accurate measurement of these
parameters.
ACKNOWLEDGEMENTS
The balloon calibration tests were performed
by Kellie Bacon as part of her vacation employment at
the University of Western Australia. Sean Galvin at the
School of Mechanical Engineering assisted with the

0.010
0.008
0.006
KB

Table 1. Parameters for the example scaling calculation


scale ratio
100
Quantity
Prototype
Model
Length (m)
50
0.5
Width (m)
50
0.5
Draft (m)
10
0.1
Skirt length (m)
10
0.1
2
Area (m )
2500
0.25
Heave Cm
1.8
1.8
P0 (Pa)
100553
981
hb (m)
10
0.1
n (rad/s)
1.264
Tn (s)
4.971
n (uncorrected) rad/s
92.908
Tn (uncorrected) s
0.068
4
KB (required) m /Pa
0.008
n (corrected)
12.722
Tn (corrected)
0.494
Lenticular balloon
E
800000
t
0.00028
Rr
0.112
0.490

No. of balloons
5
4
KB (per balloon) m /Pa
0.002
KB (total) m4/Pa
0.008

0.004
0.002
0.000
0

500

1000
1500
P0 (Pa)

2000

Figure 6. Scaling correction factor (KB) vs. excess


pressure P0.
setup and conduct of the Youngs modulus tests.
REFERENCES
Chakrabarti, S. K., Offshore Structure Modeling,
World Scientific, Singapore, 1994.
Faltinsen, O. M., 1990. Sea Loads on Ships and
Offshore Structures. Cambridge University Press.
Hill, A. B. and Sow, E., Model research programme
for hydrodynamic response of an air lifted CGS in
limited water depth. Report 185, Minerals and Energy
Research Institute of Western Australia, 1997.
Kaplan, P., Bentsen, J., and Davis, S., Dynamics and
hydrodynamics of Surface Effect Ships, SNAME
Transactions, Vol. 89, 1981, pp. 211 247.
Kapsenberg, G K and Blume P., Model tests for a
large surface effect ship at different scale ratios,
Proceedings FAST 95, 1995.
Moulijn, J. Added Resistance due to Waves of Surface
Effect Ships, Ph D Thesis, Delft University of
Technology, The Netherlands, 2000.
Nakos, D E., Nestegard, A., Ulstein, T., and
Sclavounos, P D., Seakeeping analysis of Surface
Effect Ships, Proceedings FAST 91, 1991, pp. 413
428.
Steen, S, Ulstein, T., Faltinsen, O. M., Seakeeping and
comfort of large SES, Proceedings FAST 93, 1993, pp
1077 1091.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Larry J. Doctors
The University of New South Wales, Australia
Thank you for a very interesting paper.
1.

The matter of proper scaling for model


surface effect ships is one that has been
studied before and so it is good to see a fresh
approach to the problem. Can the author
clarify the way the added mass was
estimated? Professor T.F. Ogilvie wrote a
report (University of Michigan, Department
of Naval Architecture and Marine
Engineering, Report 30, 1969) in which the
two-dimensional added mass for this
geometry was analyzed. Perhaps the results
can be adapted to the present threedimensional case.

2.

Would it be feasible to employ such


balloons at full scale to improve (that is,
reduce) the response of the ocean platform?

AUTHORS REPLY
Thanks for the comments.
1. At present, the calculation is focused on the
scaling issue, and hence the added mass was
considered a constant. It was assumed to be
similar to a closed bottom box, i.e. about
80% of displacement. We could consider
using the added mass estimation from
Professor
Ogilvies
report
as
an
improvement.
2. It would be quite difficult to control the
balloons in full scale I think.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics


St. Johns, Newfoundland and Labrador, CANADA, 8-13 August 2004

Large Eddy Simulation of the Viscous Flow


around a Ship Hull
Including the Free-Surface

E. Lillberg and U. Svennberg


(Weapons & Protection, The Swedish Defense Research Agency,
FOI, Stockholm, Sweden)

Abstract
In this study we use the Navier-Stokes Equations
(NSE) for viscous incompressible fluid flow. The NSE
are solved using Large Eddy Simulation (LES) methodology combined with an interface capturing algorithm in a finite volume framework. This enables simultaneous computations of the air-water flow around
a surface ship including the evolution of the water surface, i.e. the Kelvin wave pattern. The ability to compute the entire flow field gives the possibility to assess
design parameters. It, also, and maybe more important
in the case of surface combatants, enables the study of
the turbulent wake behind the ship. Beside this it can
be used for design and optimization issues. In this
study we use the DTMB 5415 model hull representing
a modern surface combatant with transom stern and a
sonar dome configuration below the bow. This design
has become an important case for validation of numerical ship hydrodynamics simulation methods. The
results presented in the present study are from an ongoing research project. They are in reasonable good
agreement with available experimental data.

Introduction
Traditional naval hydrodynamic design procedures
rely on regression analysis and towing tank testing using data bases. These data bases are developed from
decades of model test results correlated with full scale
ship experiences. For new ship types, the existing da-

tabases do not always provide the necessary design information. The lack of information becomes even
more critical due to increased demands for stealthy operation for future surface ships. During the past decades Computational Fluid Dynamics (CFD) has developed into an efficient and powerful tool for design and
validation. It also is an important tool for investigating
the physics of complex flow phenomena. Over the past
twenty-five years there have been several international
workshops on ship flow calculations (Larsson 1980,
Larsson et al. 1990, Kodama 1994, Larsson et al.
2000). A development from potential flow methods to
simple and more advanced Reynolds Averaged Navier-Stokes (RANS) models can be seen on these
workshops. An interest in addressing more complex
flow fields, such as including the free surface, propulsion, manoeuvring and sea-keeping can be seen at conferences like the International conferences on Numerical Ship Hydrodynamics, where the 8th was held in
South Korea last year and the Symposiums on Naval
Hydrodynamics, where this conference is the 25th.
Still, there is much to be done before all important
physical phenomena are accurately calculated or modelled. One step in this direction is to turn to LES models.
The ability to accurately predict the wave pattern of ships and submarines are of importance for
drag assessments, signatures and secondary effects
such as costal erosion. Furthermore, the behaviour of
the free water surface affects the flow field around surface ships and submerged vessels near the surface.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Their motion generates a disturbance on the water surface which is important for proper understanding of
the flow around these vessels. The free surface affects
the location and magnitude of vortices produced by the
motion of the vessel. Vortices that originate from the
bow, appendages, propulsors and rudders can greatly
affect the performance of the propulsors.
In the present study we use the NSE for viscous incompressible fluid flow. The NSE are solved
using LES methodology combined with an interface
capturing algorithm in a finite volume framework.
This enables simultaneous computation of the air and
water flow around a surface ship. This includes the
evolution of the water surface resulting from the pressure disturbance due to the ship motion, the Kelvin
waves. The ability to compute such flow fields using
LES not only gives the possibility to provide assessments of design parameters. It also, and maybe more
important in the case of surface combatants, enables
the study of the turbulent wake behind the vessel. A
ship that moves through the water does not only create
the detectable surface wave pattern but also a persistent trace of turbulence, vortices and waves underneath
the surface. This viscous wake, and its interaction with
the Kelvin wave pattern, creates a distinct trace for up
to 20 kilometres astern of the ship, which creates possible means of detection or for wake homing torpedoes.
In this study we use the 5415-model hull,
(see: Model 51415 web page in reference list), fabricated at the David Taylor Model Basin (DTMB), Naval Surface Warfare Center, Carderock Division, as an
early design variant of the DDG51 Arleigh Burkeclass destroyer currently deployed in the U.S. Naval
fleet. For this hull several experimental investigations
have been made, (Model 5415, Larsson et al. 2000),
which makes it suitable for validation of CFD models
and verification of numerical results. However, there
are some differences in the experimental data sets from
different facilities. The 5415 model represent a modern
surface combatant with transom stern and a sonar
dome configuration below the bow. The experiments,
(Model 5415), cover several Froude numbers (Fn) but
only Fn=0.28 is considered here to avoid breaking
waves and their inherited difficulties. This problem
will be addressed in future studies.
A long-term goal of the research into numerical simulations is to develop the ability to study the
performance of a full ship design. This including the
interactions of the flow fields around various appendages, sonar domes, rudders, shafts and propulsors. The
resulting complex flow field is of interest as the ship
maneuvers in response to changes in the settings of the
rudders and propulsors in various sea-states. To accomplish this goal, efficient and accurate free surface
simulations are needed. Other necessary components

to achieve this goal include capabilities to monitor the


onset of incipient cavitation, to perform simulation of
several vessels in motion relative to one another and to
adapt the grid to capture vortices and other important
flow phenomenon in the water. Other complications
are how to give physically correct boundary conditions
on the very large computational domain and how to
get sufficient computational power to do the calculations accurately.

Computational Methods
As the available computational power increases significantly every year, the frontiers of CFD is constantly expanding towards more complex problems
and more advanced models and algorithms. Thus, taking more of the physics into account and also allowing
for cross-disciplinary studies such as fluid/structure interactions,
hydro-acoustics
or
magnetohydrodynamics. Even so, the possibility to do Direct
Numerical Simulations (DNS) in which all scales of
the flow field are resolved is restricted to simple flows
at low Reynolds number (Re), emphasizing the need
for modelling at higher Re. Potential flow methods do
an excellent job if only the wave pattern on the water
surface and the wave resistance are of interest. However, the viscous effects are neglected and more advanced models are needed when the viscous effects or
the viscous flow field are of interest. Today the RANS
equations are the most commonly used flow model for
practical applications. Here, the NSE are filtered in
time to produce a model for the time-averaged (or
mean) flow. RANS is often quite adequate for mean
flow predictions, but provide only limited information
about turbulence characteristics and almost no details
on the large-scale unsteady structures of the flow field.
RANS also gives reasonably good results with short
turnover time for flows where the large eddies are almost stationary and the modelled fluctuations are isotropic. However, for more complex flows this may not
be true, and an improved method may be required. For
these types of flows a more appropriate method is
LES.
In LES, (Ferziger and Leslie 1979, Lesieur
and Metais 1996, Boris et al. 1992, Sagaut 2001), all
scales larger than the grid cells are resolved using a
space/time accurate algorithm and only the effects of
the small scales on the large scales are modelled. The
direct computation of the large energy containing eddies (being flow and geometry dependent) gives LES
more generality than RANS, utilizing models for the
entire spectrum of turbulent motions. At present, there
is hardly any study in the open literature where LES is
applied to flow around ship-hulls including the nearwake. The main reason obviously lies on the computer
resource limitation which increases with the Re-

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

number. In particular, for the near-wall region, it is estimated that the number of grid points required for a
wall-resolved LES scales as O(Re2), where Re is
the friction velocity based Re-number, (Bagget et al.
1997). Moreover, unless the grid is sufficiently fine,
the anisotropy of the flow will cause anisotropy of the
subgrid flow, necessitating subgrid models capable of
handling simultaneous flow and grid anisotropy. Consequently, high Re-number flows, especially in complex geometries, are too expensive to compute with
LES unless particular techniques are invoked to alleviate the severe resolution requirements near the wall.
Here we will follow (Fureby et al. 20003, Wikstrm et
al. 2003), and use a separate wall model, based on the
solution to the classical boundary layer equations.
The NSE (1) which are the equations governing fluid flow, (Lions 1996), consist of conservation
and balance laws for mass and momentum supplemented by constitutive equations. For a viscous, incompressible, Newtonian fluid, with density variations
they can be expressed as,
t (v )+(v v )=p+S +g,

v =0,

(1)

where v is the velocity, the density, p the pressure,


S =2D the viscous stress tensor, D= 1 2 (v +v T ) the
rate-of-strain tensor and the viscosity, and g the
gravitational acceleration. Numerical solution is the
only practical way of solving (1) due to their nonlinearity and extensive range of eddy scales. The ratio
of the largest eddies (with a size of I) to the smallest
Kolmogorov eddies (with a size of K) is related to the
Re-number as I / K =Re 3 / 4 , which implies that the
number of degrees of freedom required in DNS scale
as Re9/4. For very high Re-numbers, present-day supercomputers are not powerful enough to handle such
problems and hence alternative methods, such as potential flow models, RANS and LES, have to be devised for complex turbulent flow simulations.
For the flow around a surface ship the
mathematical model is even more complicated since
we also have to account for the motion of the free surface separating air and water. To model such interfaces
several approaches exists which broadly can be divided into either interface capturing, such as Volume of
Fluid (VOF) or Virtual Boundary Methods, or interface tracking methods, such as Arbitrary LagrangianEulerian, Immersed Boundary or Boundary Integral
methods. In the present study the well-known VOF
method is applied through the Compressible Interface
Capturing Algorithm for Arbitrary Meshes (CICSAM)
scheme, (Ubbink 1997). In the VOF approach a pas-

sive scalar, , representing the volume fraction of one


phase with respect to the other is convected with the
flow by a transport equation. All fluid properties as
well as the modelled eddy viscosity are then averaged
values in . Together with the NSE the system of
equations to be solved then becomes,
t (v )+(v v)=p+S +f +g,

v =0,
t ( )+(v )=0,

(2)

where =1 +(1) 2 is the density, S =2D the


viscous stress tensor with =1 +(1 ) 2 being the
viscosity and f the surface tension. Here, indices 1 and
2 refer to the water and the air phase, respectively.
Surface tension is a tensile force tangential to the interface separating the fluids, with a magnitude depending
mainly on the nature of the fluids. In ship flow applications, however, the surface tension can usually be
neglected. Furthermore, the definition of the indicator
function implies that is a step function and hence
is piecewise continuous. In order to model the two
fluids as one continuum, using (23), should therefore
be continuous and differentiable over the entire domain, cf. (Sussman et al. 1994). To accomplish this we
give the transitional region between the two fluids a
small but finite thickness, say , so that =1 in fluid 1,
=0 in fluid 2 and 0<<1 within the transitional region.

The Large Eddy Simulation (LES) Model


In LES the motion is separated into small and large
eddies and equations are solved for the latter. This separation is achieved by means of a low-pass filter, cf.
(Sagaut 2001). The filter is here applied by convolving
the dependent variables with the kernel G=G(x,),
where is the filter width, to extract the large scale
components. The filter width, , is usually connected
to the grid cell size. More precisely, we write v = v + v
and p= p + p , where ( ) denotes the resolved parts. Filtering the NSE yields
t (v )+(v v )= p +( S B)+f ,

v =0,
( )+( v)=b,
t

(3)

where B = ( v v v v ) the subgrid scale stress tensor


and b the subgrid flux vector which is neglected in the
method used here. B represents the effects of the subgrid flow on the resolved flow, and must be modelled
using information from the resolved flow prior to discretization at a resolution near , much more afford-

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

able than DNS. Presently, two modelling strategies exist: Functional modelling consists in modelling the action of the subgrid scales on the resolved scales. This
is basically of energetic nature so that the balance of
the energy transfers between the two scale ranges is
sufficient to describe the subgrid scale effects, cf.
(Fureby 2001b). Structural modelling consists of modelling B without incorporating any knowledge of the
nature of the interactions between the subgrid and the
resolved scales. Such models can be based either on
series expansion techniques, transport equations, scale
similarity or other deterministic approaches, e.g.
(Fureby 2001b). For the purpose of this investigation
we use functional modelling in which the energy transfer mechanism from the resolved to the subgrid scales
is assumed analogous to that of a Brownian motion
superimposed on the motion of the resolved scales.
Accordingly,
B D =B 13 ( trB)I =2 k D

(4)

where BD is the deviatoric part of B and k is the subgrid eddy viscosity. To close (4) we need models for
k and for this we assume the existence of characteristic length and velocity scales and we infer separation
between resolved and subgrid scales, viz.,
k =c /3 k ( 1)/ 2 1+/3 ,

(5)

where k is the turbulrnt kinetic energy,


c =A/( K 0 4/3 )( 32 K 0 ) ( 1)/ 2 (1 )/3 ,
with
K0=1.50,
A=0.44 and is a parameter to be varied. Different
forms of k can be obtained for different -values. For
which
together
=1
we
get k =c11/3 4/3 ,
2
3
with = ||D|| , where is a dissipative length scale,
results in k =c14/3 2/3 ||D|| . With =c the Smagorinsky
(SMG)
model,
(Fureby
2001b),
k =c D 2 ||D|| , is recovered with cD=0.02 and c=1.05.
For =0 we get k =c k k 1/ 2 , with ck=0.05. Different
ways can be used to determine k, but we prefer the
One-Equation Eddy Viscosity Model (OEEVM) in
which,

t (k )+(kv)=BD +(( + k )k ),
=k 3/ 2 /,

(6)

is solved in conjunction with (2). When =c the


standard OEEVM, (Fureby 2001b), is recovered.

Numerical Methods
LES require high-order numerical methods to avoid
masking B by the leading order truncation error. In

general, is related to the grid, i.e. |d|, where |d| is


the grid size, which makes the subgrid stress tensor an
O(|d|2) term. In LES, spectral and high-order finite
volume, element or difference methods, for spatial discretization, are usually combined with explicit or semiimplicit time-integration schemes. For complex geometries the Finite Volume (FV) method is the best
choice. Here the computational domain D is partitioned into non-overlapping cells P. The cell-average of f
over the Pth cell is f P = 1V f dV so that Gauss theorem

can be used to derive the semi-discretized LES-equations. By integrating these over time, using e.g. a
multi-step method, (Hirsch 1999), the discretized
equations become,

i t
[F C, ] n + i =0,
VP m f f
t

( i ( v ) nP + i + Vi P f [FfC, v +FfD, v +FfB, v ] n + i )=


i=0
(7)

= i (gradp ) nP + i t + i (f ) nP + i t ,
m
n + i i t
C, n + i
i = 0( i () P + VP f [Ff ] )=0,

where m, i and i are parameters of the timeFfC, =( vdA ) f ,


integration
method,
and
C, v
D, v
Ff =( vdA) f v f , Ff =(v) f dA , FfB, v =(B) f dA
and FfC, =( vdA) f f are fluxes, where the superscripts C, D, B denotes the convective, viscous and
subgrid fluxes, respectively. To close the FV-discretization the fluxes (at face f) need to be reconstructed from the variables at adjacent cells. This requires flux interpolation for the convective fluxes and
conventional difference approximations for the inner
derivatives of the viscous and subgrid fluxes. Linear
interpolation is used, resulting in a second order accurate discretization, and (7) are decoupled by combining
(71) and (72) into a Poisson equation for p , which is to
be solved together with (72). The scalar equations are
usually solved sequentially with iteration over the explicit coupling terms to obtain convergence. The segregated approach yields a Courant (Co) number restriction with Co<0.5, but a value of Co=0.2 is preferable since LES aims at resolving the dynamics of the
smallest resolved flow structures.
To capture the dynamics of the interface separating water and air a specific reconstruction algorithm
for the term FfC, =( vdA) f f in the -equation is
used. In mathematical terms this is motivated by the
elliptic-hyperbolic nature of the -equation responsible of wave-propagation effects. The CICSAM
scheme, (Ubbink 1997), is based on the donor-acceptor flux formulation that forms the basis for compressive differencing, whereby downwind information is
included through the use of non-linear flux functions,

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

e.g. (Lafaurie et al. 1994) CICSAM determines the


flux FfC, by merging the fluxes from two highresolution schemes to comply with the Convective
Boundedness Criterion (CBC), the Hyper-C (HC)
scheme, (Gaskel and Lau 1988), and the UltimateQuickest (UQ) scheme, (Leonard 1991). This blending
is accomplished by a non-linear flux-limiter which
depends on the local orientation of the interface and
the direction of flow. Given flux-functions fHC and
fUQ from the HC and UQ schemes, respectively,
FfC, =FfC, [ fHC +(1) fUQ ] . To enforce monotonicity and positivity in multiple dimensions operator splitting is used, in which each direction is handled separately, and with the final flux obtained by combining
individual pairs of sweeps from different directions.

Wall Functions
LES of wall bounded flows becomes prohibitively expensive at high Re if one attempts to resolve the small
but dynamically important eddies in the near-wall region. These structures can be captured in a wallresolved LES in which the grid is approximately such
that y+<2, x+<200 and z+<30. Here, + denotes
non-dimensionalization by the viscous length scale
/u and the friction velocity, u =1w/ 2 , where
w =(v / y)| w is the wall-shear stress. The number of
grid points (degrees of freedom) N necessary, scale as
NRe2, (Bagget et al. 1997), which is nearly the same
as for DNS. The classical remedy to this problem is to
modify the subgrid model, by means of damping functions, that provides the correct scaling of k with y+ as
the wall is approached. Following (Fureby et al. 2003,
Wikstrm et al. 2003), we here instead use an explicit
wall model together with the OEEVM. Following
(Wikstrm et al. 2003) we determine (locally) the friction velocity u from the law-of-the-wall,
y + if y + 11.225,
v + = 1
+
+
ln( y )+ B if y >11.225.

The selected testcase is the DTMB 5415 model hull


(Model 5415). It is an early design variant of the
DDG51 Arleigh Burke-class destroyer currently deployed in the U.S. Naval fleet. It represent a modern
surface combatant with transom stern and a sonar
dome configuration below the bow. For this hull several experimental investigations have been made,
(Model 5415, Larsson et al. 2000), which makes it
suitable for validation of CFD models and verification
of numerical results. It was used as one of the test
cases at the Gothenburg 2000 workshop (Larsson et al
2000). Since then it has become a standard test case in
the ship CFD community. Six RANS computations for
this case were presented at the Gothenburg 2000 workshop.
The experiments, (Model 5415), cover several
Froude numbers but only Fn=0.28 is considered. In the
present study the DTMB 5415 hull, figure 1, with
length L=5.72 m is placed in a 26.0 m long, 12.0 m
wide and 3.0 m deep canal with 1.0 m of air above the
water surface. Two domains are used; one half hull
domain with a symmetry plane along the centreline resulting in a 6.0 m wide computational domain and one
with the complete hull. In both domains the hull is
placed with the Forward Perpendicular (FP) at x=3.0
meters from the inflow plane and trim and sinkage is
applied according to measurements at DTMB, (Model
5415), for a model velocity of 4.01 knots. For these
conditions the numerically computed wetted surface,
SDWL deviates only 2% from the experimental value.
Physical values for densities and viscosities are chosen
for both water and air and yields a Re=12.6106 based
on the length of the hull, the free stream velocity and
the viscosity and density (i.e. 1 and 1, respectively)
of water.

(8)

The approximate wall boundary condition can be implemented by adding a subgrid wall-viscosity BC to
the molecular viscosity on the wall so that the subgrid viscosity becomes,
k = + BC = w /(v / y) P =u y P / v P+ ,

Geometry and Conditions

(9)

where the superscript P denotes that the quantity is to


be computationally evaluated at the first grid point
away from the wall.

Figure 1: Side view of the DTMB 5415 model hull


geometry.
The boundary conditions are those of moving
walls for all outer boundaries except the inflow and
outflow planes. The velocity and the undisturbed free
surface location are specified on the inflow plane, and
Neumann conditions are used for all quantities except
the pressure, which is held constant on the outflow
plane. On the hull, no-slip conditions are employed,
whereas symmetry conditions are used on the centre
plane (y=0) for the half-hull computations. A uniform

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

velocity, a constant pressure and the domain divided


into air and water through the field serves as initial
conditions. The computational domain is discretized
using 1106 and 2106 hexahedral cells for the half-hull
and complete-hull domains respectively, figure 2. The
distance from the hull to the first cell centre is about
2.010-4 m, resulting in a y+ value between 3 and 30
depending on the local flow velocity which is comparatively high when referring to near-wall resolved
LES but suitable for the use of wall-functions.
The LES calculations are run from uniform
initial conditions approximately one half flow-through
time equal to three flow-past times to obtain a developed flow field near the hull. It also takes some

els (Lillberg and Fureby 2001) and Experience from


other hydrodynamic computations (Alin et al 2003,
Bensow et al 2004). The investigation will be continued with a finer grid, test of other subgrid models,
other numerical schemes for the VOF method and propulsion modelling using actuator discs.
The numerical results are, in general, in reasonably good agreement with the experiments. The
differences between the half-hull and complete-hull
configurations are small with only small asymmetric
fluctuations over the centreline just behind the sonar
dome configuration. This leads one to believe that further computations for steady motion straight forward
could be carried out on a half-hull domain to save time
and computer resources. For unsteady ship motion
there would be no alternative but to use the completehull configuration. In the remaining part of the paper
only results from the complete-hull calculations are
shown.

Figure 2: Perspective view of the hull surface and selected parts of the volume grid for the half hull configuration.
time for the free surface wave pattern to develop. During this process the half domain, using the centreline
symmetry, and a coarser mesh of 500.000 cells was
used together with the Smagorinsky model. After this
initialization period the OEEVM model with wall
function (OEEVM+WM) was used throughout the remaining simulations for a total of two flow-through
times equal to ten flow-past times. The results from
this initialization process where mapped to the finer
grid and mirrored to the complete domain.

Numerical Results
The results presented here is a part of an ongoing research project. The focus of the present study is not
only to compute the evolution of the free surface but
also to give an accurate description of the sub-surface
wake and flow field resulting from the ships motion.
These different features are important not only for design optimization with respect to performance but also
with respect to the creation of a detectable trace in the
ocean. The present calculations are based on the experience from earlier computations including the free
water surface on coarser grids with other subgrid mod-

Figure 3: Subsurface view of the LES flowpattern past


the DTMB 5415 hull at Fn=0.28 showing the surface
elevation, streamlines colored with vorticity magnitude
and sliceplanes showing normalized axial velocity,
v x /v.
The complexity of the subsurface wake
makes this model geometry interesting for signature
prediction analysis. The sonar dome induces a complex flow field containing several vortices. These vortices rollup to one vertical structure on each side of the
keel and they survive for a long time. They interact
with and redistribute the flow velocities in the boundary layer along the hull. In figure 3 a perspective view
from below shows the surface elevation, streamlines
coloured with vorticity magnitude and planes showing
contours of normalized axial velocity, v x /v. In this
picture the rollup of the sonar dome vortices are

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

clearly seen. Their propagation along the hull through


the plotted cross flow planes is also clearly visible.
The strong vortical structures from the sonar
dome puts a characteristic fingerprint in the wake far
behind the ship stern that may be used for detection.
This can be accomplished by, e.g. detection of the entrained bubbles in the wake structures, the vortical
structures themselves or the augmented turbulent intensity they create in their vicinity, the electrical currents produced in salt water or from internal waves
produced by the motion in the wake region relative the
surrounding sea water. For a clearer understanding of
the predominant flow features in the wake region as
well as the evolution of the wake a perspective view of
the stern is shown in figure 4. In this figure the evolution of the flow field is shown up to three hull lengths
downstream of the stern using streamlines released
randomly on the subsurface part of the hull. These
streamlines are colored with turbulent kinetic energy.
Two crossflow planes with secondary velocity vectors
are shown to visualize the evolution and broadening of
the wake. Close to the hull, emanating from the sonar
dome, are again the characterizing pair of vortices but
here are also vortical structures closer to the hull, best
represented on the starboard side, seen as they are
formed underneath the hull and intensified at the rising
of the stern, these are also visible close to the hull in
figure 10. Also, along the ship sides close to the free
surface strong vortices are created and swept towards
the centreline at the transom stern. The recirculation
zone behind the transom stern generates turbulence
and vortical structures, which interacts with the free
surface in the wake. All these entangled structures sum
up to create the complexity of the wake and results in
long lasting structures detectable far behind the ship.

One important issue in the study of the flow


field around a surface ship is the water surface. It is a
sharp interface between air and water, were we have a
sudden change in both density and viscosity. The location of this interface is not known before hand it is not
stationary. Therefore, it is a challenging task to track
and resolve the water surface accurately. The predictions herein for the wave patterns on the hull and for a
wave cut close to the hull are good; particularly taking
into account the size of the grid. For the far field wave
pattern the grid is much too coarse to give any accurate
predictions. The grid coarsens rapidly outside the hull
boundary layer and the spacing becomes close to the
natural wavelength of the Kelvin wave pattern, which
makes numerical diffusion smears out the solution.
Typically a resolution of 30 to 40 cells over the smallest wavelength is needed to accurately capture the
shape and behaviour of the surface wave pattern. Also,
for the surface waves, time plays an important role
since the time accurate solution produced by LES
needs time to transport information through the domain. Thus, the Kelvin wave pattern gradually builds
up starting with the bow and stern waves.

Figure 5: Comparison of computed (top) and experimental (bottom) wave contours around the DTMB
5415 hull at Fn=0.28 and Re=12.6 106.

Figure 4: Perspective view of the stern of the DTMB


5415 hull at Fn=0.28 showing randomly released
streamlines coloured with turbulent kinetic energy and
two slice planes with secondary velocity vectors showing the vortical structures in the wake.

In figure 5, the computed wave pattern is presented and compared with experimental data (LES at
the top of the figure and experimental data below) using an iso-surface of the volume fraction variable at
=0.5, coloured wave height, h/L. The smearing of
the divergent waves and the stern waves due to numerical diffusion, grid stretching and comparatively
large grid spacing is easily seen, but still the amplitudes and positions of the primary wave crests close to
the hull are in good agreement with the experiments.
The differences in wave height also depend on the
value of chosen for the iso-surface. Due to the continuous constraint on the volume fraction field, the

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

values of varies from 0 to 1 over three grid cells,


hence, on this mesh over a distance of up to 0.03 m.
This is significant considering that the amplitude of the
bow-wave is only 0.10 m. Note, however, the very
good agreement in relative location, phase speed and
shape of the primary waves.
Figure 6 shows an overview of the hull and
the wave field with lines marking the position for the
wave-cuts in figures 7-9. The line closest to the hull is
the line y/L=0.096, the second line is y/L=0.172 and
the third line is y/L=0.324.

some deviations in amplitudes. In general, the results


agree reasonably well with the experiments.
Figure 8 shows the wave profile at a fixed
distance (Y/L=0.172) form the hull centreline. This is
the location for comparison of wave profiles in the
Gothenburg 200 workshop. The first wave crest is still
too high and the second too low. However the phase
and the first wave trough are accurately predicted. The
grid cells are clustered towards the hull to resolve the
boundary layer and the grid becomes coarse rapidly.
At this distance from the hull the waves are still predicted reasonably. This result is comparable with some
of the results from the Gothenburg 2000 workshop.
Its not as good as the best but still not worse than the
worst.
0.006
DTMB exp
LES
0.004

Figure 6: Overview showing the location of the wavecuts in figure 7-9.

Waveheight z/L

0.002

-0.002

-0.004
0.006
DTMB exp
LES
-0.006
0.004

-0.008
-1

Waveheight z/L

0.002

-0.5

0.5
1
Distance x/L

1.5

2.5

Figure 8: Experimental and predicted wave-cuts along


a line at y/L=0.172.

-0.002

-0.004

-0.006

-0.008
0

0.2

0.4

0.6

0.8
Distance x/L

1.2

1.4

1.6

Figure 7: Experimental and predicted wave-cuts along


a line at y/L=0.096 for the DTMB 5415 hull at
Fn=0.28 and Re=12.6 106.
Figure 7 shows the wave profile at a fixed
distance (y/L=0.096) from the hull centreline. From
these figures the differences in amplitudes become
more evident, and also in the locations of the wave
crests. The rising of the bow wave is somewhat delayed which is probably caused by the large grid spacing in front of the bow that unable the surface to react
on the gradual pressure change. This may also explain
the steepness and over prediction of the bow wave
amplitude. Further downstream the wave profile predictions agree qualitatively well with the experiments,
especially the location of crests and troughs, but with

Figure 9 shows the wave profile at a fixed


distance (Y/L=0.324) form the hull centreline. The
coarse grid affects the waves here. Both the wave amplitudes and wave lenghts differ from the experimental
data. The grid in the present computations was designed to resolve the boundary layer and the wake, not
to resolve the entire wave and flow field. Therefore,
the computed wave profile is not expected to reproduce the experimental wave profile here.
In figure 10, the time-averaged normalized
axial velocity, v x /v at x/L=0.935 is shown. An
overview of the computed results showing both the axial velocity contours and secondary velocity vectors is
shown in figure 10a. The line z/L=-0.02 used in figure
11 is also included here. A closer view of the axial velocity of the computed results in figure 10c and experimental data in figure 10b. At this section, close to
the stern, several flow features combine to form the
wake. Among the most eminent features are the vortex
pair from the sonar dome, the hull boundary layer, and
the upward-inward flow from the rising of the stern.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

0.006
DTMB exp
LES
0.004

Waveheight z/L

0.002

-0.002

-0.004

-0.006

a
-0.008
-1

2
3
Distance x/L

Figure 9: Experimental and predicted wave-cuts along


a line at y/L=0.324.
Due to the cross-sections of the hull, the bilge vortex
and its associated flow pattern is comparatively week.
Moreover, for the computed sub-surface wake velocities the amplitudes are in good agreement with the experiments but there are some deviations in the location
and strength of the main vortical structures. The redistribution of the velocities in the boundary layer due to
the sonar dome vortex pair is more concentrated in the
computations than in the experiments. Thus, the influence of the vortices is over predicted while RANS
computations often under predict the strength and effect of vortices (Larsson et al 1990, Kodama 1994,
Larsson et al 2000, Svennberg 2001). The grid spacing
at the location of the vortex is too coarse to resolve all
scales affecting its development and interaction with
the boundary layer. Also, secondary vortical structures
that affect the position of the sonar dome vortices as
well as the boundary layer are lost due to the coarse
grid. This gives a less complex flow field with fewer
and stronger vortices and reduced interaction between
them. The computed boundary layer is also thinner
than the boundary layer in the experiment. This has
been observed in other computations using the same
OEEWM model, cf. (Wikstrm et al 2003). One reason might be the studs used in the experiment at the
bow of the hull to trigger a turbulent boundary layer.
Such stripping device is not included in the calculations.
The prediction of vortex location is a very
difficult task where a small deviation in vortex
strength and distance to the hull immediately changes
its transverse position at a certain longitudinal position. This mechanism is also closely related to the prediction of the hull boundary layer, which is mainly affected by the capability of the subgrid model to handle
simultaneous grid and flow anisotropies. The wall
model is based on the law of the wall that is valid for

c
Figure 10: a: Normalized axial velocity and secondary
velocity vectors, overview with the line z/L=-0.02
used in figure 11, b: Contours of experimental, and c:
computed, normalized axial velocity, v x /v, at
x/L=0.935 for the DTMB 5415 hull at Fn=0.28.
flat surfaces without pressure gradients, while the hull
form is far from a flat surface. Therefore, the wall
model should not be expected to give completely correct predictions here. As in the experiments the boundary layer is relatively thin close to the keel and the free

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

surface whilst being thicker at mid girth. Still, the


computed mean velocity field shows the same overall
behavior and magnitudes as the experimental results.
Figure 11 show velocity profiles along the
line z/L=-0.02. This line was used for validation at the
Gothenburg 2000 Workshop (Larson et al 2000). The
velocity profiles at a fixed location as this are extremely sensitive to the location of the vortices. The
vortices are located somewhat to far from the keel and
somewhat closer to the hull than in the experiments.
This deviation may explain the difference between the
computed and experimental velocity profiles here.

1
INSEAN exp
LES
0.95

U/Uinf

0.9

0.85

0.8

Acknowledgements

0.75

0.7
0

0.01

0.02

0.03

0.04
Distance y/L

0.05

0.06

0.07

0.08

This work is supported by the Swedish Defence Materiel Administration under the grant MarinLES.

References

0.25
V/Uinf INSEAN
V/Uinf LES
W/Uinf INSEAN
W/Uinf LES

0.2

Alin N., Fureb C. and Svennberg U., LES of the Flow


past Simplified Submarine Hulls, Proceedings The 8th
International Conference on Numerical Ship Hydrodynamics, Busan , Korea, 2003.

0.15

V/Uinf and W/Uinf

side the boundary layer was also too coarse to allow


secondary vortices. This is the case at the rising of the
stern, where the sonar dome vortices moves away from
the hull. The flow field is not perfectly predicted but
the results are comparable with some of the RANS
predictions at the Gothenburg 2000 workshop (Larsson
et al 2000). The motivation to use LES instead of easier and cheaper RANS methods is that LES methods
give the large scale time dependent fluctuations while
RANS only give the time averaged flow field. The
large scale Fluctuations are generally the most energy
containing scales and responsible for the dominating
part of the flow induced signatures. The present calculations are based on the experience from earlier computations including the free water surface on coarser
grids with other subgrid models (Lillberg and Fureby
2001) and experience from other hydrodynamic computations (Alin et al 2003, Bensow et al 2004). The
investigation will be continued with a finer grid, test of
other subgrid models, other numerical schemes for the
VOF method and propulsion modelling using actuator
discs.

0.1

Bagget J. S., Jimnez J. And Kravchenko A. G.,


Resolution Requirements in Large Eddy Simulations
of Shear Flows, Ann. Res. Breiefs, CTR, NASA
Ames/Stanford University, 1997, p 55

0.05

-0.05

-0.1
0

0.01

0.02

0.03

0.04
Distance y/L

0.05

0.06

0.07

0.08

b
Figure 11: Axial (a) and secondary (b) velocity profiles along the line z/L=-0.02 in the propeller plane.

Concluding Remarks
A computation, from an ongoing research project, of
the flow field around the DTMB 5415 model hull including the free water surface has been presented. The
computational grid was concentrated around the hull to
resolve the boundary layer and the near field wake.
The wave pattern was therefore predicted with reasonably accuracy only close to the hull. The grid out-

Bensow R. E., Persson T., Fureby C., Svennberg U.


and Alin N., Large Eddy Simmulation of the Viscous
Flow around Submarine Hulls, Proceedings 25th
Symposium on Naval Hydrodynamics, St. Johns,
Newfoundland an Labrador, CANADA, 2004.
Boris, J.P., Grinstein, F.F., Oran, E.S. & Kolbe, R.L,
New Insights into Large Eddy Simulation, Fluid
Dyn. Res., 10., 1992, p 199.
Ferziger J.H. & Leslie D.C.;, Large Eddy Simulation
A Predictive Approach to Turbulent Flow Computation, AIAA paper 79-1441, 1979.
Fureby, C., Tabor, G., Weller, H. & Gosman, D., On
Differential Sub Grid Scale Stress Models in Large
Eddy Simulations, Phys. Fluids, 9, 1997, p 3578

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Fureby, C. & Almstrom, H., Hull Induced Noise. Part


1: Hydrodynamic Noise Produced by the Turbulent
Flow around a Rigid Hull, FOI-R--0130SE, 2001a.
Fureby, C., Towards Large Eddy Simulation of Complex Flows, Proc. EUROMECH Colloquium 412 on
LES of Complex Transitional and Turbulent Flows,
Kluwer Press, 2001b.
Fureby C., Towards Large Eddy Simulation of Complex Flows, In Direct and Large Eddy Simulation IV,
Eds. Friedrich R. & Rodi W., Kluwer, The Netherlands, 2001c

Lillberg E. and Fureby C., Towards Large Eddy


Simulations of Free Surface Ship Hydrodynamics,
Proceedings, Hydrodynamics and Aerodynamics in
Marine Engineering, HADMAR 2001, Bulgaria, 2001,
p. 205-214.
Lions P.L., Mathematical Topics in Fluid Mechanics, Oxford Science Publications, Oxford, 1996.
Model 5415 web page:
http://www50.dt.navy.mil/5415/5415n.html
Sagaut, P.: Large Eddy Simulation for Incompressible
Flows, Springer Verlag, Heidelberg, 2001.

Fureby C., Alin N,. Wikstrm N,. Menon S., Persson


L., and Svanstedt N., On Large Eddy Simulations of
High RE-number Wall Bounded Flows, AIAA Paper
No 2003-0066 (invited) 2003

Sussman, M., Smereka, P. & Osher, S., A Level-set


Approach for Computing Solutions to Incompressible
Two-Phase Flows, J. Comp. Phys., 114, 1994, p 146.

Gaskel, P.H. and Lau, A.K.C., Curvature Compensated Convective Transport: SMART, A new Boundedness Preserving Transport Algorithm, Int. J. Numer. Methods Fluids, 8, 1988, p 617.

Svennberg U., On Turbulence Modelling for Bilge


Vortices: A Test of Eight Models for Three Cases
Ph.D dissertation, Dept. of Naval Architecture and
Ocean Engineering, Chalmers University of Technology, Gteborg, Sweden, 2001, ISBN 91-7291-003-8

Hirsch C., Numerical Computation of Internal and


External Flows, J. Wiley & Sones, 1999
Kodama Y., Ed., CFD WORKSHOP TOKYO 1994,
Proceedings, Ship Research Institute, Tokyo, Japan,
1994
Lafaurie, B., Nardone, C., Scardovelli, R., Zaleski, S.
& Zanetti, G., Modelling Merging and Fragmentation
in Multiphase Flows with SURFER, J. Comp. Phys.,
113, 1994, p 134.

Ubbink, O., Numerical Prediction of Two Fluid Systems with Sharp Interfaces, PhD Thesis, London
University, 1997.
Wikstrm N., Svennberg U., Alin N., and Fureby C.,
Large Eddy Simulation of the Flow Past an Inclined
Prolate Spheroid, proceedings of TSP 3, Sendai, Japan, 2003

Larsson L., Ed., SSPA-ITTC Workshop on Ship


Boundary Layers, Proceedings, SSPA Report No. 90,
1980, Gteborg, Sweden.
Larsson L., Patel V. C. and Dyne G., Eds., 1990
SSPA-CTH-IIHR Workshop on Ship Viscous Flow,
Proceedings, Flowtech International AB, Research
Report No. 2. 1991 Gteborg, Sweden.
Larsson L., Stern F., Bertram V., Eds., Gothenburg
2000 A Workshop on Numerical Ship Hydrodynamics, Proceedings, Department of Naval Architecture
and Ocean Engineering, Chalmers University of Technology,
Gteborg,
Sweden,
2000,
http://www.iihr.uiowa.edu/gothenburg2000/index.html
Leonard, B.P., The ULTIMATE Conservative Differencing Scheme Applied to Unsteady One-dimensional
Advection, Comp. Meth. Appl. Mech. & Eng., 88,
1991, p 17.
Lesieur M and Metais O, New Trends in Large Eddy
Simulations of Turbulence, Annu. Rev. Fluid Mech.,
28, 1996p 45.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Takanori Hino
National Maritime Research Institute, Japan
Did you compare the resistance of the ship
with the experimental data?
Do you employ the sub-grid scale modeling
for the transport equation of VOF?
AUTHORS REPLY
No we have not computed the resistance.
This is naturally a parameter of interest for any
surface ship simulation, but not a primary objective
for this project. Also, no boundary layer profiles are
available experimentally which makes it difficult to
do any estimates on coefficients based on the wall
shear stress in the simulations.
No there is no sub-grid modelling for the
VOF transport equation. This volume fraction field is
a passive scalar field convected with the local
velocity field, which in turn, of course, is affected by
the SGS-model. The VOF transport equation is not a
physical part of the N-S equations but only a method
for capturing a discontinuous interface. The CICSAM
scheme is constructed in such a way as to minimize
diffusion and dispersion errors while preserving a
bounded solution.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Jacek Pawlowski
TRDC Inc., NL, Canada
Would you be able to indicate how much
computing effort in you computations typically is
committed to the initial transient response, and how
much is spent on free surface resolution?
AUTHORS' REPLY
The time to initiate the free surface wave
pattern is typically of the same order as the time
spent to develop the LES solution, at least for the
near field. Typically this is 2-5 flow past times for
any unsteady simulation. That is, there would be no
significant increase in the overall computational time
due to the development of the Kelvin wave pattern.
However, the computational effort in each time step
is increased by 5-10% for the free surface solution
with the method used, i.e. volume of fluid using the
CICSAM scheme.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Sung-Eun Kim
Fluent Inc., USA
1. The mesh is too coarse to capture largescale structure so what the authors did is essentially
use an algebraic turbulence model.
2. I wonder what justifies the use of LES for
these sort of applications where the boundary layers
are thin and there arent many significant large-scale
structures that impact significant hydrodynamic
performance of naval ships.
AUTHORS REPLY
Well, the grid is not too coarse to capture the
large scale structures, but it is too coarse to resolve
all important scales for their correct development.
However, The sub grid model used is a one equation
model, i.e. If the grid is far too coarse and the grid
and time dependency is ignored it will not end up
being a algebraic model it will be an one equation
RANS model. Furthermore we have seen worse
results from more advanced RANS model. Therefore,
we are not totally unhappy with the results.
The primary motivation for these
computations are the interest in wake physics and the
related signature aspects including propulsion and
manoeuvring. These phenomenon are not possible to
capture using RANS or URANS. The grid is fine
enough to capture the necessary large scale physics in
the wake using LES. But, the grid used here is not
sufficiently fine for a completely accurate flow field.
However, these computations are among our first
steps from academic test cases towards complex free
surface ship simulations of engineering interest for
the navy. We are currently working on refined grids
to improve the results. The signature computations
are generally not allowed to be published in the open
scientific literature but the underlying CFD results
are.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Michel Visonneau
Ecole Centrale Nantes/CNRS, France
The results concerning the free-surface
distribution indicate that the grid is too coarse to
capture accurately the free-surface elevation. Even if
the compressive CICSAM scheme is used, one
should build a region of refined grid where the
interface between air and water is supposed to be
located. Could you comment on that?
AUTHORS REPLY
Yes, we are aware about that the grid
resolution is not sufficient to resolve the free-surface
elevation. However, the project funding these
computations is primarily interested in the wake
physics. Therefore the grid is concentrated to the
boundary layer and the wake region. Our
computational resources were not sufficient to also
resolve the free-surface region but we are expanding
our computational resources and are working on a
finer grid. Also, local grid refinement could of course
solve the problem with the interface resolution, but
for transient simulations such as maneuvering, which
is the goal of this project, adaptive grid refinement is
very expensive.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics


St. Johns, Newfoundland and Labrador, CANADA, 8-13 August 2004

Computation of free-surface viscous flows at model and full


scale by a steady iterative approach
Hoyte C. Raven, Auke van der Ploeg, Bram Starke
(Maritime Research Institute Netherlands (MARIN))
ABSTRACT
This paper describes progress in the development of a
method to compute viscous flow with free surface
around a ship hull. The method solves the steady
RANS/FS problem by iteration, using the formulation
proposed by Raven & Van Brummelen (1999). It is
being developed to provide a more efficient alternative
for the usual time-dependent solution approaches, in
order to permit practical application on dense grids in
ship design work.
Applications are shown to a Series 60 hull at
model and full scale; and to a tanker model. Computed
scale effects on the wave pattern are confined to the
stern wave system, as expected. Good agreement with
experimental data is obtained; in particular, there is
little sign of numerical wave damping even further
from the hull. Computation times are quite modest.
A theoretical analysis of the numerical
accuracy is carried out. For the standard version of the
method, this shows that it has a numerical damping of
wave components which is of third order in the mesh
size; and a second-order numerical dispersion. This
explains the high accuracy found. Based on the analysis
an alternative scheme is designed which further
improves the accuracy.
INTRODUCTION
The development of solution methods for the problem
of viscous flow around a ship hull, including the wave
pattern at the free surface, is one of the main topics in
numerical ship hydrodynamics today. Successive
symposia and workshops, e.g. the Tokyo workshop in
1994 and the Gothenburg workshop in 2000, have
indicated the substantial progress made, which holds
promise for application of such computations in
practical ship design. The combined computation of the
viscous flow around the hull and the wave pattern is a
step beyond the traditional, separate consideration of
both aspects of the flow. It takes into account:
wave effects on the viscous flow,
viscous effects on the wave making,
and all the mutual influences between these.
Wave effects on the viscous flow take some
different forms. The wavy motion on the free surface

affects local flow directions and pressures along the


hull. E.g. the inflow of the propeller can be affected by
the wave making; wave slopes cause a vertical flow
component, wave crests and troughs cause an
additional longitudinal velocity. Also, the free surface
modifies the pressure variations along the hull, often
increasing the extremes and gradients and sometimes
causing a wave-induced flow separation near the
waterline. Since non-linear free-surface potential flow
solvers can accurately predict the wave making along
most of the hull, these effects can be largely taken into
account by first computing the inviscid wave pattern,
and then solving the viscous flow in the domain under
that surface. This so-called 'composite approach' was
introduced and validated in (Raven and Starke, 2002),
and is now being used routinely in practical viscousflow computations carried out at MARIN.
The inverse 'interaction' effect, the viscous
effect on the wave making, is essentially confined to the
afterbody and stern wave system in most cases. In
general, viscosity reduces the stern wave amplitude and
can even cause a change of flow regime for transom
sterns. Taking into account these effects in predictions
would permit a further optimisation of transom stern
shapes, a better assessment of viscous and scale effects
on the wave making and wave resistance, and
ultimately a better prediction of resistance and power at
full scale.
To compute viscous effects on the wave
making a solution of the complete problem of viscous
flow with free-surface boundary conditions is needed,
and this forms the subject of the present paper.
State of the art and objectives
A variety of solution methods for viscous flow with
free surface (RANS/FS methods) is being developed.
The general picture emerging from the literature
indicates that:
The wave profile along the hull is usually
predicted well by these methods. However, wave
cuts at a distance from the hull often show a
serious underestimation of the wave amplitude.
Grid refinement studies indicate that the wave
amplitude is often far from being grid independent.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Most computations published so far are for rather


high Froude numbers. However, practical
application will ask for computations at much
lower Froude numbers as well, since viscous
effects are larger for fuller ships at lower speeds.
The need to resolve the resulting shorter waves
will ask for even denser grids.
Almost all computations are just for model scale.
This has limited benefit for practical design, since
there is no known way to 'extrapolate' the viscous
effects on the wave making from model to full
scale.
The usual time-dependent solution approach
frequently causes problems, such as persistent
time-dependence and, in surface-fitting methods,
the need to update the grid thousands of times.
Quite substantial calculation times are usually
needed, and these actually prevent or complicate a
more widespread use in merchant ship design
today.
Therefore, a central problem is that much
denser grids are clearly needed to increase the
accuracy, but that would lead to a dramatic further
increase of the computation time. This difficulty was
already outlined in (Larsson et al., 1998). Based on the
estimate that a grid density of 50 cells per wavelength
would be needed near the free surface, they concluded
that the total number of cells would be in the range 106
- 107 at Fn=0.3, and 107 - 108 at Fn=0.15. These are
huge numbers to present standards.
The principal objectives of our developments are
thus as follows:
1. Develop a solution technique for the RANS/FS
problem that requires substantially less
computation time than present methods, and that
has a lower complexity of the computation
process;
2. Achieve a better accuracy of the wave pattern
prediction, also at a distance from the hull; by
using dense grids and refinements of the method;
3. Achieve accurate computations for full scale, and
accurate estimations of scale effects on the wave
making.
The present paper addresses these topics. In
previous papers (e.g. Raven and Van Brummelen,
1999; Van Brummelen et al., 2001; Raven and Starke,
2002) we have demonstrated the possibility to solve the
RANS/FS problem in a completely steady form by an
iterative approach, for some simple applications. This
development has been pursued, and the method has
now been extended and applied to some different ships
hulls. The method is found to be most efficient and
thereby meets the first objective.
Furthermore, the paper describes numerical
verifications and comparisons with experimental data.
These indicate quite a good accuracy of the wave

pattern even away from the hull. A theoretical analysis


has been carried out, which shows that the method used
contains a numerical damping of third order, and
indicates further possibilities for improvement. This
shows that a careful analysis may be a more efficient
option than just using denser grids.
In addition, this paper describes a computation
for full scale and compares with the corresponding
model scale computation. This is one of the first fullscale RANS/FS computations published, and the results
are quite plausible. A further full-scale computation by
the present method, for a tanker, is published in
(Starke, 2004). In the EU-funded project EFFORT
(Verkuyl and Raven, 2003), which focuses on
validation of full-scale viscous flow computations, such
computations have now been made by several
participants, and the project seems to bring about the
intended shift of attention to the real full-scale
situation.
The paper is set up as follows. First we briefly
introduce the RANS code used, focusing on some
aspects relevant for the free-surface treatment. Then we
derive and discuss the steady iterative approach, its
implementation and its behaviour in practice. The next
section then describes its application to some different
cases and compares with experimental data.
Subsequently, we briefly explain the theoretical
analysis of the numerical accuracy, the derivation of
numerical damping/dispersion; and derivation of
improved discretisation schemes.
THE RANS SOLVER
The RANS code used is PARNASSOS, a method
developed and used by MARIN as a tool for quality
assessment of hull designs (Hoekstra, 1999). It solves
the discretised Reynolds-averaged Navier-Stokes
equations for a steady, 3D incompressible flow around
a ship hull, supplemented by a turbulence model. Of
the various isotropic eddy viscosity turbulence models
available (Ea and Hoekstra, 2000) we have used the
one-equation model by Menter (1997), with a
correction for longitudinal vorticity according to
Dacles-Mariani et al. (1995). Body-fitted, generally
non-orthogonal HO-type grids are used, which are
strongly stretched towards the hull. No wall functions
are used, not even for the full-scale computation.
Most boundary conditions are conventional;
i.e. no-slip conditions at solid walls, symmetry
conditions at the ship's centreplane; Dirichlet boundary
conditions for the velocity components at the inlet
plane. However, at the external boundary we impose
Dirichlet conditions for the pressure and for the
velocity components tangential to the boundary,
derived from a free-surface potential flow calculation
using the RAPID code (Raven, 1996). At the outlet
plane, we take Neumann boundary conditions for the

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

pressure and the velocities, the former again derived


from the inviscid-flow computation.
The discretisation is of finite-difference type.
All terms in the momentum and mass conservation
equations are discretised at least second-order
accurately. Central schemes are used for the grid metric
and diffusive terms. For the convective terms and in the
continuity equation we use second-order upwind
schemes in streamwise direction, and third-order
schemes for the normal and girthwise direction. For the
gradients of the pressure in the momentum
conservation equations we use third-order schemes.
The PARNASSOS method has some distinguishing features that play a role in the free-surface
approach to be described, and are, therefore, briefly
discussed now.
Multiple-sweep marching
Reduction of the size of the equation system is obtained
by exploiting the character of the problem, with a
predominant flow direction. A solution procedure is
used in which the linearised equations are solved for
subdomains that consist of several streamwise stations
at a time. It is possible to choose these subdomains in
the range from one plane to even the complete domain.
All variables for a subdomain are solved
simultaneously. The subdomains are addressed in a
downstream sequence.
For stability and consistency, the bias of the
pressure gradient scheme must be opposite to that of
the velocity derivatives in the continuity equation, and
this requires a downstream-biased pressure derivative
in main stream direction. Necessarily, pressure values
in locations downstream of each subdomain (and the
longitudinal diffusive terms) are modelled explicitly,
i.e. they are taken from the previous downstream
sweep. Therefore the downstream marching process
must be repeated until the solution has converged. By
this multiple-sweep marching, both the non-linearity
and the elliptic character of the RANS equations are
recovered. Between the downstream sweeps, simple
algebraic upwind sweeps are performed in which only
the pressure is updated. This improves the convergence
properties. In the sequel of this paper, the process of
alternating downstream and upstream sweeps is
referred to as the global iteration.
Fully coupled formulation
In most other methods, the fact that in an
incompressible formulation the pressure does not occur
in the continuity equation motivates a reformulation of
the continuity equation, e.g. as a Poisson equation for
the pressure, a pressure-correction equation or a timedependent artificial-compressibility form. However, the
absence of the pressure from the continuity equation is
only a problem if one uncouples the equations. Instead,

in our code the full coupling is maintained. Since we


will show below that the coupling is important for the
implementation of our free-surface approach, and in the
literature the term full coupling can denote something
else, we briefly describe our approach. After
discretisation and linearisation, the three momentum
equations and the continuity equation give rise to a
matrix equation containing 4*4 blocks, containing all
elements that multiply the velocity components and
pressure in the same grid point. We use preconditioned
GMRES (Saad and Schultz, 1986) to solve this matrix
equation, using an incomplete LU-factorisation as a
preconditioner. Since the incomplete factorisation is
performed on a 4*4 block level, the construction of the
preconditioner does not break down as long as the pivot
blocks do not get singular. Hence the zero on the main
diagonal due to the absence of the pressure in the
continuity equation does not give any problems.
This fully coupled solution has been found to
be robust and quite insensitive to the mesh aspect ratio.
This permits to easily carry out even full-scale
calculations without using wall functions. More about
solution and preconditioning can be found in (Van der
Ploeg, Ea and Hoekstra, 2000).
Grid sequencing
In order to obtain a faster propagation of pressure
influences in upstream direction we use so-called gridsequencing, i.e. we start the computation on a mesh
coarsened in the mainstream direction and refine
repeatedly by a factor of 2. Particularly on grids that
are dense in main stream direction, it is a time-saving
and convergence-accelerating technique. In the sequel
we denote the grids with a j-fold mesh size as the
istepj-grid, for j=8,4,2 and 1.
THE STEADY ITERATIVE FORMULATION
The approach used for the incorporation of the free
surface is completely different from what other
methods use, being steady instead of time-dependent.
Therefore it will be described at some length and will
be contrasted with the time-dependent approach.
Free surface boundary condition
The method is of a free-surface fitting type; the upper
boundary of the computational domain coincides with
(an approximation of) the wave surface all the time,
and therefore needs to be updated repeatedly. On this
upper boundary, free-surface boundary conditions
(FSBC's) are imposed. If we denote the velocity
components (in a (x,y,z)-coordinate system fixed to the
ship, with x positive aft and z upward) by u,v,w, the
wave height by (x,y), and non-dimensionalise all
quantities using ship speed U, a reference length Lpp,
and gravity acceleration g, the free-surface boundary
conditions are:

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

a kinematic condition that the free surface moves


with the flow;
(1)
t + u x + v y w = 0 at z =

a normal component of the dynamic condition,


requiring that at the surface the pressure is
atmospheric (p=0); neglecting surface tension and
viscous contributions this takes the form
(2)
Fn 2 = 0 at z =
in which =(p+gz)/(U2) is the non-dimensional
hydrodynamic pressure.
two tangential components of the dynamic
condition, requiring that no shear stress is exerted
on the water surface in two directions.

The time-dependent approach


The free-surface boundary conditions are non-linear
and must be imposed on a surface not known
beforehand. To solve this problem, virtually all other
methods for computing viscous free-surface flows
follow a time-dependent procedure: the time-dependent
form of the problem is integrated in time, starting from
an initial condition, and this process is continued until a
steady result is obtained.
In each time step:
I.
the RANS equations are solved subject to the
dynamic conditions imposed at the current
wave surface;
II.
next, the wave surface and grid are updated
using the kinematic free-surface boundary
condition, or a similar algorithm is applied for
the convection of the wave surface in surfacecapturing methods.
However, as is discussed in (Raven and
Starke, 2002), this procedure is the origin of various
difficulties. The time-dependent formulation involves
the physics of transient waves that often cause a
substantial delay of the process: e.g. slow formation of
the steady wave pattern, persistent time dependence,
reflection of waves at artificial boundaries, contact line
problems. Moreover, the uncoupling of kinematic and
dynamic conditions often imposes severe time step
restrictions, and only small updates of the wave surface
per time step are allowed.
The steady iterative approach
In order to avoid these drawbacks, we use the approach
first proposed in (Raven and Van Brummelen, 1999)
and derived in more detail in (Van Brummelen et al.,
2001). This method avoids all unsteadiness and solves
a strictly steady form of the problem directly by
iteration. The term t in (1) is omitted, and also in the
momentum equations all time-dependence is absent.
The process only admits wave solutions that satisfy the
steady dispersion relation, which excludes most of the
problems mentioned above.

With the omission of all time-derivatives, the


question rises how the solution algorithm should be set
up. One might think of alternatingly imposing the
normal dynamic and the kinematic boundary condition;
but such a scheme would suffer from a drawback it
shares with the usual time-stepping methods: it
uncouples the kinematic and normal dynamic
condition. None of these two conditions has any wavelike character by itself: wave solutions, a dispersion
relation, group velocity and all other properties of freesurface waves only arise from the combination of both
conditions. Uncoupling leaves the task of generating a
wave pattern to the iterative algorithm; and we believe
we can do better by already building it into our RANS
solution at each iteration.
The key to a successful iterative solution of
this steady RANS/FS problem is to impose a particular,
combined form of the kinematic and normal dynamic
FSBC. This is obtained by substituting the wave
elevation from the dynamic condition into the
kinematic condition, which yields:
(3)
Fn 2 ( u x + v y + w z ) w = 0 at z = .
This combined condition needs to be supplemented by
the three dynamic conditions, to give a set of conditions
that corresponds exactly with the original set.
An iterative solution procedure then is defined as:
In each iteration:
I.
the RANS equations are solved subject to the
combined condition (3) and the tangential
dynamic conditions, imposed at the current
wave surface;
II.
next, the wave surface and grid are updated
using the normal dynamic condition (2).
Upon convergence the pressure deviation, normal
velocity and shear stress vanish at the wave surface and
the solution of the steady RANS/FS problem has been
obtained.
The combined FSBC in itself demands neither
a zero normal velocity nor a constant pressure at the
estimated wave surface; but imposes a relation between
the normal velocity and the pressure field, such that the
velocity at the estimated wave surface is everywhere
parallel to an isobar surface. In (Raven and Starke,
2002) the correspondence of the combined condition
with e.g. the Kelvin free-surface condition is pointed
out. Without any iteration this already embodies the
proper wave-like behaviour, and as a matter of fact,
even without any free-surface update we usually find
fairly accurate solutions of the viscous flow with free
surface. The free-surface updates are needed for some
of the non-linear terms and add accuracy, but are not
needed to get the essence of the wave pattern.
Implementation of free-surface boundary conditions
Let (1,2,3) denote the curvilinear system which is

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Fn 2 V 1 1 + V 2 2 + V 3 3 w = 0 .

(4)

Imposing such a relation between the


velocities and the pressure is no problem, because we
use the fully coupled approach described above. From
theoretical considerations and practical experience it is
clear that the derivative of in 1-direction must be
modelled by a backward-biased difference scheme,
which fits nicely in the marching scheme described
above. We use a third-order 4-point scheme for it.
The contravariant velocity components are
taken from the previous step of the global iteration. For
the pressure derivatives only an upwind, first-order part
is taken implicitly. By using a deferred correction, at
the end of the global iteration the scheme becomes
third-order accurate. The term with V2 in (4) is treated
similarly. From numerical experiments it appeared that
using a second-order linearisation does not improve the
convergence behaviour and a more stable
implementation is obtained if the term with V3 is
neglected. This does not influence the finally computed
solution, since V3 is the velocity component normal to
the free surface which in a converged solution is zero.
The requirement that the tangential stress
components are zero, gives two more boundary
conditions. The derivatives in these equations are
discretised using a second-order 3-point scheme.
Although we can impose only three boundary
conditions in the RANS-equations, we need a fourth
equation at the free surface in order to close the system.
This fourth boundary equation has to be sufficiently
weak. Imposing that the derivative of in 3-direction
is zero, for example, is too strong and gives wrong
results. Therefore, we impose the momentum equation
in 3-direction, in which viscous terms have been
neglected. The robustness and convergence behaviour
have been found to be strongly influenced by the
discretisation and linearisation of this equation.

computed wave elevation can occur. This is illustrated


by Fig.1, which shows the computed wave elevation
near a ship's hull, before the first update of the grid, for
the Series 60 case to be discussed in the next section.
The strong variations in transverse direction
take the appearance of short, sharply diverging waves
which actually propagate into the wave pattern away
from the hull. These short waves can be computed
since the local mesh size in wall-normal direction is
extremely small, as required in order to capture the
boundary layer. Such waves should have a very quick
decay in vertical direction, but this decay cannot be
resolved since the vertical mesh spacing is much larger.
Therefore the short waves will be severely distorted,
which may explain their occurrence in the computation.
Moreover, they are irrelevant for the wave resistance. It
is, therefore, desired and permitted to adjust the
formulation such that they are suppressed or filtered
out, provided that this does not affect longer wave
components and the viscous flow around the hull.
In order to improve the convergence
behaviour we use a modification of the equation for the
pressure quite close to the waterline. In a small region
near this line, at the free surface only, we replace the
momentum equation in 3 -direction by the requirement
that the second derivative of in wall-normal direction
has to be zero. Hence we do not modify the three real
boundary conditions at the free surface, but only the
fourth equation which is required in order to close the
system. This has the desired effect and reduces the
short waves. For the Series 60 case it is possible to let
the computation converge without this modification,
which gives us the opportunity to study its effect on a
converged solution. Fig. 2 shows that the modification
hardly affects the solution.
0.5

0.4

0.3

y/Lpp

conforming to the boundaries of the domain. Herein 1


is roughly aligned with the streamwise direction, 2
with the wall-normal direction and 3 with the girthwise
direction. Near the free surface, the latter is also the
direction normal to the free surface. The transformation
relations between the Cartesian coordinates x, y, z and
1, 2, 3 are such that a uniform, rectangular mesh in
(1,2,3)-space with all mesh sizes equal to one defines
the grid in physical space. Let V1, V2 and V3 denote the
contravariant velocity component in, longitudinal, wallnormal and girthwise direction, respectively. Equation
(3) in contravariant velocity components reads

0.2

0.1

-0.5

Behaviour in the boundary layer


From numerical experiments it appears that if we
impose these boundary conditions in all points at the
free surface, the computation is hard to converge.
Strong variations in wall-normal direction in the

-0.4

-0.3

-0.2

-0.1

0.1

x/L pp

Figure 1: Isolines of computed wave elevation prior to


first grid update; showing short diverging waves near
the hull.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

0 .0 0 5

da sh e d (re d )- m o d ifica tio n n e a r w a te rline


so lid (b la ck)- no m o d ifica tio n
ste rn

z/Lpp

-0 .0 0 5

m id ship

-0 .0 1

0 .1

0 .2

0 .3

0 .4

y/L pp

Figure 2: Transverse wave cuts for the Series 60 case at


the stern and midship, comparing the results obtained
with and without modification near the waterline.

Free surface updates


In our method, a surface-fitting approach is used in
order to compute the free surface, i.e. in the final
solution the mesh has to be aligned with the free
surface. We usually start with a grid up to the
undisturbed water surface, and this is updated several
times until a converged solution has been obtained.
Each step of the free-surface iteration incorporates a
complete global iteration process, which can be
continued until a converged solution of the RANSequations has been obtained. In the next step of the
free-surface iteration the free surface is updated using
the computed pressure coefficient at z= and (2), and
a new mesh is generated beneath this new surface. The
usual elliptic grid generation tools are applied, keeping
the number of cells in all directions equal to that in the
previous step. The global iteration is then restarted
using the previous solution as an initial estimate of the
pressure and velocity field.
Solution strategy
Summarizing, we have four nested iteration processes:
The outer-most iteration is the grid-sequencing
described above. In this iteration the grid is refined
several times in main stream direction.
For each value of istep (each grid density), the
free-surface iteration described above is carried
out.
Within each free-surface iteration, the global
iteration to deal with the non-linearity and
ellipticity in the RANS-equations is performed.
As the inner-most iteration we use preconditioned
GMRES to solve the systems of linear equations.

This nesting was found efficient, providing already a


good approximation of the wave surface and pressure
field on the coarser grids, which require little
computation time. An additional benefit is that with
little more work a converged solution is obtained for
each value of istep, which enables a grid-dependency
study.
In the study of the speed of convergence of the
free surface iteration, discussed in the next section, we
have used a different approach: the free-surface and
grid updates were performed on the finest grid only.
This was found to be less efficient however. Another
alternative is to make a free-surface update after each
global iteration, as proposed by Lewis (2004), who
found this to reduce the computation time strongly.
There are still several other possibilities to reduce the
total CPU-time, e.g. by starting with an estimate of the
free surface computed by a free-surface potential flow
method. Also, better matching the required
convergence levels for the different processes and
stages of the computation probably can save much
time. This is left for future study.
APPLICATIONS
In previous publications on the present development,
viscous-flow computations with free surface were
applied to a 2D bottom bump, a 3D pressure patch and
the Series 60 hull. Extremely fast convergence of the
free-surface was shown for the bottom bump, for which
only nine mesh updates were sufficient, despite the
significant non-linearity of the wave system. In the case
of the pressure patch and the Series 60 no free-surface
updates were performed at that time. In the present
section results including free-surface updates will be
presented for two hull forms; the Series 60 and the
Dyne tanker. In the case of the Series 60 computations
are presented both for model scale and for full-scale
Reynolds number. This provides the opportunity to
study scale effects on the stern wave system. The Dyne
tanker is a more challenging test case due to its fuller
hull form and low Froude number. The model scale
computation is shown below, the full-scale
computation in Starke (2004). The results of both cases
are validated against available experiments.
Series 60
Results are presented for the Series 60 Cb=0.6 at model
scale Reynolds number Rn=3.4x106 and Froude
number Fn=0.316. The computational mesh consisted
of 321 x 121 x 45 grid nodes in the streamwise, wallnormal and girthwise direction, respectively, adding up
to 1.8M grid nodes in total. The grid density in
streamwise direction corresponds to approximately 103
grid cells per wave length. The governing equations
were integrated down to the wall, i.e. no wall-functions
were used.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

top of the stern wave reaches 94% of its final height at


the initial mesh. The second update shows an overshoot
of the stern wave before it reaches its final height.
Between these updates the location where the
maximum stern wave height occurs shifts slightly back
and forth. The convergence history for a wave cut at
y/Lpp=0.3507 is shown in Fig. 4. Here, the height of the
second wave crest increases monotonically from
56.5%, 82.7%, 94.2% to 98.8% of its final value
between the four successive free-surface updates, while
its location shows a slight shift upstream. After the
fourth update both the wave elevation and the wave
length remain practically constant.

S e r ie s 6 0
3 ,4 ,5
2
1

2
3 ,4 ,5

- 0 .5

F .P .P .

x / L pp

0 .5

A .P .P .

Figure 3: Convergence of the waterline in successive


free-surface updates.
S e r ie s 6 0
y / L pp = 0 .3 5 0 7

4 ,5
3
2
1
3 ,4 ,5
2
1

- 0 .5

F .P .P .

x / L pp

0 .5

A .P .P .

Figure 4: Convergence of the free-surface updates for a


longitudinal wave cut at y/Lpp=0.3507.

Convergence of the free-surface updates


Fig. 3 shows the convergence of the hull wave profile.
The computation started on a mesh beneath the
undisturbed free surface, indicated by the index zero
in the figure. In the result obtained on this mesh, a
wave elevation is predicted as indicated by the index
one. The bow wave reaches already 83% of its final
height. In the first free-surface update, the grid is
adjusted to match this predicted wave surface, and a
new solution is found (index two). In this second
update the bow wave height increases to 96.5%, and it
remains practically constant after the third update. The

The results confirm the extremely fast convergence that


can be obtained with the steady iterative formulation of
the free-surface boundary conditions. Already in the
first step, which can be regarded as a partial
linearisation of the free-surface boundary conditions
with respect to the undisturbed free surface, most of the
wave system has been captured. The following updates
merely serve to capture the remaining non-linear
effects on the wave system. To quantify the
convergence rate of the free-surface iteration, the
maximum changes in the free-surface elevation are
found to decrease by a factor 1.5 to 2.5 per iteration.
Grid-dependency and validation
To investigate the effect of discretisation errors on the
solution, results obtained on four meshes that were
successively refined by a factor two are compared. The
mesh is refined in the streamwise direction only, after it
had been observed in initial computations that the grid
resolution in this direction has a significant effect on
the solution. The number of grid nodes in streamwise
direction is equal to 41, 81, 161 and 321 on the meshes
that are referred to as istep8, istep4, istep2 and istep1,
respectively. At each of the meshes two free-surface
updates were performed, which proved to be sufficient
to reduce the maximum change in the free surface to
less than 6% of the maximum wave height at the
waterline and less than 0.5% of the maximum wave
height at a longitudinal wave cut at y/Lpp=0.3507. The
relatively large change at the waterline is related to
small changes in the predicted streamwise location of
the top of the stern wave.
Fig. 5 shows the variation of the waterline
with increasing mesh density. On the coarsest mesh,
istep8, the first wave trough and the following wave
crest which are located around mid ship are underpredicted. The result improves considerably on the
second coarsest mesh, istep4, and the waterline has
become practically grid independent on the second
finest mesh.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Only further downstream, beyond x/Lpp=0.85 there is


still a change in the wave height with increasing mesh
density. This may either be an effect of the local grid
density or an effect of the outflow boundary conditions,
a subject which is left for future research.

negligible. Although most notably around x/Lpp=0.5


small changes in the solution can still be observed with
increasing mesh density.
Also shown in these figures are experimental
results from Toda et al (1991). As expected the
predicted waterline shows good agreement with the
experiments, although the top of the second wave crest
near x/Lpp=0.1 appears to be located somewhat further
downstream in the computation. At y/Lpp=0.2395
excellent agreement is found with the experiments.
A more global picture of the comparison
between the computation and the experimental results
is given in Fig 7. Again, the general comparison is
quite good, with only locally small differences in the
shorter wave components. Furthermore it is evident that
our result does not suffer from any appreciable
numerical damping away from the hull.

S eries 6 0

0 .0 1 5

istep 4 ,2 ,1
istep 8
0 .0 1

z / Lpp

0 .0 0 5

istep 2 ,1
istep 4

istep 8

-0 .0 0 5

Parnassos

-0 .0 1

-0 .0 1 5

-0 .5

F .P .P .

x / L pp

0 .5

A .P .P .

Figure 5: Grid dependence of the waterline. The open


circles correspond to experiments by Toda et al.
S eries 6 0

0 .0 1 5

y / L p p = 0 .2 3 9 5
0 .0 1

istep 1
istep 2
istep 4

z / Lpp

0 .0 0 5

istep 8

Experiment

Figure 7: Validation of the free surface at model scale.

-0 .0 0 5

-0 .0 1

-0 .0 1 5

-0 .5

F .P .P .

x / L pp

0 .5
A .P .P .

Figure 6: Grid dependence of a longitudinal wave cut at


y/Lpp=0.2395. Experiments by Toda et al.

Fig. 6 gives the results for a wave cut at


y/Lpp=0.2395. On istep8 the wave amplitude is strongly
under-estimated, even though the wave length is
captured reasonably well. Upon grid refinement the
solution rapidly improves and differences between the
finest and the second finest mesh are practically

Computational effort
All computations have been performed on a PC with a
single Intel Pentium IV 2.4 GHz processor. As stated
earlier, in the grid study two free-surface updates were
performed on each of the four meshes. The CPU time
required for the RANS/FS computation on each mesh
was 2.1 hours on the coarsest mesh, 2.6 hours on the
second coarsest mesh, 7.7 hours on the second finest
mesh and 23.9 hours on the finest mesh, respectively,
adding up to 36.3 hours CPU time for the entire grid
study.
Scale effect on the wave system.
To study the Reynolds effect on the wave system the
flow around the Series 60 Cb=0.6 at full scale Reynolds
number Rn=8.4x108 and Froude number Fn=0.316 was

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

simulated numerically. The same case has been


computed by Schweighofer (2003). Compared to the
computation at model scale, the number of grid nodes
in wall-normal direction was increased to capture the
stronger gradients that occur in a boundary layer at full
scale. The computational mesh consisted of 321 x 161
x 45 grid nodes in the streamwise, wall-normal and
girthwise direction, respectively, adding up to 2.3M
grid nodes in total. Again no wall-functions were used.
The number of grid updates required to converge the
free-surface elevation was similar to the model scale
case.
Fig. 8 illustrates the scale effect on the wave
system of the Series 60. Differences between model
and full scale are confined to the region behind the
stern. Upstream both the inviscid and the viscous-flow
results coincide, thus confirming the absence of viscous
effects on the wave system in most of the domain.
Compared to the over-estimated stern wave in the
inviscid result the model scale viscous-flow
computation shows just the right reduction to the level
found in the experiments. At full scale the depth of the
first wave trough behind the stern, located near
x/Lpp=0.8 in this wave cut, lies between the model scale
result and the result for inviscid flow. This was to be
expected, since viscous damping of the wave system
will be smaller at full scale compared to model scale.
It is also evident from the results that the
magnitude of the scale effect is quite modest for the
Series 60, despite the relatively small Reynolds number
at which the experiments (and the model-scale
computation) were performed. This confirms the small
viscous-inviscid interaction effects for slender hull
forms, but it also indicates that the free-surface
elevation has to be predicted very accurately in a study
of scale effects.

D y n e ta n k e r

1
2
3 ,4
0

2
3 ,4
1

0 .0 1 5

2 ,3 ,4

y / L p p = 0 .0 7 5 5

z / Lpp

-0 .0 0 5

-0 .0 1

x / L pp

0 .5
A .P .P .

Figure 8: Scale effect on a wave cut at y/Lpp=0.0755.

- 0 .5

F .P .P .

x / L pp

0 .5

A .P .P .

Figure 10: Convergence of the free-surface updates for


a longitudinal wave cut at y/Lpp=0.2182.

F .P .P .

0 .0 0 5

-0 .5

0 .5
A .P .P .

y / L pp = 0 .2 1 8 2

0 .0 1

-0 .0 1 5

x / L pp

D y n e ta n k e r

in viscid flow

`h / L ;Y /L ;V 3
viscou s (m od el)
`h / L ;Y /L ;V 3
viscou s (sh ip)
`h / L ;Y /L ;V 3
experim ents (T od a et al.)
`h / L ;Y /L ;V 3

F .P .P .

Figure 9: Convergence of the waterline in successive


free-surface updates.

S eries 6 0

0 .0 2

- 0 .5

Dyne tanker
Most RANS/FS solutions published have been for
rather slender ships, e.g. the Series 60, the Kriso
container ship (KCS) or the DTMB model 5415.
However, just for fuller ships the interaction between
the viscous flow and the wave system is important.
Furthermore simulations at low Froude numbers
require grids that would lead to excessive CPU-times
for most methods. In this section results are presented
for the Dyne tanker at model scale Reynolds number
Rn=8.5x106 and a Froude number Fn=0.165. The ship
has a cylindrical bow, a block coefficient Cb=0.87 and
it was used as a test case in (Larsson et al., 1991). The

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

computational mesh in the present study consisted of


553 x 121 x 45 grid nodes in the streamwise, wallnormal and girthwise direction, respectively, adding up
to 3M grid nodes in total. The grid density in
streamwise direction corresponds to approximately 44
grid cells per transverse wave length. No wall functions
were used.
Convergence of the free-surface updates
Fig. 9 shows the convergence of the waterline. As in
the case of the Series 60 fast convergence of the free
surface is found. The solution obtained on the initial
mesh beneath an undisturbed free surface captures all
essential features of the wave system. The following
free-surface updates most notably result in an
adjustment of the magnitude of the wave crests and
troughs. The waterline remains practically unchanged
after the third mesh update. Similar fast convergence of
the free-surface elevation is found for a longitudinal
wave cut at y/Lpp=0.2182, as shown in Fig. 10.
Grid-dependency and validation
Fig. 11 compares the solution for several longitudinal
wave cuts with experiments available from Lundgren &

hman (1994). Although there are some differences the


general comparison is quite good. Especially the
magnitude of the stern wave system has improved
considerably compared to the inviscid-flow prediction.
An indication of the level of grid independence of the
present solution is given in Fig. 12. Here, two wave
cuts are shown, obtained on the present mesh (istep1)
and on a mesh which has been coarsened by a factor
two in streamwise direction (istep2). Although
reasonable agreement is found, especially the wave cut
furthest from the hull still shows an upstream shift of
the wave system and a reduction of the wave length.
This indicates the presence of non-negligible numerical
dispersion in the solution at larger distances from the
hull. Note, however, that the total number of grid nodes
in streamwise direction is already quite high: 553 in the
finest mesh. An improvement of the solution by adding
more grid nodes will most likely not be practical, due
to the increase of the computational effort, which is
already considerable in the present case. An alternative
way to improve the accuracy is the use of more
accurate discretisation schemes for the free-surface
boundary conditions. This will be investigated in the
next section.

D y n e ta n k e r

- 0 .5
F .P .P .

0 .5

0
x / L

pp

y / L

pp

= 0 .2 1 8 2

y / L

pp

= 0 .1 4 7 0

y / L

pp

= 0 .1 0 4 3

y / L

pp

= 0 .0 9 0 0

A .P .P .

Figure 11: Validation of the free-surface elevation at four wave cuts. Experiments by Lundgren & hman. Thick
solid line: viscous-flow computation, thin solid line: inviscid-flow computation.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

y / Lpp = 0.0900

istep1
istep2

-0.5

x / Lpp

istep1
istep2

0.5

y / Lpp = 0.2182

provides a single wave component. For the continuous


problem, this evidently represents a plane wave with a
wave number k0 = 1/Fn2, which is the steady 2D (or
transverse) wave. For the discretised formulation, again
a single wave mode is found, but with a slightly
different wave number that depends on the mesh
spacing. The difference between the wave numbers of
the continuous and discretised form determines the
numerical damping and dispersion.
We now will briefly describe the principal
steps made and give some of the results. We consider a
2D flow with free surface, in an x-z coordinate system.
Small perturbations of a uniform flow are considered,
and we linearise in these perturbations. The
perturbations are represented by Fourier components:

)
u
u

)
q w = w e kx + sz dkds

)


-0.5

x / Lpp

0.5

(5)

Figure 12: Grid dependence of the wave system. Solid


line: istep1, dashed line: istep2.
ANALYSIS OF NUMERICAL DISPERSION AND
DAMPING
A remarkable aspect of the results shown in the
previous section is the fact that the numerical damping
appears to be very limited; as observed both from the
grid refinement studies and from the comparison with
experimental wave cuts at larger distances from the
hull. While a visible numerical wave damping is rather
characteristic of RANS/FS computations, little of that
is observed here. The fact that the grid used is denser
than what is common, seems in itself insufficient to
explain the improvement.
To study the properties of the method we have
carried out a theoretical analysis of its numerical errors.
The analysis was proposed by Van Brummelen (2000),
who essentially derived the numerical dispersion and
damping of a variant of the present method already;
and was again applied to the present method by Lewis
(2004). Below, we extend the analysis, we analyze the
asymptotic behaviour, we demonstrate the role of
various difference schemes and derive a new
discretisation with even less numerical damping and
dispersion.
Derivation
In simple terms, the analysis considers a small 2D
perturbation of a uniform flow. This perturbation must
satisfy linearised RANS equations. A set of
eigensolutions of these equations is derived.
Substitution into the free-surface boundary condition

Substituting this into the linearised momentum


and continuity equations and changing the order of
operations based on the linearity we get
kx + sz dkds = 0
Lh q = Lh (k , s ) qe

(6)

in which the operator Lh represents the discrete


momentum and continuity equations.
The Fourier symbol of the operator, Lh , is
easily derived. The equations contain derivatives of the
pressure and velocities. For the continuous case, a
derivative with respect to x gives a factor ik in the
appropriate position in Lh , a derivative with respect to
z gives a factor s. However, in the discretised form
these are replaced by the Discrete Fourier Transforms
of the difference schemes. E.g. for the 3-point secondorder backward scheme in the momentum equations,

1 3
1 2

D x =
(7)
2e + e
2
x 2

where = k x . In general all these DFT's are functions


of and of = s z , i.e. functions of the number of
cells per wavelength. Since in our method several
different schemes are used, we need to define several
other DFT's as well:
D cx is the DFT of the d/dx scheme in the continuity
equation;
D x+ is the DFT of the forward d/dx scheme for the
pressure gradient;
and similarly for the z-differences.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Non-trivial solutions of the homogeneous


equations exist if the determinant of Lh vanishes. An
example of such a solution of the homogeneous
continuous problem is a steady transverse wave. It is
argued in (Van Brummelen, 2000) that the main
eigensolution of interest concerns the inviscid modes,
which are unaffected by the viscous terms in the
momentum equations. The corresponding zero of the
determinant provides an expression relating vertical
and horizontal derivatives:

D cz D z+ + D cx D x+ = 0

(8)

For the continuous problem, substituting the


Fourier symbols D x = ik and D z = s directly leads to:
s2 - k2 = 0. The root s = k correponds with the wellknown exponential decay of wave disturbances with
depth beneath the water surface. In the discretised
version, however, this relation still provides the vertical
wave number as s = s(k), but the relation is affected by
the finite mesh sizes, and thus depends on and .
The corresponding eigenmodes have a form
(Lewis, 2004)
+
u Dx

+ kx + s ( k ) z
w ~ Dz e

Dx

(9)

To find the eigenmode that also satisfies the FSBC, and


therefore is a solution of the homogeneous problem, we
substitute this into the free-surface boundary condition,
which produces:
D x , fs D x k0 D z+ = 0

(10)

in which D x, fs is the DFT of the difference scheme for

x in the combined free-surface condition, and k0 is


again the wave number of the continuous problem; as is
easily checked by filling in D x = ik and D z = s =k.
Using the relation for s(k) (8), and the condition (10),
we can solve for the ratio k/k0, i.e. the ratio between the
wave number predicted by the numerical method, and
the theoretical value.
Results
For given values of and we can now evaluate k/k0
numerically. This ratio in general is a complex number.
The real part represents a numerical
dispersion, i.e. an error in the dispersion relation that
links wave length and wave speed. In a numerical
solution of the steady problem, the numerical

dispersion will show up as an error in the length of a


wave which depends on the mesh size; or for a ship
wave pattern, possibly as an error in the direction of
wave crests.
The imaginary part is a numerical damping, if
the sign is positive; or a numerical amplification, in
case of a negative imaginary part. The waves are
damped by a factor exp(-2.Im(k/k0)) per wavelength.
The numerical damping will appear as an
underestimation of the wave amplitude at a distance
(aft or sideways) from the hull.
We shall now evaluate the wave number ratio
in two different ways: an asymptotic analysis for
vanishing mesh spacing, and a more general evaluation
for finite mesh spacing. In addition, the outcome has
been checked by grid refinement studies for a simple
test case.
In the asymptotic analysis, we take the limit
for 0, 0 , i.e. the number of cells per wave
length tending to infinity. By Taylor-expanding the
Discrete Fourier Transforms of all difference schemes
we obtain estimates of the leading-order numerical
dispersion and damping. In one of the studies we have
done this for various alternative difference schemes for
the x term in the combined free-surface condition;
keeping all other schemes the same. This yields the
following wavenumber ratio's for the complete
RANS/FS method (i.e. including the effect of all
difference schemes in the momentum and continuity
equations and the free-surface boundary condition):
for a 2-point first-order upwind scheme in the
FSBC:
1
1
1
k / k0 = 1 + 3 + 3 + O ( 4 )
2
12
12
for a 3-point second-order upwind scheme:
1
1
5
k / k0 = 1 2 3 + 3 + O ( 4 )
2
12
12
for the 4-point upwind-biased third-order scheme,
as used in the applications shown:
1
1
1
k / k0 = 1 2 3 + 3 + O ( 4 )
6
12
4
In the evaluation for finite mesh sizes, the
analysis was simplified by disregarding the effect of the
vertical discretisation, assuming that the d/dz schemes
introduce no errors. The result for the wave number
ratio is plotted in Fig. 13 to a base of kdx. For the real
part (top figure), the deviation from a horizontal line at
k/k0 = 1 indicates the numerical dispersion. The
imaginary part (bottom figure) is positive for all
schemes, indicating a numerical damping. The
behaviour confirms the results of the asymptotic
analysis but gives additional information on the
properties for finite mesh densities.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

To illustrate the use of the figure, let us


consider a computation on a grid with 50 cells per
wavelength. This means k.dx = 2 /50 = 0.126.

Re(k/k0)

1
0.8
0.6
0.4
0.2
0
0

k.dx

Im(k/k0)

1
0.8
0.6
0.4
0.2
0
0

k.dx

Figure 13: Wavenumber ratio's for various


discretisations. Top: real part (indicating numerical
dispersion); bottom: imaginary part (numerical
damping). Dashed line: first-order upwind scheme in
FSBC; dot-dash line: 2nd-order upwind scheme; grey
line: 4-point third-order scheme; solid black line:
balanced schemes in FSBC and pressure gradient.

Re(k/k0)

1
third order

0.9
0.8

balanced scheme

first order

0.7
second order

0.6

40

20

10

Gridnodes / wavelength
0.4

Im(k/k0)

0.3
third order

0.2

first order
second order

0.1

balanced scheme

40

20

10

Gridnodes / wavelength

Figure 14 Wavenumber ratio's for various


discretisations; behaviour for higher grid densities
(close-up of Fig. 13).

If a two-point upwind scheme is used in the


FSBC, we find k/k0 = 0.9960 + 0.0629 i, so there is a
0.4% error in the wavelength. The imaginary part
however means a loss of one-third of the amplitude per
wavelength!
The standard discretisation used above, with a
4-point scheme in the FSBC, results in k/k0 = 0.9975 +
0.00037i at this mesh density, so there is a 0.25% error
in the wavelength and an amplitude loss per
wavelength of just 0.23%.
Both evaluations clearly indicate that, in
combination with all other discretisations in the
method,
a 2-point upwind scheme for the free-surface
condition gives a drastic first-order numerical
damping, but little numerical dispersion;
the 3-point scheme reduces the numerical damping
to third order, but adds a second-order numerical
dispersion;
the 4-point scheme used in the applications shown
before has, on denser grids, a smaller numerical
dispersion, and a somewhat smaller third-order
damping. This explains the favourable properties
observed, in particular the small damping that
permits to come close to grid independence on
reasonable grid densities.
It is important to note that these conclusions cannot
simply be generalised to other RANS methods. E.g. the
low dispersion in the 2-point scheme is due to a
fortuitous canceling of errors, not a property of the
scheme itself. The details of most of the difference
schemes (used in momentum and continuity equations
and free-surface boundary conditions) play a role in the
analysis. What can be applied to other methods is the
analysis method, not its results.
A very interesting possibility is that now the
free-surface difference scheme can be designed to meet
certain requirements; similarly to what has been done
for a free-surface potential flow method in (Raven,
1998). From the analysis we could identify a particular
combination of d/dx schemes in momentum equations
and FSBC that makes several leading-order error terms
cancel. Combined with all other discretisations used in
PARNASSOS, for this balanced scheme the expression
for the wave number becomes:
1
k / k0 = 1 3 + O ( 4 )
12
i.e. the second-order numerical dispersion and thirdorder damping have been eliminated. Dispersion is
reduced to a small third-order term in dz, and numerical
damping is reduced to fifth order. Fig. 14 shows that
the scheme in particular has a far smaller numerical
damping. Even at 25 cells per wavelength, the

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

amplitude loss for this scheme would be just 0.032%,


the wavelength error 0.11%.
first order

Series 60 case. Comparison with Fig. 6 confirms that it


further improves the accuracy, and grid dependence is
now limited to the shorter, divergent wave near
x/Lpp=0.5. Similarly, Fig. 17 shows this for the Dyne
tanker, and demonstrates that the grid dependence that
was visible in the results for the standard discretisation
now has largely been removed.

second order

y/Lpp = 0.2395

third order

balanced scheme

istep1

istep8
istep2
istep4

x / Lpp
Figure 15: Grid dependence study for various
discretisations. Pressure patch, longitudinal cut at
y/Lpp=1, for grids with 6, 12, 25 and 50 cells per
wavelength.

Fig. 15 provides a confirmation of the analysis, by grid


refinement studies for a simple 3D test case of a freesurface pressure patch travelling at Fn=0.6; given by:
p = 0.05 exp[-4(x2 + y2)/Lpp2],
the same case as considered in (Raven and Starke,
2002). A longitudinal cut through the wave pattern is
shown, at a distance from the symmetry plane y/Lpp =1.
The finest mesh had 50 cells per wavelength. Results
are shown on the meshes used in the grid sequencing,
i.e. 6, 12, 25 and 50 cells per wavelength. The twopoint first-order scheme is very poor, giving still a
much too low wave amplitude for a grid density of 50
cells per wavelength. The second-order three-point
scheme is much better but has appreciable numerical
dispersion. The usual scheme (third-order) gives an
almost grid-independent result on the finest grids.
However, the winner is the new, balanced scheme,
which already produces quite reasonable results on the
istep4 grid (12 cells per wavelength!) and a (visually)
grid-independent solution with just 25 cells per
wavelength.
While the analysis was found to be extremely
useful, it is a simplification. In the near field, there are
additional resolution requirements and the analysis
gives only a partial assessment. Moreover, a ship wave
pattern contains diverging waves which were not
incorporated in the 2D analysis. As a first check, Fig.16
shows the behaviour of the balanced scheme in the

-0.5

0.5

x / Lpp
Figure 16: Grid dependence for a wave cut at
y/Lpp=0.2395. Series 60 model. Computation using
balanced discretisation schemes.

Dyne tanker, y/B=1.44, third-order scheme


hx

2hx

-0.5

x / Lpp

0.5

0.5

Dyne tanker, y/B=1.44, balanced scheme


2hx

hx

-0.5

x / Lpp

Figure 17: Grid dependence for a wave cut at


y/Lpp=0.2182, Dyne tanker, using third-order scheme in
FSBC, and using balanced difference schemes.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Therefore, this theoretical accuracy analysis


shows that there is much to gain by using higher-order
schemes, and the right combination of schemes; and is
a most useful tool to improve RANS/FS methods
without simply 'adding more cells'. If equal accuracy
can be achieved on a coarser grid, this means a vast
improvement in efficiency and applicability.
DISCUSSION AND CONCLUSIONS
Three main objectives of our development have been
mentioned in the Introduction: reduce computation
time, increase accuracy, and compute scale effects. The
work reported largely meets the first two objectives,
and contributes to the third one.
The steady iterative approach has been found
to perform very well for ship flows. While the
robustness of the process for ship applications still asks
attention, the absence of any transient waves removes
several problems reported for time-dependent methods.
The fact that only few grid updates are needed, means
that occasional user intervention in the grid updates
may be acceptable, which also eases the application to
complicated geometries.
The good accuracy of the wave pattern
prediction even at a distance from the hull was a
welcome surprise. The theoretical accuracy analysis
discussed in the last section explains the favourable
behaviour, and has indicated a new scheme in which
the accuracy is even better. This shows that there may
be other ways to improve the accuracy of RANS/FS
methods than only using denser grids.
The results for the flow at full scale indicate that it
is possible to predict scale effects on the wave pattern;
but they also stress that scale effect predictions with
good relative accuracy demand a very large numerical
accuracy of the wave pattern predictions in the first
place. Uncertainty analysis is required in order to
control this accuracy.
Further development of our method is needed.
Improvement of the robustness of the method is desired
in order to make its practical use more reliable. Also,
an extension to modeling dry and wetted transom stern
flows is required and has been undertaken.
The principal conclusions of this paper can then be
summarised as follows:
1. The steady iterative approach, based on the use of
a combined free-surface boundary condition, was
found to solve the steady free-surface ship viscous
flow problem efficiently. In particular, the freesurface iteration converges extremely fast.
Computation times are quite limited, even on a
single-processor PC.
2. Computations for model and full scale have
indicated scale effects on the wave pattern, which
were confined to the stern wave system. The full-

3.

4.

scale wave pattern was found to be intermediate


between the model-scale pattern and that computed
by a potential-flow approximation, as expected.
Good agreement with experimental data was
obtained. Also the wave amplitude at a distance
from the ship agrees well with the data. For a
tanker at Froude number of 0.165, fairly gridindependent results could still be obtained on a
grid of 3 million cells.
A theoretical analysis has indicated that the
method had a second-order numerical dispersion
and third-order damping. Based on the analysis, a
discretisation has been designed which reduces the
dispersion to third order and damping to fifth
order. Practical computations confirm almost all
conclusions from the theoretical study, and
indicate the drastic effect that the details of the
discretisation can have on the result.

ACKNOWLEDGEMENT
Part of the full-scale computations for the Series 60
case were carried out under the EC-funded EFFORT
project, 'European Full-scale Flow Research and
Technology', G3RD-CT-2002-00810. The financial
support is gratefully acknowledged.
The Dyne tanker wave data were kindly provided by
Dr. C.-E. Janson of Chalmers University Gothenburg.

REFERENCES
Dacles-Mariani, J., Zilliac, G.G., Chow, J.S.
and Bradshaw, P., "Numerical/experimental study of a
wingtip vortex in the near field", AIAA Journal,
Vol.33, Sept. 1995, pp. 1561 - 1568.
Ea, L. and Hoekstra, M., "Numerical
prediction of scale effects in ship stern flows with
eddy-viscosity turbulence models", 23rd Symp. Naval
Hydrodynamics, Val de Rueil, France, Sept. 2000.
Hoekstra, M., "Numerical simulation of ship
stern flows with a space-marching Navier Stokes
method", Thesis, Technical University of Delft,
October 1999.
Larsson, L., Patel, V.C. and Dyne, G. (Eds.),
"SSPA-CTH-IIHR workshop on viscous flow",
Flowtech Research Report 2, Flowtech Int. AB,
Gothenburg, Sweden, 1991.
Larsson, L., Regnstrm, B., Broberg, L., Li,
D.-Q. and Janson, C.-E., "Failures, fantasies and feats
in the theoretical/numerical prediction of ship
performance", 22nd Symp. Naval Hydrodynamics,
Washington D.C., U.S.A., 1998.
Lewis, M.R., "Numerical methods for water
flows with free-surface gravity waves", Thesis, Delft
University of Technology, June 2004.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Lundgren, H. and hman, M., "Experimentell


och numerisk bestmning av vgmotstand fr ett
tankfartyg (Dynetankern)", Report No. X-94/58, Dept.
Naval Architecture and Ocean Engineering, Chalmers
Univ. Techn., Gothenburg, Sweden, 1994. (In
swedish).
Menter, F.R., "Eddy-viscosity transport
equations and their relation to the k- model", Journal
of Fluids Engineering,Vol. 119, pp. 876-884, 1997.
Raven, H.C., "A solution method for the
nonlinear ship wave resistance problem", Thesis,
Technical University of Delft, June, 1996.
Raven, H.C., "Inviscid calculations of ship
wave making --- capabilities, limitations and
prospects", 22nd Symp. Naval Hydrodynamics,
Washington D.C., U.S.A., August, 1998.
Raven, H.C. and Van Brummelen, E.H., "A
new approach to computing steady free-surface viscous
flow problems", 1st MARNET-CFD workshop,
Barcelona, Spain, 1999.
Raven, H.C. and Starke, A.R., "Efficient
methods to compute ship viscous flow with free
surface", 24th Symp. Naval Hydrodynamics, Fukuoka,
Japan, 2002.
Saad, Y. and Schultz, M.H., "A generalized
minimal residual algorithm for solving nonsymmetric
linear systems", SIAM Jnl. Sci. Statist. Comput., Vol.7,
pp. 856-869, 1986.
Schweighofer, J., "Viscous-flow computations
at full-scale ship Reynolds numbers using the RANS
solver FINFLO", 6th Numerical Towing Tank
Symposium, Rome, Sept. 2003.
Starke, A.R., The prediction of scale effects
on ship wave systems using a steady iterative RANS
method, NUTTS 2004 Symposium, Hamburg, 2004.
Toda, Y., Stern, F. and Longo, J., "Mean-flow
measurements in the boundary layer and wake field of
a Series 60 Cb=.6 ship model for Froude numbers .16
and .316," IIHR Report No. 352, Iowa Institute of
Hydraulic Research, August, 1991.
Van Brummelen, E.H., "Numerical solution of
steady free-surface Navier-Stokes flow", Report MASR0018, Institute for Mathematics and Computer
Science, Amsterdam, June 2000.
Van Brummelen, E.H., Raven, H.C. and
Koren, B., "Efficient numerical solution of steady freesurface Navier-Stokes flow", Jnl. Computational
Physics, Vol. 174, 2001, pp. 120-137.
Van der Ploeg, A., Ea, L. and Hoekstra, M.,
"Combining accuracy and efficiency with robustness in
ship stern flow calculation", 23rd Symp. Naval
Hydrodynamics, Val de Rueil, France, Sept. 2000.
Verkuyl, J.-B. and Raven, H.C., "Joint
EFFORT for validation of Full-Scale Viscous-Flow
predictions", The Naval Architect, Jan. 2003.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Michel Visonneau
Ecole Centrale Nantes/CNRS, France
This paper describes a new steady coupled
free-surface fitting algorithm to compute viscous
flow with free-surface. Its efficiency in terms of
speed of convergence is remarkable and convincing
illustrations of its accuracy are provided, for instance,
with the simulation of the flow around the Dyne
tanker for a low Froude number (Fr=0.165).
My question concerns the formulation of the
equation resulting from the combination of the steady
kinematic equation and the normal dynamic
condition. I guess that the term w z included in
Eq. 3 is due to a typographic error. However, I do not
understand why V3 , 3 is still present in Eq. 4 since
V3 should vanish on the steady free-surface which is
a zero mass-flux surface at convergence. Could you
comment on that and explain also the role played by
the artificial fourth boundary condition needed to
close the system?
AUTHORS REPLY
The term w.z is required, as is seen as follows.
The dynamic condition relates the wave elevation, a
function of 2 variables, with the pressure field, a
function of 3 variables:
(x,y) = Fn2 (x,y, (x,y))
Thus the derivatives of in the kinematic condition
require an additional contribution in / z, and
using the kinematic condition again this produces the
term w.z , as follows:
u / x + v / y =
= Fn2 .[(u (/x + / z . / x ) + v ( /
y + / z. / y ) ] =
= Fn2 (u / x + v / y + w / z)
The term V3 . / 3 as a matter of fact vanishes
upon convergence, but is nonzero during the iterative
solution process; thus including it or not can affect
the convergence rate. In practice, it is now dropped
for robustness reasons.
The 4th boundary condition is not a real
boundary condition, but a condition only required to
make the set of equations complete. This equation
has to be sufficiently weak. E.g. a condition
d/dn=0, as seems to be used in some methods,
gives wrong results. Therefore, we use the
momentum equation normal to the free surface, with
some small simplifications only.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics


St. Johns, Newfoundland and Labrador, CANADA, 8-13 August 2004

The use of detached-eddy simulation in ship hydrodynamics


R.J. Pattenden, S.R. Turnock and N.W. Bressloff
(University of Southampton, U.K.)
Abstract
Accurate numerical simulation of the flow around the
hull of a ship remains a challenging task. This work
examines whether this is as a result of the inherent
unsteady turbulent flow regime influencing the
position and strength of vortex systems shed in the
wake. To this end the behaviour of the flow around
a truncated cylinder is taken as an exemplar of flow
at the stern of a ship. Wind tunnel tests of the flow
around a KVLCC2 hull form were made to measure
the flow field at the propeller plane using particle
image velocimetry. Comparison of the instantaneous
and the mean flow field indicate significant low
frequency effects with large deviations from the
mean flow. Such behaviour is also found in the wake
of an aspect ratio 1 truncated cylinder.
The
performance of a detached eddy simulation (DES)
approach at capturing the on and off body flow field
is compared with detailed experiments and previous
calculations of a Reynolds Averaged Navier Stokes
with k- model and Large Eddy Simulation (LES).
The results clearly show significant improvements
with the use of DES. Further work is on-going to
apply the DES approach to a full ship simulation.

1.

Introduction

The accurate numerical solution of the flow around


the aft end of ship hulls is of interest to naval
architects and ship builders. The majority of
techniques are currently based on the solution of the
Reynolds-Averaged
Navier-Stokes
(RANS)
equations. Closure is provided through the use of a
variety of turbulence models of varying degrees of
applicability. The solution of steady flow problems,
for example when predicting full scale ship
resistance, relies on the turbulence model to capture
the mean flow behaviour due to unsteady fluctuations
in the boundary layer and wake. The RANS approach
can give reasonably accurate predictions of the drag
and of the mean flow velocities in the propeller plane,
but the flow here is a complex shear flow with strong
longitudinal vortices created by the curvature of the
hull and so it represents a considerable challenge to

RANS solvers. This was demonstrated at the


Gothenburg 2000 workshop on numerical ship
hydrodynamics (Larsson et al., 2000) using the
KVLCC2 tanker hullform, developed at the Korean
Research Institute of Ship and Ocean Engineering
(Van et al., 1998).
Detached-eddy simulation (DES) is a hybrid
between Reynolds-averaged Navier-Stokes methods
(RANS) and large-eddy simulation (LES), developed
by Spalart et al. (1997). It is of interest for high
Reynolds number ship flows as it does not require as
fine a mesh at the wall as a pure LES approach
would, while still working like an LES model away
from the wall (ship hull) where the large-scale
structures form and convect.
Previous experimental and computational
work (Pattenden et al., 2002; Pattenden et al., 2003)
has been carried out on a cylinder of diameter/height
aspect-ratio 1 mounted on a ground plane. This
geometry was used as an exemplar on which to
investigate the use of different numerical approaches
for three-dimensional separated flows. The flow
contains a number of features that are similar to flows
around ship sterns, including a large area of attached
turbulent boundary layer flow and streamwise
vortices.
The objectives of the work presented are
threefold:
1. to carry out a wind tunnel study of the flow
around a 1:320 scaled model of the KVLCC2
hull form and make instantaneous particle image
velocimetry (PIV) measurements of the flow
field at the propeller plane;
2. to implement a validated DES method and use it
to solve the flow around the truncated cylinder
and compare with previous simulations using
RANS and LES methods.
3. to prepare for a full DES calculation of the flow
around the model scale KVLCC2 hull form.
The paper presents the details of the numerical
simulation tool and the implementation of LES and
DES used. A description is given of the KVLCC2
wind tunnel tests and results. Results are presented
for the DES simulation and comparisons made with
experiment and the RANS/LES calculations. Finally

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

the way forward for full DES calculations of ship


forms is proposed.

2.

Numerical simulation methods

The calculations presented were carried out using a


modified version of Elmore (Bressloff, 2001). This
uses a finite volume discretisation based on
structured multi-block grids using collocated
variables. Convective fluxes use the second order,
curved line advection method (CLAM) of Van Leer
(1974). Diffusive terms use second order central
differences. All simulations, including steady RANS
are solved as time dependent with a second-order,
implicit three time level discretisation.
Appendix 1
gives
details
of
the
implementation of the Spalart-Allmaras turbulence
model and the LES.
The detached-eddy simulation method was
originally proposed by Spalart et al. (1997) as a
means of overcoming the considerable computational
cost of LES of high Reynolds number flows around
real geometries, which is likely to remain prohibitive
for many years. It is based on the assumption that
RANS models are capable of modelling attached
boundary layer flows to a satisfactory degree, so that
an LES treatment is only required in the regions of
separated flow. As most flows contain large areas of
attached flow the computational cost is immediately
reduced by using a RANS method in these regions.
The DES method is one of a number of
hybrid RANS/LES techniques but is gaining in
popularity due to its simplicity. It consists of a
modified RANS model which operates as normal
close to the walls but switches its behaviour to that of
an LES subgrid model away from the walls. This is
achieved through the modification of the length scale
in the turbulence model, which is related to the
distance to the nearest wall for RANS modelling, but
changes to be related to the grid spacing away from
the wall.
The original DES method was based on the
one equation Spalart-Allmaras turbulence model
(Spalart and Allmaras, 1994) which uses the distance
to the nearest wall as the length scale, lSA. For DES
this is modified so that it depends on the grid spacing,
, as follows,

~
l SA = min(l SA , C DES ) ,
where is the largest dimension of the cell, not the
cube root of their sum,
= max(x, y, z ) .

CDES is a constant which defines the


transition point between the DES and RANS modes.
In isotropic turbulence it has been calibrated to
CDES=0.65, which has also been found to be valid for
most other flows.
The length scale is given as that to the
nearest wall boundary. This needs to be calculated
once at the start of the calculation on a particular grid
as for a multi-block mesh the length scale search
process is time consuming in its own right.
The output from an unsteady flow solution
at any instant gives a snapshot. However, it is
necessary to calculate statistical quantities, which can
be achieved by periodically saving the solution and
averaging. Unfortunately, for the typically large
meshes required for LES or DES calculations the data
storage required can become very large.
For
example, storage of the values of one variable at one
time step on a 2.6 million cell mesh requires 12Mb of
storage. To minimise the storage requirements a
running average approach was used. This was
achieved by the addition of statistical variables into
the code, which does require more memory but
avoids the need to write data to the disk and uses all
the available data.
The statistical quantities required for a three
dimensional flow are the means of the three velocity
components and the components of the Reynolds
stress tensor uiuj. These are defined as follows, for N
discrete samples, where uv represents one of the
Reynolds stress components.

U =

1
N

1
N

(U U )(V V )

uv =

N
0

It follows that at a time step n, U and


by,
Un =

uv n =

uv are given

U n 1 (n 1) + U n
n

)(

uv n 1 (n 1) + U n U n V n V n
n

It is therefore necessary to store nine additional


variables, U , V , W , uu , vv , ww , uv , uw , vw .
These variables increase the memory requirement by
4%. The values are written to disk along with the all

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

20
16
14
12
u+

10
8
6
4
2
0

-1

y+

(a) Velocity profile normalised by wall flow


1
Kim DNS
128x8
128x16
128x32
128x64
128x128

0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0

tanh ((1 )q )
y
,
= p + (1 p )1

L
tanh (q )

0.2

0.4

0.6

0.8

1.2

1.4

(b) Velocity profile normalised by outer


flow

where,

1
128x8
128x16
128x32
128x64
128x128

0.9

n
N

0.8
0.7

Here n is the current node number and N is the total


number of nodes on the edge. L is the edge length.
In this case p=0.064668 and q=1.479993 giving a
first cell size of y/=0.0056 on the finest mesh. This
corresponds to y+1. For the purposes of this study
the distribution was kept constant so that the wall
spacing increased with decreasing grid resolution.
The solver was run for 1.5 times the flow-through
time.
Figure 1 shows the profiles of velocity and
turbulent viscosity, t, at x/=100. The flow at this
point is fully developed and it can be seen that the
agreement with the DNS data of Kim et al. is very
good except for the two very coarse grids. The wall
cell spacing in the 1288 and 12816 grids was

0.6
y/

Kim DNS
128x8
128x16
u +=y +
128x32
128x64
u +=1/0.4*ln y ++5.5
128x128

18

y/

the other variables for post-processing or to restart


the solution.
The largest simulations were run on a
minimum of 6 dual-processor nodes (1 GHz Pentium
III with 512 Mb of memory). The nodes are
connected by 100Mbps Fast Ethernet connections.
Typically the truncated cylinder simulations required
approximately 1 minute per time-step. This requires
one week for 10,000 time-steps. Up to 50,000 timesteps were necessary to obtain converged statistics.
The DNS channel flow solution of Kim et
al. (1987) was used to demonstrate the performance
of the Spalart-Allmaras (S-A) model. In their
simulations the Reynolds number based on the mean
centreline velocity and the channel half-width, which
is equal to the boundary layer thickness, , was 3300
(Re=180). The grid size was 192129160 in x, y
and z, giving around 4106 cells. The dimensions of
the domain were 422 with periodic
boundaries in the streamwise and spanwise
directions. This size was shown to be sufficient
based on two-point correlations.
A two-dimensional grid was used with a
length in the x-direction of 100 to allow the
boundary layer to fully develop. The number of cells
in the streamwise direction was 128. In the ydirection the number of cells was varied from 8 to
128 using a non-uniform spacing according to the
function (Eisemann, 1979),

0.5
0.4
0.3
0.2
0.1
0

(c) Profile of t
Figure 1: Profiles of velocity t and for
channel flow with different grid resolutions

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

0.2
0.18
0.16

0.18
0.16
0.14

0.12

0.12

0.1

0.1

z/d

z/d

0.14

0.2
SA
DES
Exp

0.08

0.08

0.06

0.06

0.04

0.04

0.02

0.02

0
0.5

0.6

0.7

0.8
U/U

0.9

0
0.5

DES x/d=-0.5
DES x/d=0
DES x/d=0.5
Exp

0.6

0.7

(a) S-A and DES models compared to experiment at


x/d=-0.5

0.8
U/U

0.9

(b) Development of boundary layer using DES through


the refined mesh

Figure 2: Results of S-A and DES modelling of turbulent boundary layer


outside the laminar sub-layer and so accuracy would
be impaired despite the use of a wall-function.
One of the features of typical DES
simulations is that there are regions of boundary layer
flows where the along which the grid is refined so
that the model will switch from RANS to DES mode.
In the case of the truncated cylinder this occurs on the
ground plane approaching the cylinder, while on a
ship hull the mesh will be refined towards the stern.
It is therefore useful to assess the effect of this
transition on the prediction of the boundary layer. To
do this a simple grid was constructed with the same
length and height as the truncated cylinder domain,
that is 10.67d long by 4d high. The grid was refined
in the x-direction at the location of the cylinder. For
the RANS cases a two-dimensional grid was used,
while for the DES model a three-dimensional grid
was employed with a z-dimension of 0.05d. The grid
used is not fine enough to resolve the boundary layer
structures properly in LES mode but there is still a
region where the grid is refined around the base of
the cylinder where the DES mode will be activated.
The finest grid had a wall spacing, y+=7, and x+
ranging from 18 to 1080.
The performance of the Spalart-Allmaras
RANS and DES models is compared in Figure 2(a).
It can be seen that the SA and DES results are very
close as would be expected since the DES should be
functioning in RANS mode upstream of this point.
Figure 2(b) plots the boundary layer profiles of the
DES model at different streamwise locations. The
first curve at x/d=-0.5 is essentially the RANS
solution. Beyond this point the grid is refined so that
the model operates in LES mode. Here the turbulent

viscosity is reduced due to the reduced length scale.


If the turbulent fluctuations were simulated correctly
in this zone the velocity profile should remain
correct. However in this case the profile can be seen
to straighten slightly.
This suggests that the
fluctuations are not being simulated fully. This is
likely to be due to the lack of turbulent fluctuations
entering this region from the RANS region upstream.
This is an acknowledged weakness of DES and
applies equally to areas of separation where the
separated shear layer does not contain any turbulence
from the boundary layer (Travin et al., 2000).

3.

Wind Tunnel Experiments

The experiments for both the truncated cylinder and


KVLCC2 model were carried out in an open circuit
suction wind tunnel with a 0.9m wide by 0.6m high
and 2.4m long working section. A series of three
meshes are used to control turbulence which is
measured as 0.3% at the inlet.
Full details of the arrangements for the
cylinder tests can be found in Pattenden (2004). The
computer simulations were set to reproduce the
geometry from the leading edge of the ground plate
and include the walls in the same positions. The
cylinder has a height and diameter of 150mm. At a
speed of 20m/s this corresponds to a Reynolds No.
based on diameter of 1.96105.
Measurements were made of time varying
cylinder surface pressures, surface-flow visualisation
using oil evaporation, hot wire anemometry, particle
image velocimetry (PIV), and cylinder force
measurements.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

-0.01

-0.02

z/d

-0.03

-0.04

-0.05

-0.06
0

0.01

0.02

0.03

0.04

0.05

y/d

(a) Mean flow

(b) Instantaneous flow

Figure 4: Velocity vectors and vorticity contours of the flow in the propeller plane of the KVLCC2 model

25
DWL

z (m)

20
15
10
5
0

-25 -20 -15 -10 -5

0 5 10 15 20 25
y (m )

Figure 3: Body plan of KVLCC2 hull


So far only PIV measurements have been
made for the KVLCC2 hull model whose body plan
is given in Figure 3. A 1/320 scale model (1m long)
was commissioned, made of hardwood (Jelutong).
The model is divided longitudinally into three
sections so that the parallel mid-body can be
removed, and also allows for an extra piece to be
added above the waterline to increase the depth of the
model. The sections are connected with dowel pins.
The model was mounted on a ground plate in the
wind tunnel in the same way as the cylinder
described above. While this is not a strictly accurate
representation of the real ship due to the presence of a
boundary layer on the ground plate, the CFD model
can be set to model this as is the case with the
truncated cylinder.

It may also be the case that a double-body


model as used in the KRISO tests does not give the
same turbulence behaviour as the real ship. This is
because the eddies are free to cross the whole wake
rather than being constrained by the free-surface.
Particle image velocimetry (PIV) was used
to obtain data on the flow in the propeller plane. The
laser sheet was oriented in the y-z plane so that
vectors of V and W were measured. The data was
averaged over 500 samples. Only half of the plane
was measured to enable greater resolution. The tests
were run at a Reynolds number based on hull length
of 1.3106. The boundary layer on the surface of the
hull was tripped using a roughness strip so that the
flow is fully turbulent.
Results are presented for the full hull
mounted with a depth equal to DWL. Figure 4 shows
plots of velocity vectors with contours of vorticity,
normalised by the maximum vorticity magnitude.
In the mean flow there are the longitudinal
bilge vortices close to the hub of the propeller as
found by Van et al. (1998). However, as in the
truncated cylinder flow these mean vortices appear
very unsteady in nature revealing a chaotic
instantaneous flow with many large-scale coherent
structures. Only in the mean flow do these appear as
the two counter-rotating vortices. This is similar to
the flow found by Bearman (1997), for the flow

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Table 2: Details of truncated cylinder DES results


Re

100000

200000

80
83
-1.17
-0.92
0.47
2.2
0.71

80
84
-1.11
-0.98
0.41
2.1
0.68

CPU time / time step (secs)


Side separation, s (degrees)
Primary sep. from ground, XS1 (x/d)
Secondary sep. from ground, XS2 (x/d)
Attachment on free-end, XRT (x/d)
Attachment on ground downstream, XRF (x/d)
Local CD at z/d=0.5

Table 1: Details of DES grid

70
-1.00
-0.78
0.17
1.6
0.79

1
Exp
DES

0.8
0.6
0.4
0.2
P

324
72
52
0.0098
0.002
0.002
2,760,960

Nc
Nr
Nh
Sc
Sr
Sh
Total cells

200000
Exp.

0
-0.2
-0.4
-0.6
-0.8
-1

20

40

60

80 100 120 140 160 180

(degrees)

Figure 6: Pressure distribution around cylinder


at z/d=0.5

Figure 5: Structure of blocks used to create


truncated cylinder mesh
around a car. It is questionable whether the mean
flow produced by a RANS calculation or even
measured by time-averaging techniques is a valid
picture of the flow for many applications. For
example the flow over the propeller blades could be
more influenced by this short time-scale flow pattern
than by the mean flow. While this may not matter for
the mean thrust characteristics it may be a factor in
the simulation of vibrations and pressure fluctuations.

4. Application of DES to a truncated


cylinder
Table 1 details the mesh used to solve the DES flow
around the truncated cylinder. There were a few
minor modifications necessary, (Pattenden, 2004)
when compared to the RANS and LES meshes to
ensure that the DES solution remained stable. Figure

5 shows the block structure used. Results are


presented for one mesh run at two Reynolds number.
The key parameters characterising the flow
field for the two Reynolds numbers are listed in
Table 2. The separation from the side of the cylinder
occurs at 83 or 84 degrees which is similar to the
LES result of 81 degrees. This is also close to the 80
degrees at which the flow separates when the
boundary layer is tripped. As the flow in the DES
simulation is assumed to be turbulent everywhere it is
this value with which the results should be compared.
The extent of the primary horseshoe vortex,
defined by the separation point, XS2, on the ground
plane upstream of the cylinder, is at x/d=-1.0 which is
further forward than the measured position. This
could be due to it being in the ``grey area'' between
the RANS and DES modes where it has been seen
that the boundary layer resolution suffers. It could
also be due to weaknesses in the underlying SpalartAllmaras turbulence model. At the lower Reynolds
number the separation point is further forward.
The length of the recirculation region in the
wake is characterised by the attachment point on the
ground downstream of the body. This is predicted to
be x/d=2.1 which is exactly the same as the LES and

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

(a) Experiment

(b) DES

-0.5

-0.5

-0.4

-0.4

-0.3

-0.3

-0.2

-0.2

-0.1

-0.1

y/d

y/d

Figure 7: Surface flow visualisation on ground plane around cylinder

0.1

0.1

0.2

0.2

0.3

0.3

0.4

0.4

0.5
-0.5

0
x/d

0.5

(a) Experiment

0.5
-0.5

0
x/d

0.5

(b) DES

Figure 8: Surface flow visualisation on free-end of cylinder


k- results. This point is slightly further downstream
at the lower Reynolds number.
Figure 6 shows the pressure distribution
around the cylinder at z/d=0.5. The agreement with
the experiment is quite good with the slightly later
separation and higher base pressure.
One of the motivations for trying the DES
technique on this flow was that LES was unable to
resolve the boundary layer on the ground plane,
which seemed to cause the horseshoe vortex to be
incorrectly simulated. The DES method, acting as a
RANS model on the boundary layer could be

expected to perform as well as the k- model at


capturing the shape of the primary horseshoe vortex.
In practice however, the DES simulation
falls halfway between the LES and k- results. As
shown by the primary separation position reported
above, the horseshoe vortex extends further forward
than the one found in the experiment, but not as far as
the LES one. In fact the simulations do show a two
vortex system as is expected with the primary
separation close to the experimental position. The
main vortex however is too far upstream with XS2=0.98 compared to XS2=0.78. The vortex centre is

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

1.5
1.4

U/U

1.45

Exp
DES

-0.5

1.25

1.2

0.5

U/U

1.15
1.1

z/d=0.50

0
-0.5

1.05
0.5
U/U

1.5

Figure 9: Plot of U against z/d at x/d=0 on tip


of cylinder
further forward, and the stagnation point on the
cylinder wall appears to be lower than in the PIV
measurements. These results are an improvement
over both the LES and the k- model in terms of the
secondary vortices but not as good as the RANS at
predicting the primary vortex size and shape. This
again may be due to the grey area between the RANS
and LES modes of the DES model.
There is also uncertainty over the upstream
boundary layer, due to the lack of experimental data
with the model in position. Inaccuracies in the
approaching boundary layer would affect the
formation of the horseshoe vortex.
Additional
experimental data is therefore needed to be sure about
the behaviour of the simulations in this area.
The surface flow patterns in Figure 7
clearly show the separation line at the leading edge of
the horseshoe vortex. The flow pattern as the vortex
trails downstream shows similar features to the
experimental image in terms of the flow directions
under the vortex. The differences are caused by the
excessive length of the vortex. The separation line on
the inside of the horseshoe vortex, marking the edge
of the shear layer, diverges at an angle of 17.3
degrees compared to 15.5 degrees in the experimental
flow.
The two main aspects which characterise the
flow above the free-end of the cylinder are the
vertical and streamwise extents of the recirculation
bubble and the swirl patterns on the surface of the tip.
Figure 8(a) shows the experimental oil flow. The k-
RANS model completely fails to resolve the detail of
the swirls on the tip while the LES reproduces these
patterns well. This suggests that the simple no-slip
boundary condition with no turbulence model is the
best for this swirling boundary layer flow. It might
be expected that the DES with its RANS component

U/U

1
-0.5

z/d=0.75

1.3

z/d

1.35

0.5

0.5
z/d=0.25

0
-0.5
0.5

1.5

2
x/d

2.5

3.5

Figure 10: Plot of U against x/d along the


centreline in the wake of the cylinder
will damp out some of this motion. The extents of
the recirculation bubble on the centreplane are overpredicted by all models so far.
Looking at the surface flow patterns on the
free-end, shown in Figure 8(b), as expected the swirl
patterns are not completely captured, although the
direction of the streamlines is correct. It is just the
wrapping up of the vortex eye which is not resolved.
The prediction of this flow feature is likely to depend
strongly on the ability of the model to predict the
boundary layer flow on the top surface of the
cylinder. Due to the low speeds and strong pressure
gradients in this region the nature of this boundary
layer may be difficult to predict using RANS type
models.
The vortex on the centre-plane above the
cylinder extends too high above the cylinder in the
same way as the LES and RANS simulations.
Attachment occurs at x/d=0.41 which is further back
than the experiment. Figure 9, which plots U against
z/d at x/d=0, confirms that the peak velocity occurs
60% higher than in the experiment. The flow near
the surface is faster with a thinner boundary layer.
This is where the LES simulation performed well
suggesting that the boundary layer prediction with the
RANS model is incorrect for this area of the flow.
Figure 10 shows profiles of the U velocity
component in the streamwise direction, in the wake
of the cylinder. The numerical results can be seen to
be close to the experimental values although at the
bottom, at z/d=0.25, the length of the reversed flow is
too long, corresponding to the later attachment on the
ground.

Copyright National Academy of Sciences. All rights reserved.

0.8

0.8

0.6

0.6

z/d

z/d

Twenty-Fifth Symposium on Naval Hydrodynamics

0.4

0.4

0.2

0.2

0
-2

0
U/U

2 -0.5

0
W/U

0.5

0.05

0.1

u u

0.05

0.1

v v

0.05

0.1

w w

0.8

0.8

0.6

0.6

z/d

z/d

(a) x/d=0.83

0.4

0.4

0.2

0.2

0.5
U/U

1 -0.4

-0.2
W/U

0.02
u u

0.04 0

0.05
v v

0.1

0.05

0.1

w w

(b) x/d=2.5
Figure 11: Profiles of velocity and normal stress components in the wake of the cylinder (+ Experiment, DES)
Figure 11 shows profiles of velocity
components and Reynolds stresses with respect to z/d
at various locations in the x/d direction. The profiles
of U velocity are quite close to the experimental
values except in the shear layer near the cylinder
where the peak rise to the free-stream velocity is
slower. The later attachment means that the flow
near the ground is reversed for longer and is also
stronger due to the flatter vortex. The W profiles
follow the shape of the experimental curves but the
maximum downwash is less due to the shallower
angle of the shear layer. The agreement in the lower
half of the profiles is quite good. Looking at the
normal stresses, u'u' is over-predicted in the shear
layer close to the cylinder, as is w'w'. v'v' is under-

predicted almost everywhere, until after attachment.


The difference in the vertical location of the shear
layer is evident from the profiles of u'u', which show
the peak in the experimental values about 0.2d lower
in the experiment than in the numerical results at
x/d=0.83. Although not shown here, the u'w' shear
stress agrees quite well with the experimental data
with the peaks in the shear layer being very close. In
general these profiles show better agreement to the
experiment than the LES simulations.
The vector plots in Figure 12 show
reasonable agreement particularly downstream where
the large streamwise vortices have formed. The tip
vortices off the free-end are captured. In the same

Copyright National Academy of Sciences. All rights reserved.

0.8

0.8

0.6

0.6

z/d

z/d

Twenty-Fifth Symposium on Naval Hydrodynamics

0.4

0.4

0.2

0.2

-0.6

-0.4 -0.2

0
y/d

0.2

0.4

0.6

-0.6

0.8

0.8

0.6

0.6

0.4

0.4

0.2

0.2

-0.6

-0.4 -0.2

0
y/d

0.2

0
y/d

0.2

0.4

0.6

0.4

0.6

(b) x/d=0.83 (DES)

z/d

z/d

(a) x/d=0.83 (Exp)

-0.4 -0.2

0.4

0.6

(c) x/d=2.50 (Exp)

-0.6

-0.4 -0.2

0
y/d

0.2

(d) x/d=2.50 (DES)

Figure 12: Vector plots of the flow in transverse planes in the wake of the cylinder
way as the previous simulations, the differences can
be attributed to the longer recirculation region.

5.

Why use DES for ship flows?

While many of the features found in the truncated


cylinder flow have similarities with the flows found
around the sterns of ships, there are certain
differences. In particular the Reynolds number
around large tankers is typically 109, four orders of
magnitude greater than that considered here. There is
a long length of attached flow along the hull with a
turbulent boundary layer. This boundary layer
interacts with the bilge vortex towards the stern to
produce a region of unsteady turbulent flow. The
turbulent kinetic energy in the propeller plane is
lower than that found behind the truncated cylinder
k = 14%U , compared to k = 44%U but is
still large. The other difference is that the transverse
normal stresses, vv and ww are around half of the

streamwise stress uu , whereas in the cylinder wake

the streamwise stress is half the transverse stresses.


However the flow is still anisotropic and therefore
difficult for eddy-viscosity models to predict.
The larger scale structures in the flow in this
region are likely to have a characteristic frequency of
order 0.01Hz assuming that the non-dimensional
frequency is of order 0.1 and that it scales with
breadth and ship speed. This corresponds to a period
of 100 seconds so the propeller could experience
variations in inflow velocity lasting for long periods.
It would therefore be useful to be able to simulate this
flow as experiments are difficult to achieve at
Reynolds numbers anywhere near the full-scale
values.
Both the LES and DES results on the
truncated cylinder geometry showed that while the
agreement between the mean velocity and Reynolds
stress profiles was not so good in the recirculating
flow close to the cylinder, the results were quite good
further downstream, where the flow is dominated by
streamwise vortices. It is this part of the flow that
probably has most in common with the flow around a
ship's stern where the flow undergoes three-

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

dimensional separation to create streamwise vortices


but the dominant flow is still in the streamwise
direction.
The main question to be resolved is whether
the DES model will behave correctly as the grid is
refined towards the stern. It has been seen that there
is a grey area in the boundary layer modelling around
this interface which can result in an under-resolved
boundary layer and high levels of turbulent viscosity
being carried into the DES region. DES has been
used successfully on a prolate spheroid geometry
(Constantinescu et al., 2002) which also has much in
common with ship hulls.
As an intermediate step before simulating a
full ship geometry, the truncated cylinder grid was
stretched in the x-direction to give an elliptical shape
with the same length/breadth ratio as the KVLCC2
tanker (Pattenden et al., 2003). The flow in this case
is more like the ship flow in that there is no massive
separation at the back of the body, but due to the
sharp corner at the tip, a pair of longitudinal vortices
is formed, equivalent to the bilge vortices. In the
mean flow results this pair of longitudinal vortices
was clearly visible, but the instantaneous flow is
composed of a number of large-scale vortex
structures indicating that the longitudinal vortices are
unsteady. This again highlights the fact the these
longitudinal vortex structures are in fact made up of
many smaller chaotic vortices and indicates the need
to examine such structures in ship flows as well.

6.

Conclusions

DES simulations have been performed which show


good agreement with the experimental data in
predicting flow around a cylinder of aspect ratio 1.
In particular the streamwise vortices trailing
downstream beyond the region of massive separation
are well predicted with the Reynolds stresses
agreeing closely with the experimental data.
The quality of the mesh was found to be very
important to even get the code to run. In particular
attention has to be paid to the cell expansion ratio
which should not exceed 1.2.
While the boundary layer upstream was
modelled correctly the primary separation from the
ground was 22% too far upstream while the
secondary separation was 71% too far upstream,
resulting in the horseshoe vortex being stretched.
This could be due to the switch from RANS to DES
mode in the vicinity of the cylinder due to the refined
grid. It has been shown that this area poses a
problem as the turbulent viscosity due to the
RANS/SGS model dies away before the turbulent
fluctuations start to develop. Despite this problem

the horseshoe vortex is still better than the LES


model.
The regions of separated flow, where the DES is
functioning as an LES subgrid model, are very close
to the LES results. In the wake region the DES
appears to give better agreement with the velocity
profiles, which could be due to the improved grid
used for these computations. The Reynolds stresses
though are perhaps slightly worse than the LES ones.
The success of the DES simulations on this geometry
has indicated that the DES method could be used to
simulate the flow around the stern region of fullbodied ships such as tankers and bulk-carriers. Here
it should give increased understanding of the nature
of the wake flow in the region of the propeller which
will be valuable in predicting the performance of the
propeller.
Wind tunnel experiments have been carried out
at a Reynolds number of 1.3106 on a 1m segmented
model of the KVLCC2 hullform. The segmented
model allows the effect on the stern flow of length
and depth to be investigated. PIV measurements have
been made of the flow in the stern region, particularly
around the propeller plane to investigate the structure
of the turbulence here. These show the unsteady
nature of the bilge vortex.
It is shown that the cylinder flow has certain
features in common with the flow in the propeller
plane of the KVLCC2 tanker hull, notably the
streamwise vortices, high turbulence intensities and
anisotropic stresses.
It is estimated that the energy containing vortices
in the wake will have time-scales of the order of 100s
which could have an effect on the performance of the
propeller and rudder. This is a motivation for the
study of the turbulent flow in this region.

7.

References

Bearman, P. W., "Near wake flows behind twoand three-dimensional bluff bodies," Journal of
Wind Engineering and Industrial Aerodynamics,
No. 69-71, 1997, pp. 33-54.
Bressloff, N. W., "A parallel pressure implicit
splitting of operators algorithm applied to flow
at all speeds," Int. Jnl of Numerical Methods in
Fluids, Vol. 36, 2001, pp. 497-518.
Constantinescu, G. S., Pasinato, H., Wang, Y.Q. and Squires, K. D., "Numerical investigation
of flow past a prolate spheroid," Aerospace
Sciences Meeting, Reno, Nevada, AIAA, 2002.
Eisemann, P. R., "A multi-surface method of coordinate generation," Journal of Computational
Physics, Vol. 33, 1979, pp. 118-150.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Kim, J., Moin, P. and Moser, R., "Turbulence


statistics in fully developed channel flow at low
Reynolds number," Journal of Fluid Mechanics,
Vol. 177, 1987, pp. 133-166.
Larsson, L., Stern, F. and Bertram, V.,
Gothenburg 2000: A Workshop on Numerical
Ship Hydrodynamics, Gothenburg, Sweden,
2000.
Mtais, O. and Lesieur, M., "Spectral large-eddy
simulations of isotropic and stably-stratified
turbulence," Journal of Fluid Mechanics, Vol.
239, 1992, pp. 157-94.
Pattenden, R. J., Turnock, S. R. and Bressloff,
N. W., "An experimental and computational
study of three-dimensional unsteady flow
features found behind a truncated cylinder," 24th
Symposium on Naval Hydrodynamics, Fukuoka,
Japan, 2002.
Pattenden, R. J., Turnock, S. R. and Bressloff,
N. W., "Developments in the use of large-eddy
simulation for ship hydrodynamics," CFD 2003:
Computational fluid dynamics technology in
ship hydrodynamics, London, U.K., RINA,
2003.
Pattenden, R. J., An investigation of the flow
around a truncated cylinder, PhD Thesis.
University of Southampton, Southampton, 2004.
Spalart, P. R. and Allmaras, S. R., "A oneequation turbulence model for aerodynamic
flows," La Recherche Arospatiale, Vol. 1,
1994, pp. 5-21.
Spalart, P. R., Jou, W.-H., Strelets, M. and
Allmaras, S. R., "Comments on the feasibility of
LES for wings, and on a hybrid RANS/LES
approach," Advances in DNS/LES, 1st AFOSR
Int. Conf. on DNS/LES, Columbus Oh.,
Greyden Press, 1997.
Travin, A., Shur, M., Strelets, M. and Spalart, P.,
"Detached-eddy simulations past a circular
cylinder," Flow, Turbulence and Combustion,
Vol. 63, 2000, pp. 293-313.
Van Leer, B., "Towards the ultimate
conservative difference scheme. II.
Monotonicity and conservation combined in a
second order scheme," Journal of Computational
Physics, Vol. 14, 1974, pp. 361-370.
Van, S. H., Kim, W. J., Yim, D. H., Kim, G. T.,
Lee, C. J. and Eom, J. Y., "Flow Measurement
Around a 300K VLCC Model," Proceedings of
the Annual Spring Meeting, SNAK, Ulsan,
1998, pp. 185-188.

8. Appendix
formulations
8.1

S-A

and

LES

Spalart-Allmaras model

The Spalart-Allmaras model (Spalart and Allmaras,


1994) is a one-equation turbulence model developed
specifically for external aerodynamic type flows. It is
based upon a transport equation for the turbulent
kinematic viscosity, t . It contains a destruction term
based on the distance to the nearest walls, and an
optional trip term to allow laminar to turbulent
transition at a specified location. The equations of the
model are as follows. A transport equation is solved
for ~ , given by,

where,

with S being the magnitude of vorticity and d the


distance to the nearest wall.

8.2

Large-eddy simulation

Large-eddy simulation consists of a time-dependent


computation in which the large scale eddies in a
turbulent flow are fully resolved, while the smaller
scale eddies are modelled. In order to define the
equations that are to be solved the flow field must be
filtered to remove the smaller scales which are to be

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

modelled. This operation is typically defined as


follows:

where G(x,x) describes the filter function. In the


present studies a box filter is used of width, , equal
to the cell size. Any eddies smaller than the filter
width are regarded as small eddies and are modelled.
The filtered Navier-Stokes equations are
written as:

The approximation,

is made and is called the subgrid scale Reynolds


stress.
In most LES methods this stress is
approximated by an eddy-viscosity model:

There have been many models proposed for the


calculation of t in this equation. The one used in the
simulations presented here is the structure-function
model as proposed by Mtais and Lesieur (1992).
Here the eddy viscosity is given by,

where,

CK is the Kolmogorov constant and x is the cell size.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Michel Visonneau
Ecole Centrale Nantes/CNRS, France
This paper presents an experimental and
computational study of the flow around a truncated
cylinder of aspect ratio 1 mounted on a ground plane.
This study illustrates the potentialities and
weaknesses of DES turbulence closure. Then, based
on a PIV visualisation of the flow at the stern of a
KVLCC2 model mounted on a ground plane, the
authors argue about the necessity of using DES to
compute ship flows.
I should say that I am fully convinced by the
arguments against the use of DES for ship flows
given by the authors in paragraph 5. A ship flow is
fundamentally a boundary layer flow which is
gradually evolving towards a thick boundary layer
giving rise to steady longitudinal vortices because of
the regular geometrical modifications of the hull. I
have therefore the feeling that the bad behaviour of
DES on pure boundary layer flow (illustrated by the
authors in their computations of the flow around the
truncated cylinder) does not predispose this closure
for the computation of turbulent flows over elongated
bodies. Moreover, when full scale free-surface flows
on realistic hulls with complicated appendages are
now reasonably well computed with steady RANSE
solvers using sophisticated turbulence closures
(EASM or RSM closures, for instance), the cost in
terms of CPU time of an unsteady DES approach will
be too high for industrial applications with limited
benefits. Could you comment on that and convince
me about the potentialities of DES for ship flows in a
near future?

realistic a model of the actual flow as possible. It is


also worth noting that as the flow becomes
progressively less symmetrical when the hull is
manoeuvring for instance that the capture of the offhull system (with some similarities to that found on
the top of the truncated cylinder) will be much better
captured with the LES approach. Again, the
difficulties remain with the transition between the
two flow regimes.

AUTHORS REPLY
M. Visonneau is understandably sceptical
about the application of DES to ship flows. The
authors concede that the weakness in the boundary
layer prediction in the region between the RANS and
LES modes shown in this paper is a major hurdle to
its application to ship flows, which are dominated by
boundary layer flows. However, if this problem can
be overcome there are potential benefits in the
amount in information that can be obtained on the
turbulence in the wake of the vessel. This could be
particularly useful for investigating the noise and
vibration induced on the hull and propeller due to
turbulence. From an industrial point of view the
current RANS methods are good enough to predict
the time-averaged flow around the wake and will
continue to be used for some time to come. From a
research point of view however it is important to
investigate all the options in order to obtain as

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics


St. Johns, Newfoundland and Labrabor, CANADA, 8-13, August 2004

Numerical Study on Turbulent Flow around Ship Models


Using a Large-Eddy Simulation Technique
H.-H. Chun, J.-C. Park, H.-J. Choi, H.-S. Yoon, D.-H. Kang
(Pusan National University, Korea)
Multiple Data Programs). This means that each
processor connected by high efficiency network
produces different data by running simultaneously
several single programs. In terms of parallelization
using domain decomposition, the MPI is employed to
exchange the data between neighbor domains, which
satisfies the continuity of interfaces between domains
and define the boundary condition within each domain.
For cross-checking the grid-dependency of ship
flow by LES, 1/5-, 1/2-, 1- and 2-million grid points
are used per each ship model at the Reynolds number
of a million. As the results of simulation, it seems that
the accuracy of LESs are closely related to the number
of grid points (or fine grid size), and the good
agreements with the experiments for the velocity,
pressure and wake distributions around ship models
can be obtained using the fine grid.

ABSTRACT
A Large-Eddy Simulation (LES) technique based on a
finite-volume method is applied to investigate the
turbulent flow around a Wigley and a Series
60(Cb=0.6) double models. Four types of turbulence
models in the present LES are tested through the
numerical tests for the turbulent channel flow at
Re = 180 . For the simulation of turbulent ship flow,
1/5-, 1/2-, 1- and 2-million grid points are used, and all
computations are performed on a 24-node PC-cluster
parallel machine composed of 2.6GHz CPU, which are
installed in the Advanced Ship Engineering Research
Center(ASERC, http://www.aserc.pusan.ac.kr) at Pusan
National University, Korea.
INTRODUCTION

LES - GOVERNING EQUATIONS AND


MATHEMATICAL FORMULATION

Numerical simulation techniques currently


available for turbulent flows include direct numerical
simulation (DNS), large-eddy simulation (LES), and
Reynolds-averaged Navier-Stokes simulation (RANS).
RANS, whose computational requirement is much less
demanding than DNS and LES, has been used
extensively for engineering calculations. In general,
RANS works reasonably well for attached turbulent
flows, since the empirical parameters in RANS models
are carefully tuned for such configurations. However,
for massively separated flows, such as that over an
airfoil with high attack angles, RANS models fail to
produce reliable and consistent results. LES has been
known as a promising tool in studying practical
turbulent flows at high Reynolds numbers. Thus, LES
approach is employed to simulate turbulent flow
around ships and whose results will be compared with
those of RANS obtained by previous other researchers.
Here, we have a parallelized LES code, which
allows us to perform larger-scale computations. The
methodology of parallelization using MPI(Message
Passing Interface) is the SPMD(Single Program

Governing equation
In LES, each flow variable f is decomposed into as
follows,
f = f + f

(1)

where, f is the resolvable grid scale(GS) component


and f is the residual sub-grid scale(SGS) component.
We denote the grid scale field as,
f ( x1 ,x2 ,x3 ) = G( x1 x1',x2 x2 ',x3 x3 ')
f ( x1',x2 ',x3 ')dx1' dx2 ' dx3 '
Here, Top-Hat filter is used as G,

Copyright National Academy of Sciences. All rights reserved.

(2)

Twenty-Fifth Symposium on Naval Hydrodynamics

G ( x1 , x2 , x3 ) = 1 / 3 , (3 = x1x2 x3 )

Discretization

(3)

The governing equations are solved in the


framework of the curvilinear coordinate system, where
the physical space (x1 , x2 , x3 ) is mapped to the
computational space (1 , 2 , 3 ) . The finite-volume
approach with the collocated primitive variable
arrangement is adopted for the discretization. The
overall procedure of discretization is similar to that of
Miyata et al.(1997).
The transformed governing equations are

After applying the filtering operation denoted by


the bar ( ) to the momentum equations and the
continuity equation, we get following equations for the
filtered field.

( ui u j ) p
ui

+
ij + 2 Sij
t
x j
xi x j

u i
=0
xi

(4)

(5)

d
udV = T dS
dt CV
CS

(8)

u dS = 0

(9)

where, S ij indicates the SG components of deformed


velocity tensor as follows;

CS

u u j

S ij = i +
x

x
j
i

T = uu PI +

(6)

(10)

By implementing the conservation laws with these


equations, the solution is achieved by the modified
MAC(Marker-And-Cell)-type
time-marching
algorithm.
For the differencing of the convective terms the
MUSCL-type third order upwind scheme with TVD
limiter is used, and for the time differencing the
2nd-order Adams-Bashforth method is used so that it
will suit the time-marching simulation procedure.

stress tensor in LES, and must be modeled in terms of


the resolved scale velocity, u i , in order to obtain a
closure for Eq. (4).
Spatial filtering is assumed for LES, ij can be
decomposed as

ij = (u i + ui )(u j + u j ) u i u j
= u i u j u i u j + u i u j + u iu j + u iu j
14243 14243 {
Cij

1
T
u + ( u )

Re

+ u u

and, ij is referred to as the sub-grid scale (SGS)

Lij

(7)

LES models

Rij

Smagorinsky model (S-model)


The Smagorinski model (1963) is most basic
model in LES, and the assumption for the eddy
viscosity, s , is derived as

where, the prime () denotes the fluctuating part of


the variable from the filtered value, and the terms, Rij ,

Lij and Cij , are respectively SGS Reynolds Stress


term, Leonard term and Cross term which must be
modeled in terms of grid scale components to precede
the calculation.
Several closure models for the subgrid scale stress
have been proposed. They fall into one of the following
general categories: eddy viscosity models, scale
similarity model, mixed models and Lagrangian based
models.
The filtered equations (4) and (5) are discretized in a
curvilinear coordinate system.

ij = 2 L2S S ij S ij
S ij =

1 u i u j
+
2 x j xi

(11)

(12)

where, S ij is the strain rate tensor defined as


2

S ij = 2S ij2

Copyright National Academy of Sciences. All rights reserved.

(13)

Twenty-Fifth Symposium on Naval Hydrodynamics

Here, is the filter width associated with the grid

and the length scale, LS , is set at the minimum value


among the grid size and the minimum spacing
following Takakura et al. (1991):
LS = min[C S , CT (x1 , x2 , x3 )]

(12)

=3 V

(13)

filter, and is the filter width associated with the test


filter. Also, < > indicates the ensemble averaging
operation, and the filter width ratio, , is set at 2.
The discretized filtering operation for the
finite-volume method is given by
u i = ui + 2

Here, the Smagorinskys coefficient, C S , is set at 0.1,


and the Takakuras coefficient, CT , set at 1.0 in this
study.
Near the walls, Smagorinsky model is not
appropriate because of the nonhomogeneous effect of
the wall turbulence. Therefore the universal model of
by Spalding(1961) is employed as wall damping
function written as,

(0.4u )
+

u =

e = C

(0.4u )
+

24

u
y+
and u + =
y Re
u

(14)

< L*ij M ij >


2

2 < M ij M ij >

u i
u j
+ ejl

xl
xl

(20)

~
~
3C B (u i u i )(u j u j )
S
~
~
(u u k )(u k u k )

(21)

ij = eil

(15)

eil = C s 2
2

Dynamic SGS model (D-model)


The D-model suggested by Germano et al.(1991) is
not a model but is better referred as a procedure to
evaluate the coefficient of SGS models. In the dynamic
procedure, the grid filter and the test filter are defined
to reduce the high frequency fluctuation and to obtain
smoothly the value of C S in the region of lower
frequency. The former corresponds to the
computational grid, and the later corresponds to a
coarser. While the grid filtered field is explicitly solved,
the test filtered field is used to determine the model
coefficients.
In the present study, the following model is used
according to the Lillys least square technique(1992).
C=

(19)

Generalized Normal Stress (GNS) model


The GNS model was proposed by Horiuti(1993)
as follows.

(0.4u )

(18)

where, = 3 V , and =2.0 is used in this study.


Finally, the turbulent eddy viscosity, e , is defined as

y + = u + + 0.1108 e 0.4 u 1 0.4u +

2 u i +1 2u i + u i 1
+ O(4 )
24
(x1 ) 2

where,

~
~
3C B (u i u i )(u j u j )
~
~
(u u k )(u k u k )

is for damping, and

C B =1.0.
Structure Function (SF) model
The SF model was suggested by Metais &
Lesiur (1992), and the turbulent eddy viscosity can be
derived from energy spectrum as follows.
u j

ij = e ij il

xl

u l
x j

(22)

e ij (x, x ) = 0.105C k x j [F2 il (x, x )] jl

(16)

(i = j )
(i j )

(23)

where,
M ij = S S ij S S
2

ij

with =

where, C k =1.4, and 2nd-order accuracy of SF model


F2 il (x, x ) is given as

(17)

{ [u (x ) u (x + x )]
+ [u (x ) u (x x )] }

F2 il (x, x ) =

1
6

j =1

Copyright National Academy of Sciences. All rights reserved.

(24)

Twenty-Fifth Symposium on Naval Hydrodynamics

Here,

is taken the sum of data neighboring 6

15

grids in each directions.

Ideal Speed
6
0.20*10
6
1.00*10

12

Speed-Up Ratio

MPI PROGRAMMING TECHNIQUE AND


PERFORMANCE TEST
In this research, the virtual topology method
(domain decomposition) as one of the MPI
programming strategies was employed. Figure 1 shows
the virtual topology as above mentioned. Each
processor has its own coordinates independently by
virtually disposing processors for parallel computation
as 2 or 3 dimensional arrangement. For example,
Processor 0 has (1,1) coordinate, Processor 1 has (1,2)
coordinate, Processor 2 has (2,2) coordinate and
Processor 3 has (2,1) coordinate as shown in figure 1.
As using this approach, the data transfer during parallel
computation is much more simply and the number of
processor taken part in parallel computation can be
handled flexibly.
Generally, the efficiency test of parallel program
uses the speed-up ratio. The speed-up ratio represents
the synchronization and the network effect on parallel
algorism. The performance tests of present parallel
code were performed for solving a channel flow. As
shown in figure 2, with the number of processor
increasing, the calculation performance becomes worse
than ideal speed-up ratio, and it seems to be related to
the interface speed for data exchange between the PC
machines used. For the different grid points, it seems
that the optimal number of process exists to obtaining
the best performance of computations.

P ro ce sso r 1

P ro ce sso r 2

P ro ce sso r 0

P ro ce sso r 3

co m p u ta tio n a l d o m a in

1,000,000
grids

200,000
grids

12

15

Number of Processor

Figure 2: Speedup ratio.


PRELIMINARY RESULTS FOR TURBULENT
CHANNEL FLOW
Here, we simulated turbulent channel flow using
different models. The LES was carried out at first
with a computation domain of 3 2 as
shown in Fig.3. The grid clustering was made near the
wall to capture the turbulence characteristics. The
minimum size of +2 is 0.75.0 in the viscous wall
unit. The time increment is 0.001. The total number of
grid points is 32 65 32 and 64 65 64. The initial
condition was given by the Moin and Kim(1980) with
the 10% turbulent fluctuation of the centerline velocity
in both y and z directions, where the centerline
velocity,
C
, is given by Dean(1978);
1/ 7
C = 7.764 Re
with Re = 180 here. The
computation was made up to t=50 and the statistical
analysis was carried out between t=40 and 50. The
CPU time using a single node of PC-cluster machine
was about 4 hours for the case of 32 65 32 grid
points.
The mean velocity profile can be seen in Fig.4.
The SF- and D-model show the larger value in the log
region, but the S- and GNS-model are in good
agreement considering their simplicity of procedure. In
Fig.5, the present LES with S-model is compared to the
DNS by Kim et al.(1987) and the LES by Tafti et
al.(1997). The turbulence intensity in three directions
was compared in Figs.68. The present LES is
qualitatively closer to the DNS using finer grid system.
It seems that the finer grid system is required for
obtaining better agreement. The total Reynolds shear
stress was compared in Fig.9. The present LES is
agreed with DNS in the shear stress.

th e d o m a in w ith
g h o st p o in ts

Figure 1: Virtual topology for MPI programming.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

In the simulations above, four types of turbulence


models are reviewed each other by comparing with
DNS and other LES. However, it seems not easy to
conclude that one model is much better than the others
according to the numerical tests. The present results
could be reversed for practical engineering problems,
such as ship flows.

urms

Ly=2

Top wall

20

40

y
z

Present LES, 32x65x32


Present LES, 64x65x64
Kim et al. (1987, DNS)
Tafti et al. (1997, LES)

80

Figure 6: Comparison of the profiles of the turbulence


intensity in the streamwise direction.

Lz=

Bottom wall

60

x
Present LES, 32x65x32
Present LES, 64x65x64
Kim et al. (1987, DNS)
Tafti et al. (1997, LES)

Lx=3

Figure 3: Flow definition in a channel

vrms

20

15
+

u =y

u = 2.5ln(y )+ 5.5

S-Model
DS-Model
GNS-Model
SF-Model

0 0
10

101

20

40

102

Present LES, 32x65x32


Present LES, 64x65x64
Kim et al. (1987, DNS)
Tafti et al. (1997, LES)

20
0
u+ = y+

10 0

40

60

80

Figure 8: Comparison of the profiles of the turbulence


intensity in the spanwise direction.

Present LES, 32x65x32


Present LES, 64x65x64
Kim et al. (1987, DNS)
Tafti et al. (1997, LES)
Map 5
Map 6

10 1

20

u+ = 2.5lny+ + 5.5

shear stress

u+

10

80

Figure 7: Comparison of the profiles of the turbulence


intensity normal to the wall.

Figure 4: Computed mean velocity profiles using


different turbulence models.

15

60

y+

w rms

u+

10

102

Figure 5: Comparison of the computed mean velocity


profiles.

Present, LES, 32*65*32


Present, LES, 64*65*64
Kim et al. (1987, DNS)
Tafti et al. (1997, LES)

0.5
0
-0.5
-1
-1

-0.5

0.5

Figure 9: Comparison of the total Reynolds shear


stress including the SGS contributions.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

molecular viscosity at both sections.

LES OF TURBULENT FLOW AROUND SHIP


MODELS

Table 1: Computed resistance components with


different grid points.

Turbulent flow around Wigley double model at


Re=4.5106

Grid
points

The well-known Wigley hull form was chosen for


the
present
LES,
whose
dimension
is
L B d=1 0.1L 0.0625L. The Reynolds number
was 4.5 106. The computational domain was set at
1.5L~2.5L in the longitudinal direction, 0~L in the
lateral direction and L~0 in the vertical direction. As
shown in Fig.10, the OH-type grid system was used
generating by the GridGen S/W. The minimum size
of grid in the lateral direction, 1 , was fixed to
for all computations, which are
0.5 10-5
corresponding to y + 5 in the viscous wall unit. The
time increment was 2 10-4.
In this study, 1/5-, 1/2-, 1- and 2-million grid
points were used, and all computations were performed
on a 24-node PC-cluster parallel machine composed of
2.6GHz-CPU, which installed in the Advanced Ship
Engineering Research Center(ASERC, http://www.
aserc.pusan.ac.kr) at Pusan National University, Korea.
The acceleration time was t=1 and computation
was made up to t=5. The total CPU time in case of
2-million is about 7.5 hours using 12-node PC-cluster
parallel machine, and it corresponds to about 10 times
speed-up ratio.
For the turbulence model, the S-model was
employed in this study because the model itself is very
simple but reasonable in accuracy and applicable to
applications.
Table.1 shows the resistance components with
different grid points. As increasing the number of grid
points it is clearly seen that all components of
resistance are converged at certain values, as shown in
Fig.11. From the result, we use a million grid points
from now.
Fig.12 shows the comparison of pressure
coefficient along the waterline and keel with the
experiments by Sarda(1986) and Shearer &
Cross(1965). Overall agreement is quite good, but
small discrepancy between the simulation and the
experiments is partially observed near bow and stern
area. In Fig.13, the wake distribution in the transverse
sections at x/L=0.899 & 1.0 are compared to the
experiments by Sarda(1986). The good agreement can
be obtained at x/L=0.899, but the computed results are
bulged more compared to the experiment at x/L=0.977.
The wake distribution would be closely related to the
distribution of the turbulent eddy viscosity, , as
shown in Fig.14. The computed turbulent eddy
viscosity is in maximum 20 times much higher than the

CT

Cp

Cf

2105

3.366 10 3

0.144 10 3

3.222 10 3

5105

2.750 10 3

0.165 10 3

2.585 10 3

10105

2.601 10 3

0.170 10 3

2.431 10 3

20105

2.583 10 3

0.172 10 3

2.411 10 3

Figure 10: Grid system of Wigley hull.


4

0.2

CT & Cf ( 10 3)

0.16

CT
Cp
Cf

0.14

2.5

Cp ( 10 3)

0.18

3.5

0.12

0.5

1.5

grid points ( 106)

0.1
2.5

Figure 11: Drag components of Wigley hull vs. grid


points used.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

0.00

Cp
0.6

8
16

12

x/L = 0.899

28
24

0.02
20

Present, Waterline

20

Exp. Sarda

0.4

Present, Keel

20

Exp. Shearer & Cross

0.04

20

z
0.2

0.06

0.0

0.0

0.08

0.2

0.4

0.6

0.8

0.00

x/L

0.02

0.04

0.06

0.08

0.10

Figure 12: Pressure coefficient along the waterline and


keel.

0.00

0.00
0.9

0.95

0.02

x/L = 1.000

15
15

0.8

0.7
0.6
0.02

0.04 12
9

present cal.
exp. (Sarda)

0.04

0.06

x/L = 0.899
0.06

0.08
0.08

0.00
0.00

0.02

0.04

0.06

0.08

0.6 0.7 0.8

0.10

0.06

0.08

0.10

Figure 14: The computed turbulent eddy viscosity,

(as a multiple of the molecular viscosity).

0.9

0.02

Turbulent flow around Series-60(Cb=0.6) double


model at Re=4106

present cal.
exp. (Sarda)

0.04

The LES were performed for the Siries 60(Cb=0.6)


hull form at Re=4 106. In the preesent study, a
half-million points were used, and other conditions of
simulation were exactly same as the previous Wigley
case. The grid system used for this study is shown in
Fig.15.
The comparison of wake distribution at x/L=1.0
(AP) is made with the measurements by Toda(1988) in
Fig.16. The overall agreement is quite reasonable, but
the hook-shaped bulge is observed excessively in the
present LES.
In Fig.17, the pressure contribution on the hull
surface in the present LES is compared with
measurements by Toda(1988) with good agreement,

x/L = 1.000
0.06

0.08

0.00

0.04

y
0.00

0.02

0.02

0.04

0.06

0.08

0.10

Figure 13: Wake Distribution at the transverse


sections.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

except for the area neighborhood of keel and stern end.


Fig.18 shows the limiting streamline at stern. Due
to the excessive bulge of wake the reveres flow area is
widely distributed.
The total drag resistance and frictional resistance
are compared with measurements. The present result is
plotted in the reasonable accuracy range of the
experiments and RANS by Kodama(1992).

0.8

0.9
0.0

-0.04

0.02

0.04

1.0
0.06 0.08

0.1

0.12

-0.02

-0.06
Computed
Measured

Figure 17: Comparison of measured and computed


pressure contours at stern.

Figure 18: Limiting streamline on hull surface at stern.

6.0x10

-03

5.0x10-03

4.0x10

-03

3.0x10-03

2.0x10

-03

1.0x10

-03

0.0x10

+00

NKK
IHI
SHI
Hiroshima Univ(4.0M)
AL Mitsui
Univ. of Tokyo
Kyushu Univ.
Yokohama Univ.
Hiroshima Univ(1.8M)
Kodama Cal. CT
Kodama Cal. Cf
Scheonherr
Present Cal. CT
Present Cal. Cf
6

10

X
X

10

Reynolds number R e

Figure 19: Comparison of measured and computed


total drag of Series 60 (Cb=0.6) hull; the all data for
experiments taken from Kodama(1992).

Figure 15: Grid system of Series 60(Cb=0.6) hull (top:


bow and bottom: stern).

CONCLUDING REMRKS
In the present study, a LES with MPI-parallelizing
algorithm was applied to investigate the turbulent flow
around a Wigley and a Series 60(Cb=0.6) double
models. Through the comparison with experiments and
other numerical results, it can be seen that the present
technique has been validated with reasonable accuracy.
However, the present LES is required more validations
for the wider range of Reynolds number and the
various shapes of ship hull form near future.

0.8
u=0.9

Computed
Measured

Figure 16: Wake distribution at x/L=1.0.


ACKNOWLEDGEMENT

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

This work was supported by the Advanced Ship


Research Center (ASERC) of the Korea Science and
Engineering Foundation.

Tafti, D.K. and Vanka, S.P., Large Eddy Simulation of


Channel Flow using Finite-Difference Techniques,
CFD Lab., Dep. Me. & Ind. Eng., UIUC, Report No.
CFD 90-01., 1990.

REFERENCES
Germano, M., Piomelli, U., Moin, P. and Cabot, W.H.,
A dynamic sub-scale eddy viscosity model, Physics
of Fluids, A Vol. 3, No. 7, 1991, pp. 1760-1765.
Horiuti, K., A proper velocity scale for modeling
subgrid-scale eddy viscosity in large eddy simulation,
Physics of Fluids, A , Vol. 5, No. 1, 1993, pp. 146-157.
Kim, J., Moin, P. and Moser, R., Turbulence statistics
in fully developed channel flow at low reynolds
number, J. Fluid Mechanics, Vol. 177, 1987, pp.
133-166.
Kodama, Y. , Computation of Ships Resistance Using
an NS Solver with Global Conservation - Flat Plate and
Series 60 (CB=0.6) Hull - J. of the Society of Naval
Architects of Japan, Vol. 172, 1992, pp. 147-155.
Lilly, D.K., A proposed modification of the Germano
subgrid-scale closure model, Physics of Fluids, A, Vol.
4, No. 4, 1992, pp. 633-635.
Miyata, H., Akimoto, H. and Hiroshima, F. , CFD
Performance Prediction Simulation for Hull-Form
Design of Sailing Yachts, J. Mar. Sci. Technol., Vol.2,
1997, pp. 257-267.
Sarda, O.P., Turbulent Flow Past Ship Hulls An
Experimental and Computational Study, Ph.D. Thesis,
1986, Mech. Eng., Univ. of Iowa.
Shearer, J.R. and Cross, J.J., The Experimental
Determination of the Components of Ship Resistance
for a Mathematical Model, Trans. Royal Institute of
Navel Archi., Vol. 107. 1965.
Smagorinsky, J., General Circulation Experiments
with the Primitive Equations. I. The Basic
Experiment, Monthly Weather Review, Vol. 91, 1963,
pp. 99-164.
Spalding, D.B., A Single Formula for the Law of the
Wall, J. Applied Mech, 1961, pp.455-458.
Takakura, Y., Ogawa, S. and Ishiguro, T., Turbulence
Models for Transsonic Viscous Flow, AIAA paper,
1989, No. 89-1952CP.
Toda, T. et al., Mean Flow Measurements in the
Boundary Layer and Wake of a Series 60 CB=0.6
Model, IIHR Report No.326., 1988

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics


St. Johns, Newfoundland and Labrador, CANADA, 8-13 August 2004

An Investigation of Propeller Inflow for Naval


Surface Combatants
Joseph J. Gorski, Ronald W. Miller and Roderick M. Coleman
(Naval Surface Warfare Center, Carderock Division)

ABSTRACT
Calculations using the Reynolds Averaged
Navier-Stokes (RANS) equations are performed for
two naval combatants, an aircraft carrier and a
destroyer. The carrier configuration consists of a bare
hull with skeg, bilge keels and outboard propeller shaft.
The destroyer consists of the hull, shafts and shaft
struts. These computations demonstrate the significant
effect propeller shafts and struts have on the propeller
inflow as compared to bare hull configurations. This
is demonstrated at both model and full scale Reynolds
numbers with the model scale computations comparing
well with experimental data. Without the propeller
shaft the dominant effect is the forward bilge and bow
dome vortex, particularly at model scale.
For
calculations at full scale the flow entering the
propellers is very nearly inviscid without the shafts
present. The shaft wake is shown to have a significant
impact on the flow entering the propeller and shaft
rotation further impacts this inflow.

transom stern, bilge keels, a docking skeg, propeller


shafts and struts (not shown). The destroyer is shown
in Fig. 2 with the major features again including a
sonar dome, transom stern, propeller shafts and struts.
For the given geometries the destroyer configuration
has a more pronounced bow dome, which leads to
stronger vortices. The destroyer has a rounded hull
form as compared to the carrier, which has a large flat
section on the bottom. This large flat section along
with the vertical wall sided hull leads to a more
pronounced bilge over the forward part of the hull that
can produce more significant bilge vortices.
Consequently, the bare hull flow generated by these
configurations can be quite different.

INTRODUCTION
To better understand the flow into the
propeller of naval combatants a Reynolds-Averaged
Navier-Stokes (RANS) code is used to compute the
flow fields about both a carrier and a destroyer
configuration. For naval vehicles the flow into the
propellers is significantly influenced by the upstream
hull form, Gorski (2001). This includes the boundary
layer generated on the hull as well as any vortical flow
that may form such as that from the bow/sonar dome or
bilges. Additionally, wakes are formed from upstream
bilge keels, shafts, and supporting struts. Of particular
significance for many naval combatants are the shaft
and strut wakes, which are immediately upstream of
the propellers. RANS codes provide a means of
predicting such flow fields and have been demonstrated
for a variety of complicated ship flow fields, Gorski
(2002).
Major features of the carrier configuration
are shown in Fig. 1. These include a bulbous bow, a

Fig. 1 Carrier hull form.

Fig. 2 Destroyer hull form.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

For both configurations only the flow below


the waterline, depicted in the figures, is computed in
the current effort. There have been few predictions of
carrier flow fields in the past and it is particularly
challenging at full scale due to the extremely high
Reynolds numbers. The destroyer flow field has been
computed numerous times as David Taylor Model
Basin Model 5415 and was one of the hull forms used
for the Gothenburg 2000 workshop, Larsson et al
(2003). However, the computations have largely been
for the bare hull configuration and do not address the
complexities of the shafts and struts. Calculations
done here are for both bare hull and appended
configurations to highlight differences to the propeller
inflow due to the shafts and struts. The current effort
uses structured grids which makes it very difficult to
include all of this complexity in the calculations.
Consequently, the final computations for the carrier
contain the outer shaft, but do not contain the inboard
shaft/strut arrangement or the outboard struts
supporting the shaft. Although only the outer shaft is
included in the carrier calculations it is shown to have a
significant impact on the flow into the outboard
propeller. For the destroyer the shafts and struts are
included in the final calculations. In the future
including all the complexity of interest should become
more feasible as the use of unstructured or Chimera
RANS codes becomes more routine.
To better estimate the quality of a simulation
there have been efforts to develop uncertainty estimates
and validation procedures for computations (e.g. AIAA
1998; Roache, 1998). This is an area of significant
importance as the computational community tries to
provide metrics for the quality of a computation and
more work needs to be done in this area. To this end
though, it is becoming largely recognized that one
cannot declare a particular code validated, even with
good predictions for a particular flow, because
individual solutions depend on many factors including:
geometry definition, grid quality, turbulence modeling
and user experience among other variables. However,
when applying well developed RANS codes properly,
experienced users can obtain very meaningful
information on complicated hull flows, including hull
modification and scale effects, without a formal
validation procedure. For the complicated geometries
studied here it is difficult to go through a formal
uncertainty procedure with the structured grids.
Nonetheless, grid resolution studies are included to
indicate sensitivity of the solutions for Model 5415.
Calculations are carried out at model scale
conditions for comparison of the computational
predictions against available experimental data. The
calculations of interest are also carried out on the full
scale hull to produce understanding of the flow the real

ship would experience. For Model 5415 all of the


predictions are for the nominal wake, but the carrier
calculations are also performed with the propeller
modeled as an actuator disk and with the shaft rotating
to highlight these individual effects.
RANS calculations for predicting full-scale
ship flows are becoming more routine. However, there
are issues involved in the full scale calculations in
addition to the need for more grid points for boundary
layer resolution. Surface roughness is often ignored,
Patel (1998), as in the current full scale simulations,
but can be important for the actual ship. Models tested
are often smooth allowing surface roughness to be
ignored. Real ships are considerably rougher as built
and only worsen with time at sea. Hence, the full scale
computations probably predict thinner boundary layers
than actually exist and care must be taken when
interpreting the results.
FLOW SOLVER
To compute the viscous flow field the
incompressible Reynolds Averaged Navier-Stokes
equations are solved using the Mississippi State
University code, UNCLE, developed by Taylor et al
(1991, 1995). The UNCLE code is one of two RANS
codes used for the ONR Surface Combatant
Accelerated Hydro S&T Initiative to provide
documented computational solutions for innovative
propulsor/hull concepts of interest for DD(X) and
beyond, Gorski et al (2002b).
The equations are
solved using the pseudo-compressibility approach of
Chorin (1967) where an artificial time term is added to
the continuity equation and all of the equations are
marched in this artificial time to convergence. Only
steady state computations are performed for this effort.
For the present calculations a third-order upwind
biased discretization, based on the MUSCL approach
of Van Leer et al (1987), is used for the convective
terms. The equations are solved implicitly using a
discretized Newton-relaxation method, Whitfield and
Taylor (1991), with multigrid techniques implemented
for faster convergence. The turbulence model used for
the present calculations is a k- model. An important
factor in being able to compute and evaluate the hull
modifications and operating conditions of interest is the
implementation of a parallel version of the UNCLE
code. The code uses MPI for message passing due to
its portability. To run in parallel the computational
grid is decomposed into various blocks, which are sent
to different processors. Load balancing is obtained by
making the blocks as equal in size as possible. More
details of the solver can be found in the various
references provided.

2
Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Computations are carried out with both


double-model and free-surface conditions applied at
the water surface.
The double-model condition
simulates the flow about the double body formed by
reflecting the hull about the undisturbed water level.
The undisturbed water level is treated as a symmetry
plane where the vertical velocity is set to zero. The
double-model condition is a good approximation to a
free-surface condition if the speed of the hull is low or
if the flow region of interest is sufficiently far from the
water surface. When the free surface is included only
inviscid approximations to the free-surface boundary
conditions are applied since free-surface viscous layers
are not of interest here. The shape of the water surface
is computed subject to the conditions that the flow is
tangential to this surface and that the pressure is
atmospheric. The linearized free-surface option of the
code is used where these conditions are applied at the
undisturbed water level rather than at the actual
computed free-surface level. This option avoids the
complication of having to move the grid to conform to
the free-surface shape as it is computed. Similar to the
double-model
approximation,
the
linearized
assumption is a good approximation if the wave slopes
are small or if the flow region of interest is sufficiently
far from the water surface. Comparison of results
obtained with the double-model condition with those
obtained with the linearized free-surface condition
shows that for the purposes of this propeller inflow
study, use of the double-model condition is quite
sufficient.
DESTROYER COMPUTATIONS
Perhaps the most extensively measured naval
combatant is the bare hull version of DTMB Model
5415 (Ratcliffe, 1998), and its geosyms. Tow tank data
include: resistance, sinkage and trim, wave profiles,
near- and far-field wave elevations, mean flow (taken
with pitot probe and LDV), and turbulence data. A
unique aspect of this model is that is has been used in a
collaborative effort between the David Taylor Model
Basin, the Iowa Institute of Hydraulic Research, and
the Italian Ship Model Basin. This was done to obtain
a complete data set, at various conditions, as well as
provide information on facility biases as there were a
number of repeat runs between the facilities. Because
the data is meant for CFD validation, much care was
taken in obtaining the data and uncertainty estimates
are provided based on ITTC recommendations.
However, as pointed out by Stern et al (2000), in a
review of the effort, uncertainty estimates by individual
facilities are often too optimistic and there appear to be
unaccounted for bias and precision limits in the tests.
Considering the care taken in these tests one can only
assume such issues also exist with other data sets from

other facilities. However, this is an excellent data set


for code comparison and simply means care must be
taken about forming absolute conclusions in
comparisons with computations. A fully appended
model was also tested at DTMB including shafts,
struts, and rudders. The model was run straight ahead
at a speed of 4.0 knots for a corresponding model scale
Reynolds number of 12 million, based on length, and
Froude number of 0.277. LDV data was obtained near
the propeller plane, with and without the propeller
operating.
A number of configurations have been
calculated for this study. An initial calculation is done
for the bare hull with bow dome and transom stern as a
baseline. These bare hull calculations were performed
with both double-model and linearized free surface
boundary conditions. Numerous computations for the
bare hull Model 5415 have been done in the past so the
emphasis in this study is on the appended hull with
shafts and struts. Since there is little difference in the
computed flow at the propeller disk with the different
free-surface boundary conditions the appended
calculations are all done with the double-model
approximation. In addition, an appended hull with
shafts only, and no supporting struts, is included to
isolate the impact of the shaft from the combined
shaft/strut geometry into the propeller disk. These
calculations are run at the model scale Reynolds
number of 12 million and compared with the
experimental data of Chesnakas. Additionally, a fullscale calculation is performed for the appended hull
with shafts and struts at a Reynolds number 1.4 billion,
based on length. For all calculations the hull is sunk
and trimmed by amounts given from the experiment for
the corresponding Froude number of 0.277.
Grid Generation
For all calculations structured grids are used.
For the bare hull geometry this does not pose a
problem. The grid about the bare hull is the typical HO type structured grid. It consists of 65 points radially,
89 points circumferentially, and 177 points along the
hull. The total number of points is about 2 million for
the ship with port/starboard symmetry assumed. The
spacing of the first point off the wall is about 1.0x10-6
resulting in y+ less than 1.0 over most of the hull. The
grid extends one ship length forward of the hull, two
ship lengths aft and one ship length radially outward.
At all far field boundaries, characteristic boundary
conditions are applied and on the mean free surface,
the double model or linearized free surface boundary
condition is applied.
The symmetry boundary
condition is applied on the centerplane where port-

3
Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

starboard symmetry is assumed. The grids are created


using Pointwises grid generator GRIDGEN.
The grid about the appended hull is much more
complicated. Including the shafts and struts in the
calculation using structured grids is a very time
consuming process. The necessity of point-to-point
matching and the acuteness of the shaft-hull
intersection can result in very highly skewed grids near
the region of the shaft-hull intersection.
The
developing upstream flow containing the bow dome
vortex and the hull boundary layer must be able to pass
through this region adequately. The fully appended
grid generated here is a variation of the barehull grid.
A bubble or blister region is created to contain the shaft
and struts. The bubble begins on the hull a small
distance upstream of the hull-shaft intersection and
continues to the downstream farfield boundary. The
hull surface upstream of the bubble and the outer
surface of the bubble combine to give a new surface
about which an H-O grid is created. The complex
blocking arrangement on the hull surface is illustrated
in Fig. 3.

supporting struts Grid 1 is used with the struts


collapsed.

Fig. 4 Cross section of grids near the shaft.


Model Scale Computations
For Model 5415 the flow near the bow has a
large downward component resulting in the creation of
a bow dome vortex. This disturbance grows as it is
convected downstream and combines with the hull
boundary layer. This vortex may also be enhanced by
the general downward and inward flow over the
forward bilge. For the bare hull configuration this is
the major flow disturbance into the propeller plane. A
computation showing axial velocity contours along the
hull is shown in Fig. 5. The decrease in draft at the
stern accentuates this wake as the propeller plane is
approached.

Fig. 3 Surface grid for Model 5415.


Topologically the grid inside of the bubble is
cylindrical with one end on the hull and the other end
at the farfield boundary. O-grids are created, inside the
bubble, around the propeller shaft. As the crosssections or O-grids march toward the hull they begin to
turn rapidly toward the hull which can result in highly
skewed cells. Two grids are created for the shaft/strut
configuration to check for grid independence of the
solution. A cross section in the bubble region for both
grids is shown in Fig. 4, illustrating the resolution
difference. The bubble region of Grid 2 consists of
twice the number of points of Grid 1, in both the radial
and tangential directions, while keeping the wall
spacing the same. The total number of points for Grid
1 is 7 million points while Grid 2 contains 16.6 million
points. Each grid is decomposed into 67 blocks for
parallel processing.
For the calculation without

Fig. 5 Computed axial velocity contours for


Model 5415.
When shafts and struts are included, the
upward and inward flow due to the decrease in draft
and width of the hull, results in a shaft wake above the
shaft and beneath the hull. As seen in Fig. 6 the bow
dome vortex still persists at the propeller plane.
However, the shaft and strut wakes combine with the
bow dome vortex and the hull boundary layer to further
complicate the flow into the propeller plane, Fig. 7.

4
Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

propeller inflow, a calculation is also performed


without the struts. The results, shown in Fig. 11,
indicate that the struts tend to widen the wake
immediately above the shaft. It appears that the wake
of the shaft joins with the inboard strut wake, which is
more in the cross-flow direction, thickening the overall
disturbance. However, the comparison with the shaft
alone demonstrates that the shaft wake dominates
significantly over the strut wakes as long as the struts
are aligned with the flow.

Fig. 6 Axial velocity contours for Model 5415


with shafts and struts.

Fig. 8 Computed and measured axial velocity at


the propeller plane.

Fig. 7 Computed shaft and strut wakes for


Model 5415.
A comparison of the calculated axial velocity
at the propeller plane with the measured data is shown
in Fig. 8. The calculation captures the angled v-shaped
wake formed from the combination of the bow dome
vortex, hull boundary layer and the shaft wake. The
calculation also shows the presence of strut wakes in
the flow field. The struts wakes tend to sharpen the
wakes v-shape. This feature is missing in the
experimental data, probably because of the coarseness
of the measurement data locations as seen in Fig. 9.
There are almost no measurement locations that can
pick up the thin strut wake, whereas the computation
has resolved these wakes much better. This is further
evident in Fig. 10 where the calculated results are
interpolated to the data measurement locations. In this
figure the computed strut wakes have not been
preserved by interpolation to the measurement
locations. To evaluate the effect of the struts on the

Fig. 9 Axial velocity with computational and


experimental grids overlaid.

5
Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Fig. 12 Comparison of computed axial velocity


on Grids 1 and 2.
Full Scale Calculations

Fig. 10 Computed axial velocity interpolated


onto the experimental grid.

For full scale conditions the Reynolds number,


based on length, is approximately two orders of
magnitude larger than the model scale Reynolds
numbers and the Froude number is kept the same. For
the bare hull calculation the flow field itself is very
similar to the model scale flow, shown in Fig. 5, but
with a thinner boundary layer and smaller vortex
structure formed over the front of the hull. This leads
to a thinner wake at the propeller plane as shown in
Fig. 13. Here a comparison with the model scale
predictions indicates that there is almost no hull wake
entering the propeller disk, indicated by the circle, at
full scale compared to what would enter the disk at
model scale. Consequently, care must be exercised
when using model scale experimental data to infer full
scale propeller inflow.

Fig. 11 Computed axial velocity without the


struts as compared to the experimental data.
Finally, a calculation is performed to check
for grid independence of the solution. The difference
in grid resolutions is shown in Fig. 4 and the computed
axial velocity at the propeller plane for each grid is
shown in Fig. 12. Grid 2 is the finer of the two grids
by several million points. The two solutions are
basically the same, however Grid 2 tends to resolve
some features better. For example, the inboard strut
wake is more clearly defined and the contours between
the centerplane and inboard strut tend to curve away
from the centerplane, as in the experiment.

Fig. 13 Bare hull model and full scale


prediction at the propeller plane.

6
Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

The full-scale calculation performed for the


shafts-and-struts configuration also exhibits significant
differences from the model scale predictions in the
flow at the propeller plane. Here the Reynolds number
is set to 1.4 billion, based on length. Grid 1 is modified
by decreasing the spacing of the first point off of the
wall to about 10-7 to maintain a good y+ value. As a
result of this new clustering some of the grid resolution
away from the hull is sacrificed. Results for this
calculation are shown in Fig. 14. The shaft wake is
still evident as are the wakes of the struts. However, it
is seen that the size of these wakes is reduced
significantly, as compared to the model scale
computation shown previously in Fig. 7. Additionally,
the hull boundary layer and vortex structure is reduced
significantly. Consequently, only a small portion of
the propeller disk experiences a flow disturbance as
almost none of the flow from the hull boundary
layer/wake enters it. This is consistent with the bare
hull calculation of Fig. 13. However, the shaft and
strut wakes are still present and are the only significant
deviation from the free stream flow entering the
propeller disk, Fig. 15.

Fig. 14 Computed axial velocity contours at full


scale.

Fig. 15 Computed axial velocity at the propeller


plane at full scale.
Comparison of Wake Fraction
In this section, a comparison of the averaged
axial velocity through the propeller disk area is made.
To integrate the velocity, a grid is created covering the
disk area, upon which the measured and calculated data
are interpolated. The data is then integrated to
calculate an average axial velocity. The interpolated
data and average axial velocities are shown in Fig. 16.
The bare hull shows the least disturbance (Uave =
0.950) for the model scale flows since there is no wake
deficit from the shaft or struts. The fact that it is only
about 0.5% greater than the experimental data seems
misleading. Some of this may be just coincidence,
since the wake itself is significantly different than the
measured weak. It may also be be due to the fact that
the experiment is missing the strut wakes as already
discussed. The results for the calculated shafts-andstruts case is the same between the two grids with Uave
= 0.939 with Grid 2 and Uave = 0.938 with Grid 1.
Because of the added disturbance from both the shafts
and struts the average velocity is the smallest and is
less than the measured value. However, it could
conceivably be more accurate than the measured value
due to the better resolution of the strut wakes. The
results for the strutless configuration, where the wake
is due to the shaft alone, raises the average axial
velocity to Uave = 0.944, which is very close to the
experimental value of 0.945. Finally, the full-scale
results for the shafts-and-struts configuration, where
the wakes provide less velocity deficit than the model
scale case, raises the average velocity to Uave = 0.972.

7
Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

CARRIER COMPUTATIONS
The carrier configuration consists of a bulbous
bow, a transom stern, bilge keels, a docking skeg,
propeller shafts and struts. The hull is sunk and
trimmed appropriately, based on experimentally
obtained values, and run at model and full scale
Reynolds numbers, based on hull length, of 35 x 106
and 48 x 108, respectively. Both cases are for a Froude
number of 0.277 when using the linearized free surface
boundary condition with the grid generated up to the
design waterline. Individual hull effects are evaluated
by performing a number of calculations.
These
include a coarse grid bare hull calculation without bilge
keels and a fine grid calculation of the bare hull with
bilge keels. Both of these bare hull calculations
include the skeg and are performed at model and full
scale. The outboard propeller shaft is also included for
a number of calculations. Unfortunately, it was not
possible to include the supporting struts or inboard
shaft arrangement in the current calculations. With the
outboard propeller shaft present propelled calculations
are done using an actuator disk model with and without
the shaft rotating.
Again, these calculations are
performed at both model and full scale.

a) Experiment; Uave = 0.945

b) Bare hull calculation; Uave = 0.950.

Grid Generation
c) Shaft and strut calculation; Uave = 0.939 (Grid 1),
Uave = 0.938 (Grid 2).

d) Calculation without struts; Uave = 0.944.

e) Full scale calculation; Uave = 0.972.


Fig. 16 Comparison of flow at the propeller disk
and average axial velocity.

As with the destroyer computations an


important element of the current effort is obtaining a
good computational grid. For the given predictions the
computations are limited to one-half the flow region by
taking advantage of the port/starboard symmetry.
Various structured grids are employed ranging from
one of about 2.5 million points, used for preliminary
computations of the hull with neither bilge keels nor
propeller shafts, one of 5.9 million points, which
includes bilge keels, and one of about 7.2 million
points including the outboard shaft. All grids are of HO topology with grid lines wrapping around the hull
from the center plane to the undisturbed water level in
transverse cross sections. For the finer grids this
consists of 129 points. Radial lines, which have 129
points for the finer grids, extend from the hull outward
to a far field boundary about one body length from the
hull centerline. In horizontal sections grid lines wrap
around the bow but continue past the transom to the
downstream boundary about two body lengths from the
stern. A transom grid section fills in the trough
extending from the transom to the downstream
boundary. For the 5.9 million point grid there are 225
points along the length of the hull. For the grid with
the outboard shaft 305 points are along the length of
the hull with the additional points inserted in the shaft
region for better definition. The surface grid on the
hull in the stern region for the geometry with the shaft

8
Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

is shown in Fig. 17. Here the clustering toward the


skeg and bilge keel can be seen.

which compares the bare hull grid to the grid with the
outboard shaft. Again, both grids have 129 points
going out radially from the hull to the far field. Only a
single grid is generated for the case with shaft. A
second grid reclustered to the walls for full scale
calculations is not generated. However, full scale
calculations are performed on this grid and the results
presented in later sections. Again, only the outboard
shaft is included. This grid with the outboard shaft is
split into 160 blocks for parallel processing.

Fig. 17 Surface grid with the outboard shaft,


view from underneath the hull.
The nearest points to the hull are about 10-6
body lengths from the wall for the model scale grid.
This tight clustering is needed to ensure that the
minimum spacing normal to the body surface is less
than a y+ value of 3, for good behavior of the
turbulence model. As one goes to higher Reynolds
numbers, it is required that the relative distance of the
first point off of the wall become smaller. For the
current full scale computations these nearest points are
about 10-8 body lengths from the wall to yield a y+
value less than 5. Although generating the bare hull
grid is rather straightforward such tight spacing at the
wall for the full-scale computations can pose
significant problems, particularly around the skeg,
transom and other regions where grid points need to be
concentrated due to geometric complexity. Although
these difficulties can be overcome, as demonstrated
here, it becomes necessary to increase the overall
number of grid points or sacrifice the grid resolution in
the boundary layer. An initial grid of approximately
2.5 million points, which is split into 44 blocks for
parallel processing, is used for the bare hull
computations at model scale only. The finer grid of 5.9
million points, split into 84 blocks for parallel
processing, is used when the bilge keels are added.
This grid is used for both model and full-scale
calculations with the points clustered more to the wall
for the full scale as already discussed. This tighter
clustering pulls more points closer to the wall, which
leads to coarser grids away from the hull. Hence
resolution of far field flow features can be
compromised.
The final grid of 7.2 million points
includes the outboard shaft. To include the shaft,
cylindrical grids wrapping around the shafts are
imbedded into the hull-conforming grid. Here some of
the grid surfaces that are near the hull upstream of the
shaft leave the hull at the start of the shaft and move
outward away from the hull with the shaft to maintain
grid clustering around the shaft sufficient to capture
flow physics. This clustering is displayed in Fig. 18,

Fig. 18 Cross section of grid with and without


outboard shaft.
Bare Hull Model Scale Calculations
Model scale calculations, at Reynolds number
of 35 x 106 based on length, are first done with the 2.5
million point grid without bilge keels. These are for a
Froude number of 0.277 while using the linearized free
surface boundary condition with the grid generated up
to the design waterline. The computed free surface
elevations for the model scale prediction are shown in
Fig. 19. This computed free surface height along the
hull agrees well with experimentally obtained data1
shown in Fig. 20. Computed full-scale wave heights
are very similar to those for model-scale. This was
shown previously by Krueger et al (2002) for Model
5415 where the only noticeable difference in computed
free surface heights between model and full scale
predictions was in the area behind the transom. The
inclusion of the outboard shaft has no noticeable
impact on the computed free surface heights.

Experimental data provided courtesy of Code


5200, Naval Surface Warfare Center, Carderock
Division, West Bethesda, MD.

9
Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

and the forming bilge vortex.


This vortex travels
downstream and is the dominant feature in the
propeller plane area as seen in Fig. 24. As shown, the
wake retains much of the hull shape as it propagates
downstream and takes on a very flat behavior at the
stern on the underside except for the skeg wake. Due
to the free surface interaction, there is a downward
component of the flow at the stern, and then an upward
flow toward the transom. The bilge/bilge keel vortex
continues as a dominant feature into the wake.

Fig. 19 Computed wave heights for Fr = 0.277.

Fig. 21 Surface streamlines and axial velocity


contours for the model scale calculation.

Fig. 20 Computed hull free surface height as


compared with experimental data.
Results for the 5.9 million point grid, which
also includes bilge keels, are shown in Fig. 21. This
figure shows surface streamlines and axial velocity
contours at various locations along the length of the
hull. It can be seen that the boundary layers are very
thin even at model scale over most of the hull. Flow is
generally downward over the forward part of the hull
and then runs axially along the hull mid-section. An
enlarged view of the bow area, Fig. 22, indicates that
the downward flow creates a vortex over the bow dome
that convects down the hull centerline. However, this
vortex is rather weak and has negligible impact on the
hull boundary layer midway down the length of the
hull. The hull has rather sharp bilges and an expanded
view of the midships region in Fig. 23 indicates a bilge
vortex forming toward the stern of the hull. There also
appears to be some interaction between the bilge keel

Fig. 22 Flow field in the bow region.

Fig. 23 Flow field in the midships region.

10
Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Fig. 24 Flow field in the stern region at model


scale.
The bilge keel can affect the wake at the
propeller disk. A cross section at X/L = 0.5, Fig. 25,
indicates that there is still some secondary flow from
the bow vortex which creates flow outward from the
centerline towards the bilges. It can be seen in a cross
section at X/L = 0.55, Fig. 26, that the generally
outward and upward flow around the bilge creates the
downstream vortical flow and it appears the bilge keel
effects this. This bilge/bilge keel vortex is the
dominant flow into the propeller plane at model scale
when the shafts and struts are not present as shown in
Fig. 27. The bilge keel effect may be smaller than the
effect due to the bilge itself. This is discerned from a
comparison of the axial velocity contours for the
calculation without bilge keels on the 2.5 million point
grid, with those of the computation on the 5.9 million
point grid with the bilge keels. There are some
differences between the two, but the differences are
small and quite likely due to the grid resolution
differences as much as the bilge keel effect. The finer
grid appears to predict a somewhat stronger bilge
vortex as well as a more pronounced skeg effect, which
is consistent with better grid resolution. Details are
reported in Gorski et al (2002a).

Fig. 26 Axial velocity contours and secondary


flow streamlines at X/L = 0.55.

Fig. 27 Axial velocity contours at the outboard


propeller plane with the bilge keels present.
Bare Hull Full Scale Calculations
Full-scale calculations, at a Reynolds number
48 x 108 based on length, are done with the 5.9 million
point grid, which includes bilge keels. This grid has
the same structure as the model scale grid, but has been
more tightly clustered to the hull as already discussed.
At full scale the boundary layers are relatively thinner
than at model scale so the bow vortex is less noticeable
than that for the model scale computation. A view at
the stern, shown in Fig. 28, indicates the formation of a
bilge/bilge keel vortex. However, this is the only
feature that stands out and the boundary layer wake of
the hull is almost nonexistent. Again, there is some
interaction between the bilge keel and bilge vortex
forming. The effect of the sharp bilge appears to be
more significant than effects due to the bilge keel, but

Fig. 25 Axial velocity contours and secondary


flow streamlines at X/L = 0.5.

11
Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

it is not conclusive. There is a skeg wake, but it does


not interfere with the propeller disks. The surface
streamlines look very similar to those at model scale,
but the flow at the propeller planes is significantly
different than that seen at model scale. At the inboard
plane the propeller disk basically experiences inviscid
flow, Fig. 29, at the local angle of attack due to the hull
shape. At the outboard propeller plane there is a slight
perturbation to the inviscid flow, due to the bilge
vortex, but this is negligible, Fig. 30. From these
calculations it is obvious that there is a tremendous
difference between model and full-scale propeller
inflow.
The computations and turbulence models
should be able to predict the correct trends with
Reynolds number so some of this is the reality of the
scale effects, although a grid resolution study has not
been performed at this full scale Reynolds number.
One must also consider the fact that such full-scale
calculations do not account for roughness effects, as
already mentioned, which can be significant on a real
ship.

Fig. 28 Axial velocity contours and surface


streamlines for the full scale calculation.

Fig. 29 Axial velocity contours at the inboard


propeller plane at full scale.

Fig. 30 Axial velocity contours at the outboard


propeller plane at full scale.
Outboard Shaft Effects
Calculations are also performed at both model
and full scale, which include the outboard shaft. The
inboard shaft is not modeled nor the supporting struts
for the outboard shaft. The grid consists of 7.2 million
grid points for half the geometry. Computed axial
velocity contours near the stern of this hull are shown
in Fig. 31. It is seen that the bilge/bilge keel vortex
flow that is the dominant feature at the stern for the
bare hull configuration flows directly into the outboard
shaft.
The shaft changes this flow significantly
creating an additional wake deficit in the outboard
propeller disk. Also seen in Fig. 31 is a noticeable
effect on the flow downstream of the outboard
propeller shaft, due to the interaction of the bilge
vortex and shaft.

Fig. 31 Computed axial velocity contours with


the outboard shaft at model scale.

12
Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

A comparison of the bare hull axial velocity


contours and those of the current configuration with
shaft are shown in Fig. 32. There is a velocity deficit
due to the shaft wake between the hull and shaft, which
is not present in the bare hull calculation. As with the
destroyer, the flow at this location has a strong upward
component due to the hull draft decreasing toward the
stern. The flow follows the hull form and its vertical
component creates a wake due to the shaft, between the
shaft and the hull. This added velocity deficit is a
significant feature of the propeller inflow.
A
comparison of axial velocity at the outboard propeller
disk between experiment2, the bare hull computation,
and this computation with the shaft shows a much
better comparison with the shaft present. This is
demonstrated in Fig. 33. The bare hull flow has none
of the sharp wake-like velocity deficit.
The
computation with the shaft agrees well with the
measured data, which also has the wake-like deficit.
The measured data has a somewhat wider wake, which
is likely due to the struts in the experiment. Here
again the shaft wake is a dominant feature into the
propeller disk and the inclusion of the shaft alone has
provided a reasonable comparison with the
experimental data as compared to the bare hull
computation. However, the average velocity for the
calculation with the shaft alone, Uave = 0.951, is still
higher than the measured value of 0.92 indicating the
inclusion of the struts for this computation is important
to the propeller inflow prediction.

Fig. 32 Comparison of axial velocity at


outboard propeller plane with/without shaft.

Experimental data provided courtesy of Code


5200, Naval Surface Warfare Center, Carderock
Division, West Bethesda, MD.

Fig. 33 Computed and measured axial velocity


at the outboard propeller plane.
The full-scale calculations with shaft are done
on the same grid as used for the model scale
calculations so the clustering near the wall is less than
satisfactory. This calculation again showed largely
inviscid flow at the propeller with some slight effect
due to the shaft. These features will be discussed
further in conjunction with the propelled calculations.
Propeller Modeling
Calculations are performed at both model and
full scale with the outboard propeller simulated as an
actuator disk. These calculations are done with the
outboard shaft present, but not rotating.
With the
actuator disk model a single grid plane is chosen at the
propeller plane and a constant Kt and Kq imposed over
the disk in this plane. With this approach the total Kt
and Kq over the disk is at the correct level, but a
constant value is imposed so there is no accounting for
radial distributions of force. Additionally, different
levels of thrust and torque produced in the inner part of
the disk interacting with the hull/shaft wake are not
modeled differently than that exposed to the free
stream flow. Shown in Fig. 34 are the computed axial
velocity contours at the outboard propeller plane at
model scale. The propeller modeling accelerates the
flow to above free stream levels over most of the disk
area. However, the wake of the hull and outboard
shaft are still present and provide a velocity deficit over
parts of the disk area. The secondary flow is also
changed due to the rotation of the propeller, but the
shaft wake stays in much the same position as in the
unpropelled case. The full-scale calculation, Fig. 35,
shows a largely inviscid flow as expected. The axial
velocity has increased to similar levels as in the model
scale calculation. However, the velocity deficit due to
the hull/shaft wake which exists at model scale is
largely nonexistent at full-scale, which is partially due
to the coarse grid at full scale with the shaft present.

13
Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

rotation, and the subsequent no-slip boundary condition


on the shaft, pulling more low momentum flow from
the area between the shaft and the hull down toward
the shaft. Here it is also apparent that this shaft wake
has rotated around with the shaft from its previous
position when the shaft is fixed. This wake rotation is
also apparent with the full-scale calculation, Fig. 38.
Here a wake defect appears due to the shaft, which is
not apparent in the full-scale case without shaft
rotation. This velocity deficit is very thin, but does
change the flow from the cleaner free-stream type flow
seen without the shaft rotation at full scale. More
details of the carrier computations are found in Gorski
et al (2002a).

Fig. 34 Model scale axial velocity with the


outboard propeller modeled.

Fig. 36 Axial velocity contours with the shaft


rotating.

Fig. 35 Full scale axial velocity contours with


the outboard propeller modeled.
Shaft Rotation Effects
The previous propelled calculations are
performed without shaft rotation.
Additional
calculations are performed using the same actuator disk
model, but with the shaft rotating. The shaft rotation
has an effect on the flow. Axial velocity contours for
the model scale calculation, near the shaft, are shown
in Fig. 36. This calculation includes the propeller
modeled with an actuator disk near the end of the shaft.
As seen, the wake of the shaft between the shaft and
hull is pulled around by the shaft to create a curved
wake due to the shaft rotation. At the propeller plane
the model scale prediction, shown in Fig. 37, has a
similar overall wake deficit in the top of the propeller
disk as for the fixed shaft. However, the extent of the
wake has become thinner with lower velocity reaching
to the shaft. This appears to be a result of the shaft

Fig. 37 Model scale calculation with shaft


rotation.

14
Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Fig. 38 Full-scale calculation with shaft


rotation.
CONCLUSIONS
RANS calculations are performed for both a
carrier and a destroyer hull configuration.
These
computations are performed at model and full scale
Reynolds numbers for bare hull and appended variants
of these hulls. For the carrier configuration the hull
includes the bow dome, skeg, bilge keels, and outboard
propeller shaft. For the destroyer configuration the
geometry includes bow dome, propeller shafts and
supporting struts. Significant differences are observed
in the calculations due to changes in Reynolds number
from model to full scale. The model scale calculations
demonstrate a significant hull wake entering the
propeller planes. For the carrier, the flow entering the
outboard propeller plane is complicated by a
bilge/bilge keel vortex formed upstream.
The full
scale bare hull calculations for both configurations
show nearly inviscid flow entering the propeller disks
with minimal hull boundary layer and wake. From
this one could conclude that measured model scale
inflow to the propeller planes has a larger wake deficit
than seen at full scale.
The shafts significantly impact the propeller
inflow and appear to be at least as important as the hull
boundary layer and wake effects. The model-scale
calculations without the shaft show a rather benign
boundary layer flow at the propeller plane, whereas the
experimental data demonstrates a significant shaft/strut
wake for both the destroyer and carrier configuration.
This relative importance of the shaft is confirmed by
the calculations. The destroyer computations with the
shaft and struts compare well with the measured data
and are significantly different than the bare hull flow
field. In addition, a computation without the struts
indicates the shaft wake is the more dominant feature

for the destroyer. From these comparisons it appears


the added effect of the shaft supporting struts may be
secondary as long as the struts are aligned with the
flow. For the carrier, model scale computations with
the outboard propeller shaft present agree well with
measured data. However, here the struts appear to
have more of an impact on the average velocity at the
propeller plane and should be included in future efforts.
Again, the full scale calculations are significantly
different than the model scale calculations. For the
destroyer calculation the grid is well refined for the full
scale calculation with shafts and struts present and the
wake of these features persist. Due to the thinner
boundary layers there is a significantly higher velocity
entering the propeller disk at full scale than at model
scale. Again, this indicates that measured model scale
inflow to the propeller planes has a larger wake deficit
than seen at full scale. For the carrier computations the
full scale grid is coarser than desired with the shaft
present.
However, it is seen that shaft rotation
impacts the wake at model scale and enhances the shaft
wake at full scale. This rotation created a wake
between the hull and shaft producing a wake deficit not
seen in the stationary shaft calculations at full scale.
It is apparent that the local flow due to the
presence of the shafts and struts is very important to
the propeller inflow and needs to be included in
calculations relevant to the real hull form. The current
work also raises additional questions about the details
of the flow into the propeller and upstream influences
at different scales, which require further study. One
thing that needs to be considered when using RANS
codes for such complicated geometries is the ability to
quickly grid the various important details. The current
geometries with their shafts and struts do require
significant amounts of time to be gridded properly and
the solutions are only as good as the grids used for the
computations.
However, with good grids and
appropriate resolution, the RANS calculations provide
very useful information concerning flow physics and
scale effects. Future work will be made more efficient
through the use of unstructured or Chimera grid
technology.
These maturing grid generation
techniques, in conjunction with the appropriate flow
solvers, may more readily provide accurate solutions
for these complex geometries.
ACKNOWLEDGEMENTS
This work was supported by NAVSEA 05H (K.
McCreight) for the carrier effort and ONR (P. Purtell)
for the Model 5415 effort. This work was also
supported in part by grants of computer time provided
by the DOD High Performance Computing
Modernization Office at the Arctic Region

15
Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Supercomputing Center, the Maui High Performance


Computing Center and the Aeronautical Systems
Center at Wright Patterson Airforce Base under the
Project CFD for Naval Vehicles. Software licenses for
GRIDGEN were provided by the U.S. Navy
Hydrodynamic/Hydroacoustic Technology Center.

Larrson, L., Stern, F., and Bertram, V., Benchmarking


of Computational Fluid Dynamics for Ship Flows: The
Gothenburg 2000 Workshop, Journal of Ship
Research, Vol. 47, No. 1, March 2003, pp. 63-81.

REFERENCES

Ratcliffe, T., Validation of Free Surface Reynolds


Averaged Navier-Stokes and Potential Flow Codes,
Proc. 22nd Symposium Naval Hydrodynamics,
Washington D. C., 1998, pp. 964 980.

AIAA, Guide for the Verification and Validation of


Computational Fluid Dynamics Simulations, G-0771998, 1998.
Chesnakas, C., Model 5415, DTMB Web page,
http://conan.dt.navy.mil/5415.
Chorin, A. J., A Numerical Method for Solving
Incompressible Viscous Flow Problems, Journal of
Computational Physics, Vol. 2, 1967, pp. 12-26.
Gorski, J. J., Marine Vortices and Their
Computation, Proc. NATO RTO Applied Vehicle
Technology Panel Symposium on Advanced Flow
Management, Loen, Norway, May, 2001.
Gorski, J. J., Present State of Numerical Ship
Hydrodynamics and Validation Experiments, Journal
Offshore Mechanics and Arctic Engineering, Vol. 124,
May 2002, pp. 74-80.
Gorski, J. J., Haussling, H. J. and Coleman, R. M.,
Model and Full Scale Predictions of a Carrier Flow
Field, NSWCCD-50-TR-2002/009, January 2002a.
Gorski, J. J., Haussling, H. J., Percival, A. S.,
Shaughnessy, J. J. and Buley, G. M., The Use of a
RANS Code in the Design and Analysis of a Naval
Combatant, Proc. 24th Symposium on Naval
Hydrodynamics, Fukuoka, Japan, July 2002b.
Krueger, K. E., Gorski, J. J., and Miller, R. W.,
Appendage
Alignment
Using
RANS-Based
Numerical Methods,
NSWCCD-50-TR-2002/035,
Nov. 2002.

Patel, V. C., Perspective: Flow at High Reynolds


Number and Over Rough Surfaces Achilles Heel of
CFD, Journal of Fluids Engineering, Vol. 120, 1998,
pp. 434 444.

Roache, P. J., Verification and Validation in


Computational Science and Engineering, Hermosa
Publishers, Albuquerque, New Mexico, 1998.
Stern, F., Longo, J., Penna, R., Olivieri, A., Ratcliffe,
T., and Coleman, H., International Collaboration on
Benchmark CFD Validation Data for Surface
Combatant DTMB Model 5415, Proc. 23rd
Symposium Naval Hydrodynamics, Val de Reuil,
France, 2000.
Taylor, L. K. and D. L. Whitfield, Unsteady ThreeDimensional Incompressible Euler and Navier-Stokes
Solver for Stationary and Dynamic Grids, AIAA
Paper No. 91-1650, June, 1991.
Taylor, L. K., A. Arabshahi, and D. L. Whitfield,
Unsteady Three-Dimensional Incompressible NavierStokes Computations for a Prolate Spheroid
Undergoing Time-Dependent Maneuvers,
AIAA
Paper No. 95-0313, Jan. 1995.
Van Leer, B., J. L. Thomas, P. L. Roe, and R. W.
Newsome,
A Comparison of Numerical Flux
Formulas for the Euler and Navier-Stokes Equations,
AIAA Paper No. 87-1104-CP, June 1987.
Whitfield, D. L. and L. K. Taylor, Discretized
Newton-Relaxation Solution of High Resolution FluxDifference Split Schemes, AIAA Paper No. 91-1539,
June 1991.

16
Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Stephane Cordier
Bassin dessais des Carnes, France
Reynolds number effects: You show large
differences in wake fraction due to Reynolds number
effect. Can these be correlated to powering data at
model and full-scale?
Shaft wake deficits: Your comparison of
wake deficit due to the shaft between computation
and experiments. Given that the shaft wake is mostly
due to hull boundary layer being entrained by the
longitudinal vortices along the shaft, can you point to
some explanations on these differences?
AUTHORS REPLY
No attempt was made in the present effort to
correlate the Reynolds number differences in wake
fraction with powering data from model to full-scale.
In fact, the difference is probably larger than would
be experienced by a real ship as no surface roughness
effects are included in the present full-scale
calculation.
For Model 5415 the shaft wake is well
captured in the computation from the comparisons
shown in Figs. 8, 10, and 16.
For the carrier
computation the shaft is included in the computation,
but the mounting struts are not.
Although the shaft
wake may be mostly due to the hull boundary layer
being entrained by the longitudinal vortices along the
shaft the strut wakes can accentuate the overall
shaft/strut wake. For the carrier this involves two
pair of struts for each shaft, which can accentuate the
wake deficit into the propeller considerably. The
authors feel the lack of mounting struts in the carrier
computation is the main reason for the discrepancies
between the computation and experiment.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics


St. Johns, Newfoundland and Labrador, CANADA, 813 August 2004

The Wakes of Idealized Propeller Shafts with Cross-flow and


Rotation
D. Hally (Defence R&D Canada Atlantic, Canada)

ABSTRACT

flat plates were calculated for different shaft angles, incident boundary layer thicknesses and Reynolds numbers
(Hally, 2001a). The numerical predictions were compared favourably with the experimental data of Pinard
(Hally, 2001b). An additional set of calculations added
an adverse pressure gradient (Hally, 2001c).
In none of the aforementioned calculations was
there any cross-flow across the shaft: i.e. the plate was
the xy plane, the flow was in the x direction, and the
shaft, initially lying along the x axis, was tilted up by rotating around the y axis by an angle : see Figures 1 and
2 for a description of the shaft geometry. The present paper describes a new series of calculations which include
flow across the shaft: i.e. the shaft is also rotated through
an angle about the z axis. The wake was calculated for
a shaft angle of 10 , Reynolds number of 105 based on
shaft diameter, and flow angles of = 0 , 2 , 4 , and 6 .
The effect of shaft rotation has also been examined. The rotation rate of the shaft is quantified by the
non-dimensional rotation parameter :

In order to understand the effect of cross-flow and shaft


rotation on the wake in the propeller disk, a series of calculations was performed on idealized propeller shafts:
rotating cylinders extruding from a flat plate. The angle
of inclination of the shaft was a constant 10 , but both
the angle of cross-flow and shaft rotation rate were varied.
Both cross-flow and shaft rotation cause the
wake to assume a bowed shape. A vortex is also generated soon after the shaft emerges from the plate; it
then detaches from the shaft and is convected by the free
stream. The vortex convects velocity deficit from the
plate boundary layer into the shaft wake. When there is
no rotation the vortex is weak, it is stronger when there
is rotation but no cross-flow, and it is strongest when
the cross-flow and rotation combine. In the latter case,
which is typical of a ship with propellers turning inward
over the top, the vortex can cause a significant augmentation of the wake fraction.

Rs
(1)
V
where is the shaft rotation rate, Rs is the shaft radius,
and V is the speed of inflow. Therefore is the speed of
the fluid at the the shaft surface relative to the speed of
inflow. For a ship the rotation parameter can be related
to the advance coefficient, J, by
=

INTRODUCTION
The flow into a propeller significantly affects the extent
to which it cavitates and consequently the noise radiated
by the ship. Extra velocity deficit in the propeller disk,
due either to the hull boundary layer or wakes from the
propeller shaft or appendages, tends to promote cavitation and increase radiated noise.
Pinard (1997) and Cordier et al. (1997) have
shown that there can be a significant interaction between
the wake of the propeller shaft and the hull boundary
layer of a twin screw ship resulting in an increase in
wake fraction. In an attempt to understand this interaction, the wakes of inclined cylinders protruding from

J=

Rs
Rp

(2)

where R p is the radius of the propeller. Typical values


for are in the range 0.5 to 1.0.

y
y
x

Incident flow

Figure 2: The shaft seen from directly below. The z


coordinate extrudes vertically from the page.

Figure 1: The shaft


1

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

1.2
1
H

0.8

Velocity

Figure 3: The skewed box containing the computational


region.

0.6
0.4
0.2

CALCULATION SET-UP

0
0

The flow region was a skewed box of length L in the x


direction, width W in the y direction, and height H in
the z direction: see Figure 3. The skew angle was equal
to the shaft inclination . For each grid L/D s = 230,
W /Ds = 150 and H/Ds = 25 where Ds is the shaft diameter. These dimensions are sufficient to keep the surface of the shaft a distance of at least 20Ds from the side
boundaries. The incident flow was parallel to the sides
of the box; the shaft was rotated through about the
vertical (z) axis to generate the cross-flow: see Figure 2.
The upper boundary of the box (z = 0) was
treated as a solid wall (the plate) and shaft surface as
a moving boundary with prescribed velocity set by its
rotation. Near the shaft/plate junction the rotation velocity was reduced to remove the discontinuity in velocity
between the shaft and the plate:
v=

v 
if z d
z
z
v
2
if z d
d
d

0.5

1
1.5
2
Distance to Wall

2.5

Figure 4: The velocity profile of the incident boundary


layer. The wall distance is in units of shaft diameter.

RESULTS
The nominal propeller disk
For the purpose of displaying results in a familiar way,
a nominal propeller disk was chosen at a distance of 20
diameters along the shaft from the shaft/plate junction.
The diameter of the disk is 5 shaft diameters and the
clearance of the edge of the disk from the plate is 1.01
shaft diameters (0.202 propeller diameters).
Coordinate systems
It will prove convenient to represent the velocity in two
different coordinate systems. In the propeller coordinates one axis is aligned with the axis of the shaft. The
component of the velocity in this direction is the axial
velocity. The tangential velocity is the velocity which
remains after the axial velocity has been removed.
In the plate coordinates one axis is aligned with
the flat plate and the other with the free stream flow. In
this coordinate system the tangential velocity is the velocity that remains after the component in the direction
of the free stream has been removed.

(3)

where v is the velocity due to the rotation of the shaft


and d was set to Rs .
The incident boundary layer was allowed to develop naturally along the plate. At the shaft plate junction its thickness was about 1.6 shaft diameters: see Figure 4.
The computer program TRANSOM (Hally,
1997), developed at DRDC Atlantic, was used for all calculations. It uses a flux-splitting pseudo-compressibility
scheme on structured grids following the formulation of
Rogers and Kwak (1988). The Spalart-Allmaras (1992)
turbulence model was used.
The grid was calculated in planes of constant z
(i.e. parallel to the flat plate). The planes were clustered
near the plate with closest node spacing of 3 10 4Ds ,
small enough to keep the y+ values close to 1.0 for the
closest nodes to the wall. A 680,225 node grid was used;
161 nodes were placed around the shaft, 65 along its
length between the lower and upper surfaces, and 65 radially outward from the shaft.

Inclined flow with no shaft rotation


Figure 5 shows the flow into the nominal propeller disk
at flow inclinations of = 0 , 2 , 4 and 6 . Two sets of
tangential velocities are shown: in the propeller coordinate system and in the plate coordinate system. The displayed flow region extends past the edge of the propeller
disk to 1.2 propeller radii. The edge of the propeller
disk is marked by the larger black circle, the surface of
the shaft by the smaller black circle.
The principal features of the axial wake are the
flat plate boundary layer at the top of the disk, and the
2

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Tangential
Propeller Coordinates

Axial

Tangential
Plate Coordinates

1.00
0.95
0.90
0.85

= 0

0.80
0.75
0.70
0.65
0.60
0.55
0.1 V

0.50
1.00
0.95
0.90
0.85

= 2

0.80
0.75
0.70
0.65
0.60
0.55
0.1 V

0.50
1.00
0.95
0.90
0.85

= 4

0.80
0.75
0.70
0.65
0.60
0.55
0.1 V

0.50
1.00
0.95
0.90
0.85

= 6

0.80
0.75
0.70
0.65
0.60
0.55
0.1 V

0.50

Figure 5: Flow into the nominal propeller disk at different cross-flow angles; no shaft rotation.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

shaft wake connecting the shaft at the centre of the disk


with the flat plate boundary layer. It is also possible to
discern two counter-rotating vortices in the shaft wake
just above the shaft. The tangential velocities in the plate
coordinates show clearly that in the shaft wake the fluid
moves downward toward the shaft. Velocity deficit is
convected downward into the shaft wake from the plate
boundary layer increasing the velocity deficit in the shaft
wake in the upper part of the propeller disk.
When = 0 , the shaft wake remains directly
above the shaft. As the flow angle is increased, there
is a cross-flow from left to right across the disk which
convects the shaft wake to the right. The portion of the
shaft wake in the upper part of the disk is created farther upstream than the portion near centre of the disk, so
the shaft wake is deflected further to the right as one proceeds outward from the centre. When the plate boundary
layer is approached, the convecting velocity begins to
diminish and the rightward deflection of the shaft wake
also begins to diminish. Therefore, when > 0, the shaft
wake is rotated clockwise around the shaft and is bowed
toward the centreplane.
The largest contribution to the tangential velocities in propeller coordinates is due to the projection of
the free stream onto the propeller disk. In the plate coordinate system the the free stream is entirely removed.
It should be noted that for most ships the angle
of flow across the shaft is largest where the shaft leaves
the hull and decreases as one proceeds along the shaft.
At the propeller disk the angle of cross-flow is typically
two-thirds to one half its value at the hull. However,
in the set-up reported here the cross-flow remains constant along the length of the shaft. Therefore, for a ship
wake, the portion of the tangential velocity field in propeller coordinates due to the projection of the free stream
would normally be smaller than that shown in Figure 5.
The tangential velocity fields in plate coordinates show that, as increases, a counterclockwise vortex forms near one oclock on the propeller disk. It tends
to augment the convection of velocity deficit from the
plate boundary layer into the shaft wake, as well as to
cause the boundary layer to become thinner near one
oclock on the propeller disk. The increase in wake fraction for 6 cross-flow relative to no cross-flow is about
30%.
The vortex first forms within a few shaft diameters of the shaft/plate junction. It is fully developed
about 10 shaft diameters downstream by which time
it has detached from the shaft and is convected downstream by the free stream. As one moves along the shaft
getting further from the plate, the vortex appears to drift
upward away from the shaft. By the time the propeller
disk is reached, the centre of the vortex is close to the
outer edge of the propeller disk.

1.00
0.95
0.90
0.85
0.80
0.75
0.70
0.65
0.60
0.55
0.50

0.1 V

= 0.2
1.00
0.95
0.90
0.85
0.80
0.75
0.70
0.65
0.60
0.55
0.50

0.1 V

= 0.4
1.00
0.95
0.90
0.85
0.80
0.75
0.70
0.65
0.60
0.55
0.50

0.1 V

= 0.6
1.00
0.95
0.90
0.85
0.80
0.75
0.70
0.65
0.60
0.55
0.50

0.1 V

= 0.8
1.00
0.95
0.90
0.85
0.80
0.75
0.70
0.65
0.60
0.55
0.50

0.1 V

= 1.0
Figure 6: Flow into the nominal propeller disk at different shaft rotation rates; no cross-flow.
4

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

which the dependence on rotation rate weakens. In the


cases shown here = 10 , so tan 0.176.
Figure 7 shows the variation with of the vorticity at the centre of the vortex. It confirms that the
strength of the vortex depends strongly on only for
<
0.2.
Although the vortex due to rotation tends to be
stronger than that caused by cross-flow, it is not as effective at convecting velocity deficit into the propeller disk.
This is because of its location, somewhat closer to the
shaft, and consequently further from the plate boundary
layer.

0.8
0.6

Vorticity

0.4
0.2
0
-0.2
-0.4
-0.6
-0.8
-1

-0.5

0
Omega

0.5

The combination of inclined flow and shaft rotation

Figure 7: Variation of the vorticity at the centre of the


vortex with shaft rotation rate: = 0.0.

Figure 8 shows the flow into the nominal propeller disk


when shaft rotation is combined with cross-flow. The
most striking feature of these wakes is the marked increase in velocity deficit when = 0.6 and = 6 .
In this case the direction of cross-flow and rotation are
such that each promotes generation of a counterclockwise vortex. The result is a stronger vortex which increases the rate at which velocity deficit is convected
away from the shaft and plate boundary layers. Some
of the velocity deficit is entrained by the vortex and retained after it detaches from the shaft. The ratio of wake
fractions for the cases with = 0.6 and = 0.6 is
roughly 23%.
Figure 9 tracks the evolution of the vortex for
the case with = 0.6. It shows the inflow into three
disks upstream of the propeller disk; the distance, d, of
each disk from the shaft/plate junction is 5, 10 and 15.
For the former, a close up of the region near the shaft is
shown. This figure shows that within five shaft diameters of the shaft/plate junction the vortex has formed and
the shaft boundary layer near it has begun to thicken.
When d = 10Ds , the vortex is well-developed but still
remains close to the shaft. A large bubble of velocity
deficit has already formed near the vortex centre. When
d = 15Ds , the vortex has detached from the shaft and has
moved upward and to the right taking the entrained velocity deficit with it. When the propeller disk is reached
(d = 20Ds ; see Figure 8 with = 0.6) the vortex has
reached the edge of the propeller disk. The bubble of
velocity deficit has diffused becoming broader but less
intense.

Shaft rotation with zero incidence flow


Figure 6 shows the flow into the nominal propeller disk
at shaft rotations of = 0.2, 0.4, 0.6, 0.8 and 1.0 (the
case with = 0.0 is shown in Figure 5). The tangential velocities are in the plate coordinate system. Arrows
at the shaft surface have not been drawn because they
would be very long at the higher rotations.
The main features of the flow field are similar
to those when there is cross-flow. The wake separation
points (here defined to be the points in the propeller disk
at the shaft surface where the circumferential shear stress
is zero) are shifted counterclockwise by the rotation. The
wake then leaves the shaft at an angle and curves upward
to meet the plate boundary layer. A vortex, rotating in
the sense opposite to that of the shaft, develops and detaches from the shaft. For typical rotation rates the vortex is stronger than that induced by cross-flow.
Once the rotation rate is sufficiently large, the
wake is only weakly dependent on except very close
to the shaft. This is because the velocity due to the rotation does not have time to diffuse very far from the shaft.
The main influence of the rotation is to shift the separation points for the wake. Once the rotation speed at the
shaft surface is equal to the cross-flow due to the shaft inclination, the separation points coalesce at a single point
at 9 oclock on the shaft surface. At higher rotation rates
there are no separation points; the circumferential shear
stress is non-zero all the way around the shaft. Instead,
near 9 oclock but a short distance off the surface, there
is a point where the tangential velocity in propeller coordinates is zero. Because the fluid rotation decreases very
rapidly away from the surface, this point varies very little with increasing rotation rate and the size and strength
of the vortex remains roughly constant. Therefore the
main portion of the wake changes until tan after

VARIATION WITH GRID DENSITY


To obtain an estimate of the dependence of the calculations on the grid density, the flow with = 6 and
= 0.6 was calculated on four different grids denoted
small, medium, big and very big (the standard grid used
was the big grid). The ratio of the sizes of the cells in the
5

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

1.00
0.95

1.00

0.90

0.95

0.85

0.90

0.80

0.85

0.75

0.80

0.70

0.75

0.65

0.70

0.60

0.65

0.55
0.50

0.60
0.1 V

0.55

= 0.6

0.50

0.1 V

d = 5Ds

1.00
0.95

1.00

0.90

0.95

0.85

0.90

0.80

0.85

0.75

0.80

0.70

0.75

0.65

0.70

0.60

0.65

0.55
0.50

0.60
0.1 V

0.55

= 0.0

0.50

0.1 V

d = 10Ds

1.00
0.95

1.00

0.90

0.95

0.85

0.90

0.80

0.85

0.75

0.80

0.70

0.75

0.65

0.70

0.60

0.65

0.55
0.50

0.60
0.1 V

0.55

= 0.6

0.50

0.1 V

d = 15Ds

Figure 8: Flow into the nominal propeller disk at different shaft rotation rates with cross-flow; = 6 .

Figure 9: The velocity field on three disks at different distances, d, from the shaft/plate junction: = 6 ,
= 0.6. The tangential velocities are in plate coordinates. The gray areas are above the plate outside the
region of flow.

small, medium and big


grid to the
cell sizes in the very
big grid are roughly 2 2, 2 and 2 respectively.
Figure 10 displays the difference between the
inflows calculated using the big and very big grids. A
positive difference indicates that the value was higher on
the big grid. The difference in the axial velocity is less
than 2% except in small regions close to the shaft. Most
of the discrepancy can be attributed to a small shift of the
very big grid shaft wake to the left of the big grid shaft
wake. The difference in the magnitude of the tangential
velocity is less that 1% of the free stream velocity except
in small regions close to the shaft.
Figure 11 plots the variation of the vorticity at
the centre of the vortex against relative cell size (i.e. a
cell size of 1 is the very big grid). A straight line least
squares fit to the four points has also been added. It is
clear that the vorticity at the centre of the vortex is sensitive to the density of the grid. Extrapolation back to

0.020

0.010

0.015
0.008
0.010
0.005

0.006

0.000
-0.005

0.004

-0.010
0.002
-0.015
-0.020

0.000

Figure 10: The difference between the velocity in the


nominal propeller disk as calculated on the big and verybig grids. The difference in axial velocity is on the left
and the difference in the magnitude of the tangential velocity in plate coordinates on the right.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

1.00

0.8

0.95
0.90

0.75

0.85

0.7

0.80

Vorticity

0.75

0.65

0.70
0.65

0.6

0.60
0.55

0.55

0.50

0.1 V

= 0.6

0.5
0.45

1.00

0.5

1
1.5
2
Relative cell size

2.5

0.95
0.90
0.85
0.80

Figure 11: Variation of the vorticity at the centre of the


vortex with relative cell size: = 0.6, = 0.6

0.75
0.70
0.65
0.60

a cell size of zero suggest that the vorticity predicted by


the big grid is roughly 20% to 25% too low. On the other
hand, the location of the vortex varied very little with the
grid density.
The discrepancies between the grids are small
enough to conclude that using a larger grid would not
alter the qualitative conclusions made from the results
of this study.

0.55
0.50

0.1 V

= 0.6
Figure 12: The flow in the nominal propeller disk for
= 6 and Re = 106 . The tangential velocity is in the
plate coordinate system.
Figure 12 also shows clearly that the shaft wake
in the inner part of the propeller disk is deeper when
= 0.6 than when = 0.6. This is also true at the
lower Reynolds number but cannot be seen so clearly in
Figure 8. Therefore, part of the effect of the vortex is
to redistribute the velocity deficit, removing it from the
shaft wake into the core of the vortex. The vortex then
carries its portion outward toward the edge of the disk
at the same time as it is diffused over a larger area. On
an inward turning propeller there will be relatively more
velocity deficit near the edge of the disk, but perhaps less
near the shaft. The redistribution can have repercussions
for the types of cavitation produced by the propeller, tip
vortex cavitation being most sensitive to velocity deficit
near the edge of the disk, while sheet cavitation is more
sensitive to velocity deficit somewhat closer to the shaft.

VARIATION WITH REYNOLDS NUMBER


To determine the effect of Reynolds number on the inflow, the calculations with = 6 and = 0.6 were
also done at a Reynolds number of 106 based on shaft
diameter. For these calculations a larger 161 77 77
grid was used with the distance of the closest nodes to
the wall reduced to 3105. (The effect of the Reynolds
number on the flow with no rotation or cross-flow has
been reported by Hally (2001a).)
The flow into the nominal propeller disk for
these two cases is shown in Figure 12; it can be compared with Figure 8. At both Reynolds numbers the
locations and strengths of the vortices are very similar,
lending credence to the supposition that the vortex is dependent primarily on the geometry and the location of
the separation points on the shaft. Once the rotation rate
is large enough that there is no separation, the character
of the vortex depends solely on geometry.
As should be expected, the velocity deficit in
the shaft wake is less at the higher Reynolds number. The velocity deficit entrained by the vortex is also
smaller, due to less velocity deficit being available in
the the plate and shaft boundary layers for entrainment.
Moreover, the relative difference in wake fraction between the = 0.6 and = 0.6 cases is only about
7% at the higher Reynolds number.

CONCLUDING REMARKS
The flow calculations shown in this paper suggest that
cross-flow and shaft rotation can have a significant effect on the propeller inflow by generating a vortex which
convects velocity deficit into the propeller disk from the
hull and shaft boundary layers. Both cross-flow and
shaft rotation can result in increased wake fractions. The
vortex is most intense when the cross-flow angle and rotation rate have different signs, the typical case for a ship
whose propellers turn inward over the top.
7

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Because the rotation rate for ship propellers


is typically in the range 0.5 to 1.0, where the strength
of the vortex is nearly constant for a given cross-flow,
the propeller inflow depends only weakly on the rotation rate. Therefore, only three rotational states need be
considered for a given cross-flow:

Hally, D., TRANSOM: A Multi-method Navier-Stokes


Solver: Overall Design, DREA TM 97-231, Defence
R&D Canada Atlantic, 1997
Rogers, S. E. and Kwak, D., An Upwind Differencing Scheme for the Incompressible Navier-Stokes Equations, Technical Report TN 101051, NASA, 1988.

1. the propeller turns outward over the top and the


shaft is bare;

Spalart, P. R. and Allmaras, S. R., A One-Equation Turbulence Model for Aerodynamic Flows, AIAA Paper
92-0439, 1992.

2. the shaft is protected by a sleeve so that the effective


rotation rate is zero; and
3. the propeller turns inward over the top and the shaft
is bare.
Since part of the effect of the vortex is to redistribute velocity defect in the propeller disk, and this
seems to be more important at higher Reynolds numbers, it is not possible to say definitively that propellers
that turn inward over the top will cavitate more or earlier
than propellers turning outward over the top. However,
it certainly seems that inward turning propellers are most
likely to have cavitation problems caused by the combination of cross-flow and shaft rotation.
Because the vortex is well-formed, and entrainment of velocity deficit begins, as soon as the shaft fully
clears the plate, it is likely that the vortex for inward
turning propellers could be weakened by covering the
shaft with a short sleeve extending only a little way past
the shaft/plate junction. A vortex would still form, but
would be closer in strength to that of the case with no
rotation. Flow calculations with short sleeves have not
yet been tried.

REFERENCES
Pinard, J-C., Etude experimentale et numerique du sillage en amont dune helice, Ph.D. thesis, Ecole Doctorale Sciences pour lIngenieur de Nantes, 1997.
Cordier, S., Legrand, F., and Pinard, J. C., Hull
and Shaft Wake Interaction, in Proceedings of Propeller
and Shafting 97, Arlington, USA, 1997.
Hally, D., On scaling the wake behind an inclined cylinder protruding from a flat plate, DREA ECR 2000-130,
Defence R&D Canada Atlantic, 2001a.
Hally, D., Flow past an inclined cylinder protruding
from a flat plate: comparison of numerical predictions
with experiment, DREA ECR 2000-168, Defence R&D
Canada Atlantic, 2001b.
Hally, D., On scaling the wake behind an inclined cylinder in an adverse pressure gradient, DREA ECR 2000131, Defence R&D Canada Atlantic, 2001c.
8

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Stephane Cordier
Bassin dessais des Carnes, France
Could you give further information on the
insight this model can give in terms of Reynolds
number and the presence of a shaft bossing on the
propeller wake?
AUTHORS REPLY
Although the calculations at different
Reynolds numbers have not yet been very extensive,
all those that have been performed suggest that the
principal effect of Reynolds number is to reduce the
velocity deficit in the wake; the strength and location
of the vortices vary very little with number.
Therefore, the principal conclusions, that the shaft
rotation is important and that the wakes from inward
and outward turning propellers may be significantly
different, are likely to hold at full scale as well as
model scale.
The model is too simple to shed much light
on the effect of the shaft bossing. Recent calculations
suggest that any discontinuity in rotation along the
shaft (i.e. passing from a portion of the shaft
protected by a sleeve or bossing to a portion which is
open to the flow) is sufficient to generate a vortex of
moderate strength. However, how efficiently the
vortex entrains velocity deficit from the shaft
boundary layer may well depend on the details of the
bossing geometry.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics


The
Symposium on Naval Hydrodynamics
St. Johnss, Newfoundland and Labrador, CANADA, 8-13 August 2004

Complementary RANS Equations for


Viscous Flow Computations
Kunho Kim, Ana I. Sirviente and Robert F. Beck
(University of Michigan)

ABSTRACT

Campana, et al. (1993) utilized a similar matching


idea between a RANS equations finite volume scheme
and a linear potential code. They chose a Dawson
model for the free-surface external flow. Chen and
Lee (1996) employed a nonlinear potential code combined with a RANS code using the finite analytic
method for a submerged foil with and without the
presence of free surface waves. An extension to a
surface piercing body was exploited in Chen and Lee
(1999).
Although it is clear from the point of view of
computational time that the potential/RANS matching
method is advantageous over full viscous RANS computations, this approach still involves a fairly large domain for the computation of the viscous effects. Thus
far, researchers have used a potential solver separately
either by using potential solutions as initial conditions
or by matching potential solutions to viscous solutions
in separate regions. However, little has been done to
directly couple the use of potential solutions with viscous solutions.
Recently, Dommermuth, et al. (1998) employed a
decomposition to solve the contact line problems in
bow waves. They decomposed the flow into an irrotational portion and a vortical portion. The vortical
portion was used to enforce the no-slip boundary condition on the hull and the irrotational portion was used
to impose the free-surface boundary conditions. Ferrant, et al. (2003) also proposed a potential/RANSE
(RANS Equations) formulation by using a decomposition of velocity and pressure into an incident part
and a diffracted part. In their method, the Euler equations are solved for the incident part and modified
RANS equations, so called, SWENS (Spectral Wave
Explicit Navier-Stokes) equations are solved for the
diffraction part in a submerged square cylinder problem with a free surface.
Kim, et al. (2003) developed a complementary set

In this study the complementary Reynolds-Averaged


Navier Stokes (RANS) equations developed by Kim
et al. (2003) are used to compute the laminar and turbulent flows around a two-dimensional NACA airfoil.
The impact of the order of accuracy of the discretization scheme used for the convective terms is assessed
for both laminar and turbulent flows. Furthermore the
overall effects of the potential solution on the numerical results are also analyzed and discussed.
INTRODUCTION
Numerical techniques involving the coupling of potential/RANS solvers have been proposed in the past
for ship hydrodynamic problems. Among others,
Stern, et al. (1988), Tahara, et al. (1992), Chen, et al.
(1993), Campana, et al. (1993), Chen and Lee (1996,
1999), Dommermuth, et al. (1998) and Ferrant, et al.
(2003) have all studied viscous-inviscid interactions.
Stern, et al. (1988) employed a displacement body
concept to solve the partially parabolic RANS equations. Similar techniques were utilized by Tahara, et
al. (1992) for the Wigley hull. They used the SPLASH
potential code along with a RANS equations solver
based on the finite analytic method. Although these
studies produced satisfactory results, the displacement
thickness is sensitive to small velocity changes in the
outer parts of the viscous layer. Moreover, the displacement thickness concept becomes questionable as
the boundary layer thickens or if flow separation takes
place. To overcome this problem, Chen, et al. (1993)
proposed a velocity/pressure matching scheme. They
solved the RANS equations based on the finite analytic method for the viscous near field, which was
matched with a potential solution, and computations
were performed for a Series 60 bare hull (  
  ).
1

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

mensionalized by the velocity scale   , the length


scale
, the reference pressure  , and the kinematic
viscosity .
The fundamental idea for the derivation is that the
velocity field can be decomposed into any arbitrary
two parts as follows:

of RANS equations based on a decomposition of the


velocity field into a viscous flow and a potential flow.
Such decomposition was used to develop a modified
set of RANS equations that solves for the complemen
tary velocity
defined as the difference between the
velocity at a given point minus the potential velocity.

This decomposition is expected to produce a velocity distribution, such that it will approach zero in the
far field quickly assuming the potential velocity was
chosen properly. This could possibly minimize the
size of the computational domain, which will translate into obvious computational time savings.
The equations were validated by the authors for
three different laminar flows, and the results show
good agreement with experimental data and simulations from conventional RANS equations. It was
shown that the complementary RANS formulation
gave comparable results using coarser grids and a first
order discretization upwind scheme for the convection
terms to the numerical solutions of the conventional
RANS equations using finer grids.
In this study, the complementary set of RANS
equations developed by Kim, et al. (2003) for laminar
flows will be applied to turbulent flows. Moreover,
the results of numerical experiments set up to further
investigate the advantages and disadvantages of such
numerical approaches versus the conventional RANS
equations are also discussed. To that end, the effects
of the chosen potential solution, as well as the order of accuracy of the discretization scheme used for
the convection terms, are evaluated in terms of overall
performance of the complementary RANS solver are
evaluated for both laminar and turbulent flows.
To aid in the understanding of the complementary
RANS equations, an overview of the mathematical
formulation supporting the derivation of the complementary RANS equations and the numerical methods
is presented in the following section. This is followed
by a section giving results for the two-dimensional
laminar and turbulent flows about NACA airfoils with
discussions on the effect of the discretization schemes
and the employed potential solutions.


  

(1)

where is the complementary velocity to the potential velocity vector  and 


 which corresponds to  , , and  direction, respectively. For inviscid, irrotational flows, the potential velocity vector,
 , can be expressed as the gradient of a velocity potential, ! , where ! satisfies the Laplace equation. The
velocity vector, " , must satisfy the continuity equation and the Navier-Stokes equations for incompressible viscous flows. By substituting the velocity decomposition shown in Eq. (1) in the continuity and the
Navier-Stokes equations and after some mathematical
manipulation, the following set of equations results:

# 
#


(2)

#
 
('  ' ) # '
 ' 



 3

 *,+%-/.0  12 . 54 (3)


* +
,
1%2
where

7689:
<;>= is the Reynolds number,

3 '
9(;@? A
is the Froude number, and is the Kro
#%$ 

#  &
#%$  ( ' 

 #
' ) #
#

#
%

&

necker delta. Eq. (3) includes the influence from the


potential solution in the form of two new convection
terms in the left hand side. The right hand side has the
same form as the original Navier-Stokes equations.
Using the Bernoulli equation for the potential
flow, Eq. (3) can be further simplified as follows:

#
# 

' #  '  :' # '


# 


#  *, +%-/.0



where the new pressure field, , is defined as:



B  
# 
#$ 

COMPUTATIONAL METHODS
The complementary RANS equations were derived in
Kim et al. (2003). In this section a summary of the
derivation of the equations is presented and the corre
model used for the turbulence
sponding modified
computations is discussed. All variables are nondi-

 # 
' # ' 


(4)

(5)

The set of Eqns. (4) constitute the complementary


Navier-Stokes equations.
Comparing Eq. (4) to the original momentum
equations, it can be observed that the new equations
2

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

have instead of % with two additional terms (boxed


terms) due to the introduction of the velocity potential. Both terms are convection terms. The first one,
#
#
'  ;  ' , corresponds
to the convection of 
#  # '

with the speed of and the second one,  '
;  ,

corresponds to the convection of with the speed of
 .

The complementary velocity can be decomposed into a mean part and a fluctuating part just as
 , i.e.


 

and

  

  

'

# 
#%$ 
# 
#$ 

# '
  3 '

# '  #


% 


#
# 
# 
 

'
# ' 
# '  ( ' # '



# 
 # 
#
'

#

#
#




 '
 '
 


* +  

(11)

# '


(12a)

#  # 

# '  # '  
   

 


.
  
(12b)

# 

' #
 '

# 

(7)




 # '
 
B



.

(13)

and  are defined as:






* +  



* +  


(14a)
(14b)

 is the production term and it is defined as:


 #
# ' #
#

%
%

' #  '
 #  '  #   #  '  (15)


and   and   are the effective Prandtl numbers,




which represent the diffusion rate of and .    ,
  and  are empirical constants.

(8)

Following a similar procedure to that explained


for the derivation of the complementary RANS equa  
tions, the corresponding modified
model is derived as:

(9)

where is the turbulent kinetic energy,  is the eddy


viscosity, which is dependent on flow conditions, 
is the effective kinematic viscosity defined as:

where is the turbulent energy dissipation rate and 


is the eddy viscosity, which is modeled as

the following form of the complementary RANS


equation can then be obtained:

# 

#$  '
# 

#
 


# 
' #
 '

(6)

 #



  


9/

The conventional
turbulence closure of Chen
and Patel (1985) is chosen to model turbulence in this
study, that is

It can be seen that the mean velocity, , becomes



  and the fluctuating parts are the same in both
decompositions.
The rest of the derivation of the complementary
RANS equations is the same as that followed for the
conventional RANS equations. That is, the complementary RANS equations can be obtained by substituting the decomposition shown in Eq. (7) into the
complementary Navier-Stokes equations and taking
the time average in the conventional sense, and then
using the eddy viscosity model, following the Boussinesq assumption, to model the Reynolds stresses:

is defined as:

 

It should be noted that the fluctuating part is denoted



as , not . This means that the fluctuating part of

is equal to that of " . This is clear if the decomposition is written as follows:

 

(10)

# 
 # 
'
#%$ 
# ' 

#  # 

#
 # '
 '
 
# 
#


#$  ' # ' 

#  # 

# '  # '     

 

3
Copyright National Academy of Sciences. All rights reserved.

# 

 ' # '

 

# 
 ' # '



.
  

(16a)

(16b)

Twenty-Fifth Symposium on Naval Hydrodynamics

In this study wall functions are employed to avoid


the integration of the equations through the sublayer.
The wall function method has been widely used due
to its two merits: it economizes computer time and
storage, and it allows the introduction of additional
empirical information in special cases such as when
the wall is rough (Launder and Spalding, 1974). The
  
model assumes that the log-law holds in the
inner layer, which is governed by the wall function.
The conventional wall function formulation derived
by Chen and Patel (1988) is used herein.
The potential flow around the body is considered
to be steady, inviscid and irrotational. The flow velocity potential can then be expressed as:

!
 


of the convection terms eliminates the need for explicitly adding artificial dissipation terms to the right hand
side of the momentum equations to stabilize the numerical algorithm. Following the recommendations of
Kim et al. (2003), the second additional term is combined with the convection term of the conventional
RANS equations for all the simulations presented in
this study.
An implicit fractional-step method (Chorin, 1968)
is used to integrate the RANS equations in time to
steady state. The first step in this implicit two-step
procedure, involves the implicit discretization in time
of the convection and diffusion terms in the momentum equations while the pressure gradient term is evaluated explicitly using the available pressure field at
that time step. The solution of such discretized equations obtained by using an approximate factorization
method renders an intermediate velocity field that is
not divergence free. Consequently, to satisfy the compressibility constraint, a pressure correction is computed in the second step, the continuity step, by solving a Poisson-type equation whose source is the nonzero divergence of the intermediate velocity field. The
pressure correction obtained from the continuity step
is then used to update the pressure and to correct the
intermediate velocity field computed in the first step
of the fractional-step method. The resulting velocity field is now divergence free. The solution of the
equation resultant in the second step is obtained by
also using the same approximate factorization method
(Beam and Warming, 1978) used in the first step.
In the approximate factorization method, the original
multi-dimensional difference equations are replaced
by a series of one-dimensional difference equations,
which can then be formulated by a tri-diagonal matrix equation. The method allows a speedy solution
of relatively complex systems of differential equations
(Hoffmann and Chiang, 1993).

(17)

where is a perturbation potential. The potential velocity is derived by following a similar methodology
to that described in Kim et al. (2003). A Rankine
source type desingularized method (Cao, et al. 1991,
Beck, 1994) is employed for the potential solver with
sources located inside the two-dimensional body. Unlike the traditional panel method, the kernel function
is not singular when desingularized sources are used.
The general formulation of the potential velocity in
two-dimensional flow is as follows:





.




  ' 
 '
 
 . '
'   
    ' 
  '
 . '
'   

(18)

(19)

where  and  are the components


of the po&
.
tential velocity at any field point  ,    ) are the
coordinates
& of the position vector at the field point
 , and  '   ' ) are the coordinates of the position
vector
at a source point .
The continuity and momentum equations are
transformed into the generalized curvilinear coordinates and then discretized on a non-staggered mesh.
The continuity equation is discretized in space using
three-point central finite differencing. The same discretization algorithm is used in the momentum equations for the pressure gradient and the diffusion terms.
Both first and second order accurate flux splitting
based upwind differencings are tried in this study for
the convection terms. As noted earlier the two extra
terms that appear in the complementary RANS equations are convection terms. The upwind differencing

 




The complementary RANS solver is employed


for the numerical computations for two-dimensional
flows around a NACA airfoil. Since the solver is
coded in three-dimensions, a unit width of the domain
is considered in the computations.
RESULTS AND DISCUSSION
Kim et al. (2003) presented in a previous study the
derivation of the complementary RANS equations
along with a validation of the methodology for three
steady laminar flows: the flow in a square duct, the
4

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

The boundary conditions for the conventional RANS


solver are free-stream conditions at both the inlet and
outer boundaries, a symmetry boundary condition at
the centerline, upstream and downstream of the foil,
a no-slip condition on the airfoil and a Neumann condition at the exit. The corresponding boundary conditions for the complementary RANS solver are uniform


flow distribution at the inlet (i.e.
  ), no-slip


at the wall (i.e.

 ),# zero
 # streamwise velocity gradients at the exit (i.e.
; 
), recovery
of the free-stream velocity at the outer boundary (i.e.



  ) and symmetry boundary conditions at
the centerline, upstream and downstream the foil (i.e.
#  #
; 
). The numerical results presented herein
were obtained for a CFL number of 1.65 and the code
was run until the residuals, defined by the summation of differences between the current and the previous iterations, were reduced by at least four orders of
magnitude. In these calculations a body-fitted potential, obtained by applying the desingularized method
of Beck (1994), was used and compared with the results obtained from employing a uniform potential solution. The body-fitted potential had singularities inside the foil. The simulations presented in this study
were performed with 359 sources and a wake source
(Kim, et al. 2003, Kim, 2004) was used in the computations to eliminate the stagnation point at the trailing
edge.

flow over a flat plate and the flow over a NACA 0010
foil. In this study a follow up of the latter investigation
is performed with the intent of assessing the impact of
the chosen potential on the overall performance of the
complementary RANS equations solver. The ability
of the new equations to produce comparable results
to the conventional RANS equations but in coarser
grids will be explored to assess the performance of
first and second order discretization schemes for the
convection terms in the equations. Finally the solver
is used to solve the turbulent flow over a NACA airfoil
to identify the advantages or disadvantages of the new
set of equations when dealing with turbulence.

Effect of the Discretization Order of the Convective


Terms
The study of Kim et al. (2003) concluded that the
complementary RANS solver shows less grid dependency, consequently, it gives comparable solutions to
the conventional RANS code in a less resolved grid.
These computations involved a first order accurate
flux-splitting based upwind differencing of the convection terms. In what follows a discussion is presented where the applicability of this argument is tied
with the discretization order of the convection terms.
Computations of flow around the foil were con*,+

. The comducted for a Reynolds number,


putational domain, shown in Figure 1, starts two foil
lengths upstream of the leading edge of the airfoil,
and ends five body lengths downstream of the trailing edge. The height of the domain is two and a half
body lengths. The grid shown in Figure 1 is 

nodes with a stretching ratio of about 1.0 over almost
the entire grid, except for the vicinity of the trailing edge where it becomes about double. Two other
8 (Fine)) were
grids (   8 (Medium) and  8
used to assess the impact of the grid resolution on the
solver. A coarser grid than 
 (Coarse),   8
(Very Coarse) was also used in some instances.





Figure 2 (a) and (b) show comparisons of the


streamwise velocity components at 
 computed by the complementary RANS solver for both
the uniform and the body-fitted potential respectively,
for the coarse, medium and fine grids considered in
this study. Both figures show the results corresponding to both the first and second order discretization
schemes. The effect of the accuracy of the discretization scheme as a function of the chosen potential solution is evident from the figures. Using a body-fitted
potential (Figure 2 (b)), it is really does not matter
whether the first order or second order scheme is used.



 







Comparisons between the conventional RANS


solver and the complementary RANS solver with the
body-fitted potential are presented in Figures 3 (a) and
(b). Figure 3 (a) shows comparisons of the streamwise

velocity components and   , at 
 , with a
first order discretization scheme used for the convective terms for the three considered grids. Figure 3 (b)
shows similar results for the second order discretization scheme. The comparisons of both figures clearly
show that the performance of the first order scheme

3
2.5
2



1.5
1
0.5
0
-2

-1

upstream

foil

downstream



Figure 1: Computational domain and grid for


NACA 0010 airfoil ( 
 )
5

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

u*:Coarse:First
u*:Medium:First
u*:Fine :First
u*:Coarse:Second
u*:Medium:Second
u*:Fine :Second

0.4

0.3

0.3

0.2

0.2

0.4

0.1

0
1

0.1

1.05
1.1
u*+u p

0
1

1.15

(a) with Uniform

Figure 2:



-profile at 

u :Coarse:First
u :Medium:First
u :Fine :First
u*:Coarse:First
u*:Medium:First
u*:Fine :First

(NACA 0010,

1.15

 )

* +

u :Coarse:Second
u :Medium:Second
u :Fine :Second
u*:Coarse:Second
u*:Medium:Second
u*:Fine :Second

0.4

0.3

0.3

0.2

0.2

1.05
1.1
u*+u p
(b) with Body-fitted

 

0.4

0.1

0
1

u*:Coarse:First
u*:Medium:First
u*:Fine :First
u*:Coarse:Second
u*:Medium:Second
u*:Fine :Second

0.1

1.05
1.1
u and u*+u p

0
1

1.15

1.05
1.1
u and u*+u p

(a)  upwind

Figure 3: - and

 

(b) 

(with Body-fitted ! )-profiles at 



upwind

(NACA 0010,

6
Copyright National Academy of Sciences. All rights reserved.

1.15

*,+

 )

Twenty-Fifth Symposium on Naval Hydrodynamics

shown in Table 1 that the RMS difference in


 
is well within 1% even for the coarse grid. The results
for the very coarse grid are also quite impressive con
sidering that the RMS difference in
  is over
three times better than that shown by the conventional
RANS solver for the first order scheme. The RMS

differences in   for both discretization schemes
are below 2% while that is not true for the conventional RANS solver.

used in the complementary RANS equations solver


gives remarkably accurate results, for all grids, when
compared to those from the second order scheme.
On the other hand the results from the conventional
RANS solver show significant differences among the
solutions for the various grids with the first order
scheme. The results for even the finest grid are in substantial overall disagreement with respect to those obtained for the second order scheme. To further emphasize the similarities and differences a comparison of
the root-mean-square (RMS) difference between the
first order and the second order cases for the conventional and the complementary RANS solvers are presented in Table 1. The RMS difference is defined as:

*  &
)


&

 


) . ;




Impact of the chosen potential solution


Comparisons of the results corresponding to the
complementary RANS computations around the foil
for uniform and body-fitted potentials are shown in
Figures 4 (a) and (b) for the first and second order
discretization schemes of the convective terms. Each
figure shows the results corresponding to both potentials in the coarse, medium, and fine grids considered
in this study. The effect of the chosen potential on
the ability of the numerical solver to produce an accurate solution in a less resolved grid is clearly visible in the figures. The better potential solution of the
two used, that is the body-fitted potential, aids to produce smaller discrepancies between the grids considered when compared to the results from the uniform
potential solution.
The RMS differences in the horizontal velocity
component between the coarse and fine grids are
shown in Figure 5. This figure reveals how the better potential flow definition can influence the performance of the complementary RANS solver. In particular, the differences are large at the foils trailing edge
for the computations conducted with the uniform potential solver, while they are much smaller with the
body-fitted potential. Table 2 shows the summary of
the overall RMS differences for all grids considered
along with the CPU times of the calculations. The
CPU times for the very coarse and fine grids are not
presented because the number of nodes is not appropriate for the comparison. It is obvious that the coarse
grid with the body-fitted potential gives similar, if not
better, results than the medium grid with the uniform
potential. The savings achieved by reducing the number of nodes for the complementary RANS computations easily exceed the computational time expense
caused by computing the body-fitted potential.
The analysis of the results from the complementary RANS equations for the very coarse grid showed
that important oscillations of the streamwise complementary velocity, and consequently pressure, were

(20)

where  and  are velocities from two different


computations and  is the number of nodes.

Table 1 RMS difference (%) comparison between



the u- and the u*-codes w/  and  schemes

 

 

uw.  uw.  uw.  uw.


V.Co. vs. Fi.
6.56
2.46
1.93
1.50
Co. vs. Fi.
3.70
0.57
0.80
0.42
Me. vs. Fi.
1.28
0.09
0.31
0.13
V.Co. = Very Coarse, Co. = Coarse, Me. = Medium,
Fi. = Fine; uw = upwind scheme
Table 1 includes results for the very coarse grid, as
well, for reference. The root-mean-square difference
in displays significant improvement when the second order scheme is used. On the other hand the root

mean square difference in
  shows very similar results for both discretization schemes. In fact,
the difference between the coarse and fine grids with
the first order upwind scheme for the complementary
RANS equations is already below 1%. A higher order
discretization scheme for the convection terms does
not seem to substantially increase the already good accuracy shown by the first order scheme. As mentioned
in Kim et al. (2003), the results corresponding to the
first order scheme are attributed to the smaller velocity gradients computed by the complementary RANS
solver. It is shown herein that such an advantage can
be compared to the benefits obtained by increasing
the accuracy of the discretization scheme for the convection terms in the conventional RANS solver. It is
7

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

u*:Coarse:BF
u*:Medium:BF
u*:Fine :BF
u*:Coarse:UF
u*:Medium:UF
u*:Fine :UF

0.4

0.3

0.3

0.2

0.2

0.4

0.1

0.1

0
1

1.05
1.1
u*+u p

0
1

1.15

(a)   upwind

 )

Figure 4:

*,+

u*:Coarse:BF
u*:Medium:BF
u*:Fine :BF
u*:Coarse:UF
u*:Medium:UF
u*:Fine :UF

-profiles at 



1.05
1.1
u*+u p
(b)  

1.15

upwind

for both uniform potential (UF) and body-fitted potential (BF) (NACA 0010,

RMS difference in u (%)

Table 2 RMS difference (%) in


  and
CPU time (hour) for the u*-code with body-fitted
and uniform ! s
Body-fitted !
Uniform !
  uw.  uw.   uw.  uw.
RMS difference (%)
Co. vs. Fi.
0.80
0.42
3.70
0.41
Me. vs. Fi.
0.31
0.13
1.28
0.07
CPU time (hour)
Co.
0.46
0.49
0.46
0.48
Me.
5.86
6.07
6.09
6.30

with Uniform
with Body-fitted

0
0

To accomplish a smaller-gradient potential solution, the location of the control and source points was
changed in the trailing edge region as shown in Figure 7 to obtain a modified potential solution at the
trailing edge. The purpose was to have a fairer flow
around the trailing edge, and to that end a circular arc
was used for the potential solver, the size of which was
controlled by the length of the trailing edge to the end
of the arc. The results obtained following this more
careful selection of the potential solution are shown
in Figure 8. In the figure, a comparison of the distribution of  , when the wake source is used, with the
results from the modified potential with a faired trailing edge (faired-TE potential), are presented. Figure 9
shows the distribution of the complementary velocity
at the foils trailing edge and along the centerline for

 


Figure 5: RMS difference in the horizontal velocities


between the coarse grid ( 
* +  ) and the fine grid
8 ) (NACA 0010,

)
(  8

found along the centerline in the neighborhood of the



foils trailing edge. Results for the
velocity in the
four grids are presented in Figure 6. Those results
were obtained with a second order upwind scheme
for the convection terms. While insufficient grid resolution or inadequate discretization scheme can be
thought to be the main reason for the oscillations
shown in Figure 6, it was speculated that choosing
a potential solution with smaller velocity gradients at
the trailing edge could minimize, if not overcome, this
problem.
8

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

u*-code:
u*-code:
u*-code:
u*-code:

-0.75

Very Coarse[36x21]
Coarse[71x41]
Medium[141x81]
Fine[281x161]

1.1
w/ wake source
w/ faired-TE

-0.8

up

u*

-0.85

-0.9

0.9

-0.95

0.8
-1
0.8

0.9

1.1

1.2

x
0.8

1.1

1.2



Figure 6:
in the neighborhood of the trailing edge
*,+

)
and along the centerline (NACA 0010,

Figure 8:  in the neighborhood of the foils trail* +


ing edge and along the centerline (NACA 0010,



the faired-TE potential and all grids considered in this


study. It is evident that a much smoother distribution
is obtained by comparison to that shown in Figure 6
with the wake source. The RMS comparisons for the
various grids obtained from the u*-code with the wake
source and the faired-TE potential are shown in Table 3. As shown, the RMS differences decrease by
half in the u*-code with the faired-TE potential for the
very coarse grid compared to the corresponding case
with the wake source. However, in the coarse and the
medium grids, the RMS differences become slightly
larger than the wake source case. In other words, as
the grid becomes more resolved, the benefits from the
faired-TE potential decrease. Both cases give similar
CPU times, while the faired-TE potential case gives
slightly smaller CPU times due to the decrease of the
number of source points.
x

0.9

-0.8

u*

-0.85

-0.9

-0.95

-1
0.8

0.98

1.2

x x

Table 3 RMS difference (%) and CPU time (hour)


for the u*-code with wake source and faired-TE
! w/ wake source ! w/ faired-TE
RMS difference (%)
V.Co. vs. Fi.
2.46
1.25
Co. vs. Fi.
0.41
0.58
Me. vs. Fi.
0.07
0.12
CPU time (hour)
Co.
0.48
0.47
Me.
6.30
6.19

L tail
0.96

1.1



xx
x

Figure 9:
in the neighborhood of the foils trailing edge and along the centerline with the faired-TE
*,+

)
potential (NACA 0010,

x
x

0.9

source points
control points

u*-code:Very Coarse[36x21]
u*-code:Coarse[71x41]
u*-code:Medium[141x81]
u*-code:Fine[281x161]

-0.75

1.02

Figure 7: Distribution of control points and source


points near the trailing edge in the faired-TE potential
9

Copyright National Academy of Sciences. All rights reserved.

10

RMS difference in u (%)

RMS difference in u (%)

Twenty-Fifth Symposium on Naval Hydrodynamics

Very Coarse vs. Fine


Coarse
vs. Fine
Medium
vs. Fine

8
6
4
2
0
0

10
8
6
4
2
0

(a) u-code

(b) u*-code with wake source

10
Very Coarse vs. Fine
Coarse
vs. Fine
Medium
vs. Fine

8
6

CPU time (hour)

RMS difference in u (%)

Very Coarse vs. Fine


Coarse
vs. Fine
Medium
vs. Fine

2
0

5
4
3
2
1

VCoarse Coarse Medium VCoarse Coarse Medium VCoarse Coarse Medium


<--------- u-code ---------> <- u*-code w/ wake source -> <- u*-code w/ mod. potential->

(c) u*-code with faired-TE

(d) CPU times

Figure 10: Comparison of RMS differences and CPU times (NACA 0010,

10
Copyright National Academy of Sciences. All rights reserved.

*,+

 )

Twenty-Fifth Symposium on Naval Hydrodynamics

0.3

u*-code w/ wake source


u*-code w/ faired-TE

0.1

0.1

-C p

0.2

-C p

0.2

-0.1

-0.1

-0.2

-0.2

0.8

0.9

1.1

u*-code w/ wake source


u*-code w/ faired-TE

0.3

1.2

0.8

0.9

1.1

1.2

(a) Very Coarse

(b) Medium



Figure 11: Pressure coefficient distribution along the foils trailing edge and the centerline (NACA 0010,
*,+

)
Figures 10 (a)-(c) show the overall RMS differences corresponding to the four grids for the
conventional RANS solver (10 (a)) and the complementary RANS solver with the body-fitted potential
(10 (b)) and the faired-TE potential (10 (c)). The corresponding CPU times are presented in Figure 10 (d).
The time savings and accuracy obtained with the very
coarse grid with the faired-TE potential (1.65%) is remarkable by comparison to the results for the same
grid with the uniform potential (6.56%).
A comparison of the pressure coefficient distribution along the computational domain centerline and
over the body is shown in Figures 11 (a) and (b) for
the very coarse and medium grids respectively. The
results presented in each of the figures correspond to
both the original and faired-TE potentials. As shown
in Figure 11 (a) for the very coarse grid, the u*-code
with the faired-TE potential gives smaller fluctuations
than that with the wake source, especially after the
trailing edge. It is more obvious in the medium grid
shown in Figure 11 (b).

was used along with wall functions. The computational


 grids used for this study had the first node at
a  (
  ; , where  is the wall-friction velocity and is the kinematic viscosity of the fluid) of
about 150 from the wall. The number of nodes for the
coarse, medium, and fine grids was the same as those
used for the laminar computations.
Comparison of the computed   with the experimental data of Gregory et al. (1970) and the numerical results of Rhie and Chow (1983) are presented in
Figure 12. The results from the u*-code show good
agreement with both the experimental and numerical
data, while the pressure is estimated slightly higher
than the measured data over the airfoil.
1

Gregory et al. (1970)


Rhie and Chow (1983)
u*-code w/ faired-TE

-C p

0.5

-0.5

Turbulent flow computations


The complementary RANS equations were applied to solve the turbulent flow around the
*,+
NACA 0012 airfoil at a Reynolds number,


  . The foil choice was driven by the availability of experimental and numerical data that could
be used for comparisons. As mentioned earlier in the
  
Numerical Methods Section, a modified
model

-1
0




0.25

0.5

0.75

Figure 12: Pressure coefficient, comparison of


the u*-code with experimental and numerical data
* +
(NACA 0012,


 )




11
Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

u
u
u
u
u
u

0.3

0.3

0.2

0.2

0.1

0
1

0.1

1.05

1.1

0
1

1.15

u
(a) u-code




u*:Coarse:First
u*:Medium:First
u*:Fine :First
u*:Coarse:Second
u*:Medium:Second
u*:Fine :Second

0.4

0.4

:Coarse:First
:Medium:First
:Fine :First
:Coarse:Second
:Medium:Second
:Fine :Second

Figure 13: Comparison of


*,+


  )

1.05
1.1
u*+u p

1.15

(b) u*-code

in the conventional and the complementary RANS solvers at 



(NACA 0012,

Table 4 RMS difference (%) and CPU time (hour)


  upwind
 upwind

A similar analysis to that conducted for the laminar flow was done in this case to assess the impact
of the order of accuracy of the discretization scheme
used for the convection terms and also of the chosen
potential solution. Figures 13 (a) and (b) show the
horizontal velocity profiles, at 
 , corresponding to the conventional and the complementary RANS
solvers respectively. Each figure displays the results
for both first and second order accurate discretization
schemes for the coarse, medium and fine grids with
the faired-TE potential. The results reveal that very
similar conclusions to those drawn for the laminar
flow studies can be establish for this case. The conventional RANS solver shows very large discrepancies among the various grids for the first order scheme,
and smaller but still significant errors with the second
order scheme. On the other hand the agreement is excellent even with the first order upwind scheme for the
results from the complementary RANS solver. The
data not only shows good agreement between the first
and second order schemes but also between the various grids, and especially so compared to the results
from the conventional solver.

RMS difference (%)


Co. vs. Fi.
4.19
Me. vs. Fi. 2.75
CPU time (hour)
Co.
1.63
Me.
20.78



 

0.59
0.32

0.60
0.50

0.50
0.33

1.69
21.88

1.65
19.38

1.72
20.50

Comparisons with the results from the laminar flow


computation (even though the geometry is not exactly same, i.e. NACA 0010) reveal that the RMS differences seem to increase in the turbulent results (especially, in the conventional RANS solver). The RMS
differences found between the medium and fine grids
for the turbulent case (0.5%) are more than five times
larger than those found for the laminar case (0.09%).
The use of the second order scheme for the complementary RANS solver is not justified in this turbulent computations in comparison to the conventional
RANS solver according to the results in Table 4. Using the first order scheme, however, the complementary RANS solver gives significantly smaller RMS
differences compared to the corresponding conventional RANS solver. In fact, the results from the complementary solver with the first order scheme are better than those obtained from the conventional RANS

The overall RMS differences in and


  are
shown in Table 4 along with the corresponding CPU
times. The results indicate that substantial benefits
can be obtained with the conventional RANS solver
by using the second order discretization scheme.
12

Copyright National Academy of Sciences. All rights reserved.

RMS difference in u (%)

Twenty-Fifth Symposium on Naval Hydrodynamics

grid for the conventional RANS solver. For the turbulence computations the results from the first order
scheme show less error than those found for the second order scheme of the conventional RANS solver.
The impact of the chosen potential flow solution was
also investigated in this study. It was concluded that
the closer the potential solution is to the real fluid flow,
the more beneficial the complementary RANS equations will be. Because of the smaller gradients of the

solution, a less resolved grid can be used.

with Uniform
with faired-TE

0
0

ACKNOWLEDGEMENT
This work was supported by grants from the Office
of Naval Research (ONR). The computations performed in this study were done mostly at the Computational Marine Mechanics Laboratory of the University of Michigan (funded by ONR and DARPA).
Additional support was supplied by the U.S. Department of Defense High Performance Computing Modernization Program (HPCMP). This support is gratefully acknowledged.

 

Figure 14: Comparison of RMS differences in horizontal velocities between the coarse grid ( 
*,+  )
8 ) (NACA 0012,

and the fine grid (  8



 )




solver with the second order scheme.


The main reason behind this advantage of the
complementary RANS equations is the fact that the
gradients of the velocities are larger than those cor
responding to . This was discussed by Kim et al.
(2003) for laminar flows, from which the u*-code is
expected to be even more beneficial in turbulent flows
because the -velocity gradients will be even larger.
Very similar conclusions to those established in the
previous section regarding the impact of the chosen
potential, were concluded for the turbulent flow computations. Figure 14 shows the RMS differences for
the horizontal velocity component for computations
performed with both the uniform and body-fitted potentials. The body-fitted potential results render about
a fourth of the error found for the uniform potential
solution.

REFERENCES
Beam, R.M., and Warming, R.F., An Implicit Factored Scheme for the Compressible Navier-Stokes
Equations, AIAA Journal, Vol.16, No.4, 1978,
pp.393-402.
Beck, R.F., Time-domain computations for floating
bodies, Applied Ocean Research, Vol. 16, 1994, pp.
267-282.
Campana, E., Di Mascio, A., Esposito, P.G., and Lalli,
F., Domain Decomposition in Free Surface Viscous
Flows, Proceedings, 6th International Conference on
Numerical Ship Hydrodynamics, Iowa City, U.S.A.,
1993, pp. 329-340.

CONCLUSIONS
The complementary RANS solver was shown to give
less grid dependent results not only for laminar but
also turbulent flows. In fact, due to the larger velocity gradients present in turbulent flows, the benefits of the complementary RANS solver are possibly
larger. The ability of the solver to give equally accurate results in less resolved grids than the conventional
RANS solver was contrasted with the accuracy of the
discretization scheme used for the convection terms.
It was shown that for the laminar results the first order
upwind scheme can give as good results in the coarse
grid as a second order scheme will give in the medium

Cao, Y., Schultz, W.W., and Beck, R.F., A ThreeDimensional Desingularized Boundary Integral
Method for Potential Problems, International Journal for Numerical Methods in Engineering, Vol. 11,
1991, pp. 785-803.
Chen, H.C., and Patel, V.C., Calculation of TrailingEdge, Stern and Wake Flows by a Time-Marching
Solution of the Partially-Parabolic Equations, IIHR
13

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Including the Effects of Upper Surface Roughness


Simulation Hoarfrost, National Physical Laboratory,
Teddington, England, Aero Rept. 1308, 1970.

Report No.285, April 1985.


Chen, H.C., and Patel, V.C., Neal-Wall Turbulence
Models for Complex Flows Including Separation,
AIAA Journal, Vol.26, No.6, 1988, pp.641-648.

Hoffmann, K.A., and Chiang, S.T., Computational


Fluid Dynamics for Engineers, Engineering Education System, 1993.

Chen, H.C., Lin, W.M., and Weems, K.M., Interactive Zonal Approach for Ship Flows Including Viscous and Nonlinear Wave Effects, Proceedings, 6th
International Conference on Numerical Ship Hydrodynamics, Iowa City, U.S.A., 1993, pp. 341-363.

Kim, K., Beck, R.F., and Sirviente, A.I., A ViscousInviscid Interaction Study using Complementary
RANS Equations, Proceedings, 8th International
Conference on Numerical Ship Hydrodynamics, Busan, Korea, 2003, pp. 248-260.

Chen, H.C., and Lee, S.K., Interactive


RANS/Laplace Method for Nonlinear Free Surface
Flows, Journal of Engineering Mechanics, Vol.122,
No.2, 1996, pp.153-162.

Kim, K., A Viscous-Inviscid Interaction Study using Complementary RANS Equations, Ph.D. Thesis,
Department of Naval Architecture and Marine Engineering, University of Michigan, 2004.

Chen, H.C., and Lee, S.K., RANS/Laplace Calculations of Nonlinear Waves Induced by SurfacePiercing Bodies, Journal of Engineering Mechanics,
Vol. 125, No. 11, 1999, pp. 1231-1242.

Launder, B.E., and Spalding D.B., The Numerical


Computation of Turbulent Flows, Computer Methods in Applied Mechanics and Engineering, Vol.3,
1974, pp.269-289.

Chorin, A.J., Numerical Solution of the NavierStokes Equations, Mathematics of Computation,


Vol.22, 1968, pp.745-762.

Rhie, C.M., and Chow, W.L., Numerical Study of


the Turbulent Flow Past an Airfoil with Trailing Edge
Separation, AIAA Journal, Vol.21, No.11, 1983,
pp.1525-1532.

Dommermuth, D., Innis, G., Luth, T., Novikov, E.,


Schlageter, E., and Talcott, J., Numerical Simulation of Bow Waves, Proceedings, 22nd Symposium
on Naval Hydrodynamics, Washington D.C., U.S.A.,
1998, pp.508-521.

Stern, F., Yoo, S.Y., and Patel, V.C., Interactive and


Large-Domain Solutions of Higher-Order ViscousFlow Equations, AIAA Journal, Vol. 26, No. 9, 1988,
pp. 1052-1060.

Ferrant, P., Gentaz, L., Alessandrini, B., and Le


Touze, D., A Potential/RANSE Approach for Regular Water Wave Diffraction about 2-d Structures,
Ship Technology Research, Vol.50, 2003, pp.165-171.

Tahara, Y., Stern, F., and Rosen, B., An Interactive


Approach for Calculating Ship Boundary Layers and
Wakes for Nonzero Froude Number, Journal of Computational Physics, Vol. 98, 1992, pp. 33-53.

Gregory, N., and OReilly, C.L., Low Speed Aerodynamic Characteristics of NACA 0012 Airfoil Section,

14
Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Jacek Pawlowski
TRDC Inc., NL, Canada
To preclude possible misunderstandings in
the discussion, I would like to point out that the
approach taken in the presented work depends on a
flow superposition, with the superposed flows being
defined on the same domain. This may cause some
confusion for researchers who are used to thinking in
terms of matched regions of potential (or ideal
fluid) and viscous (e.g. boundary layer) flows.
My paper in this symposium elaborates on flow
superposition in a general and formally defined
context.
AUTHORS REPLY
The authors thank Dr. Pawlowski for his
comment. The approach used in this study indeed is
based on a superposition of flows within the same
computational domain. This approach offers the
advantage of giving equally accurate results in less
resolved grids than the conventional approach to
solving RANS equations, where such superposition
of flows is not done.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

EXPERIMENTAL STUDY OF THE FLOW


FIELD AROUND A ROLLING SHIP MODEL
M. Felli, F. Di Felice, C. Lugni
INSEAN (Istituto Nazionale di Studi ed Esperienze di Architettura Navale)

ABSTRACT
The experimental survey of the 3D velocity
field of a DDG551 ship model in forced rolling motion
have been carried out at the INSEAN Circulating Water
Channel with the aim to obtain suitable data for CFD
validation. Results are presented and discussed for the
bare hull and fully appended configurations. Flow field
is resolved in phase with the roll motion to point out
flow structures by using laser Doppler velocimetry. The
effectiveness of the bilge keels in damping the roll
motion is demonstrated by the generation of strong
vortical structures which are dominating in the flow
field.
INTRODUCTION
Safety of ship operation in a seaway has the main aims
to prevent damages to the avoidance of ship and to
secure the human safety. In this context, the prediction
of the motion of a ship in seaway has been the main
goal of the hydrodynamic research for many years.
A small amount of information about the geometry of
the ship and the sea state are sufficient to predict well
the response of the ship in the vertical plane (heave and
pitch motions). The added mass, the wave damping and
the restoring terms are the main contributes to the
hydrodynamic field. In this context the potential flow
theory is able to represent both the global motion
(Sclavunos, 1996, Colagrossi et al, 2001) and the local
flow field (Iwashita et al., 2000).
The rotational terms have been taken into account to
well predict the response of the ship in the lateral plane
(sway and yaw) and a Euler model can be used to
reproduce the hydrodynamic field around the hull.
Differently, roll motion is the most difficult to predict.
Roll motion is a lightly-damped and lightly-restored
motion and the roll natural period of the conventional
ship is very close to the richest region in the wave
energy spectrum.

This may cause very large amplitude roll motions even


in moderate seastate, if the wave frequency spectrum is
narrow and tuned with ship roll natural period. Ship
capsizing is the extreme consequence of this
phenomena.
Even in more moderate conditions the roll motion is of
concern, since undesirable effects may come from
excessive lateral accelerations motions. Parameters for
estimation of these effects are the Motion induced task
Interruption Indicator (MII) and lateral force estimator
(LFE).
From the theoretical point of view the prediction of the
roll motion is a challenging task: it is governed by the
viscous effect and the damping is due to the vortex
shedding phenomena. Navier Stokes equation is the
mathematical model able to reproduce the
hydrodynamic field around the hull in roll motion. The
difficulties in predicting the roll damping of a ship
depend not only on its nonlinear characteristics but also
on its strong dependence from the presence of the bilge
keels and from the forward speed of the ship.
The flow separation due to the ship bilge keel
significatively affects the evolution of the roll motion.
Therefore, the analysis of these vortical structures and
their evolution with the roll angle as well as the
evaluation of forces and moments applied to the ship
during the roll motion play a fundamental role in the
dynamics of the ship.
The difficult of implementing this kind of model
explains the strong development of the empirical
method (Ikeda et al., 1978 ). In particular in Himeno
(1981), the effects of several contributes to the roll
damping terms have been examined.
From the numerical point of view, several efforts have
been carried out concerning the 2D viscous flow
simulations (Yeung et al., 1998; Roddier et al., 2000).
Experimental works for validating theoretical and

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

numerical model have been performed (Yeung and


Cermelli, 1996).
Rapid advancements in computational fluid dynamics
(CFD) have enabled solution of increasingly complex
ship 3D flow simulations (Broglia et al, 2003). For
development and validation of CFD codes, much more
detailed model-scale, surface ship experimental data is
required. Official issues from ITTC 1999 and 2003
required explicitly EFD and CFD to be more and more
interlaced in order to provide reliable tools for fluid
physics understanding.
More in detail, EFD should be able to provide
highly reliable, time and space refined datasets for CFD
code validation. This requires a considerable effort in
designing and executing experiments that consider
more real flow conditions and address a variety of
physical aspects of interest.
The advancement in the measurement techniques
allows for a better fulfillment of these requirements. In
the present work an experimental investigation of the
3D velocity field evolution of a ship model in forced
rolling motion is presented and discussed.
Experiments have been carried out at the
INSEAN Circulating Water Channel, using a 2D laser
Doppler velocimetry (LDV) along eight transversal
sections of the ship model. The velocity field has been
acquired in phase with the ship roll angle.

water surface. Ship model under investigation was the


DDG51, whose main characteristics are reported in
figure 1. A sketch of the experimental set-up is shown
in figure 2.
Flow velocity components were measured by means of
a two components back scatter LDV system, in which a
5W argon laser produces a radiation that is collimated
by an underwater fiber optic probe into the
measurement point. With the adopted optics
(transmission lens with 465.5 mm focal length in water)
the ellipsoidal shaped measurement volume size was
0.2 mm by 0.2 mm by 4.5 mm (principal axes of the
ellipsoid).

Figure 2 LDV experimental set up

Figure 1 The test model

EXPERIMENTAL SET-UP
The experiments were carried out at the INSEAN
cavitation channel. This depressurized free surface
facility can reach a maximum speed of 5.3 m/s. The
test section has a rectangular cross-sectional shape, and
is provided with large viewing windows on the lateral
sides. The main dimensional characteristics of the test
section are: length = 10 m, width = 3.6 m, max. water
depth = 2.25 m with 1.0 m of freeboard above the free

The frequency shift, required for the velocity versus


ambiguity removal, was provided by a 40MHz Bragg
cell. Real time velocity analysis was performed by two
TSI IFA 550 processors.
The measurement of the three components was
performed in two separate steps by adopting two
different optical configurations of the LDV system
(Figure 4):
1. An optical configuration with the laser beams
coming from a side and the optical axis at 0 with
respect to the horizontal direction, in order to
acquire the axial (U) and the vertical (W)
component of the velocity (Figure 3 top).
2. An optical configuration with the laser beams
coming from a side and the optical axis at -30 with
respect to the horizontal direction, in order to
acquire the axial (U) and the cross component along
the -30 direction (axis Z) (V) (Figure 3 bottom).
Transversal component (V) was then calculated in post
processing by combining the two independent non
orthogonal measurements of the two optical
configurations .
The solution of using a configuration with the
underwater probe trimmed of 30 in place of a
configuration with the laser beams coming from below,

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

measuring directly the transversal component and


typically used for such a kind of measurements, was
adopted for minimizing the errors due to the probe
vibrations and strut deformation.
The underwater probe was set up on a computer
controlled traversing system which allows to get a
displacement accuracy of 0.01mm in all the directions
and to achieve an high automation of the LDV system.
This was performed by using a special target with
known position with respect to the ship reference frame
OXYZ. High care was required in the initial location of
the measurement volume in order to reduce positioning
errors of the two optical configurations.

in the same LDV dataset.


Velocity and angular position signals was acquired in
random mode. Modulation frequency of the angle
signal was chosen in order to obtain an homogeneous
data rate among the 3 LDV processors.

Phase sampling technique


Phase analysis represents a suitable tool to resolve in
details the evolution of the flow around the oscillating
bodies. Phase sampling was carried out by ensemble
averaging the velocity time history over several cycles
The following software procedure has been
implemented:
1. Scan of the complete angle time history (randomly
sampled because acquired as Doppler signal by the
third LDV processor) in order to accurately
determine the beginning and end of each single
cycle. The beginning of the cycle is defined as the
angular position with the midship plane aligned
along the vertical direction (roll angle = 0). In
order to accurately evaluate the time of beginning of
the cycles a linear interpolation is used. Before this,
time evolution of the roll angle, as acquired from the
LDV processor (red line in figure 4), was preprocessed and transformed into a low pass filtered
signal of sinusoidal shape (blue and green lines in
figure 4) to eliminate any steps and high frequency
fluctuations, and to improve the zero crossing times
determination.

Figure 3: LDV optical configurations

In order to improve the Doppler signal processor data


rate and to reduce the acquisition time at point, the
tunnel water was seeded with 1-m Titanium dioxide
(TiO2). Data acquisition was accomplished by using a
low end personal computer, whereas the post
processing analysis required several Gbytes of data
storage.

Roll angle measurement


The time evolution of the angular position of the rolling
body was acquired simultaneously with the velocity
time history, by means of a 3600 pulse/revolution
optical encoder installed in the rotating axis. In
particular, a procedure was devised to make the
correspondence between the randomly acquired
velocity bursts and the angular position of the hull. This
was obtained feeding the frequency modulated signal of
the angular position of the rolling hull to a third LDV
processor. In this way the time history of the angle is
simultaneously obtained with the velocity time histories

Figure 4: Time evolution of the roll angle


reconstruction

2. The velocity and angle time signals are split into


arrays of time histories whose time lengths are equal
to a cycle. On the basis of the zero crossing times
the measured oscillation period of each cycle was
calculated and then compared and normalized with
the nominal period of the roll motion (in the
experiments set to 2 s). If the deviation is bigger
then a defined threshold the actual oscillation cycle

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

is not included in the statistical analysis


3. The time distance between two consecutive zero
crossing positions, corresponding to a complete
oscillation cycle, was divided in N slots, 2 wide. A
certain slot with center ti (0 ti T), will contain
each velocity sample acquired whose acquisition
time t* is such to respect the following belonging
condition:
(1)
ti - t* ti +
The statistical analysis was performed inside each
slot to obtain mean flow field and turbulence
intensity information. The accuracy of the statistical
analysis was improved by implementing a weighted
average procedure so that the influence of each
sample will progressively decrease with fixed law
(linear, Gaussian) as its distance from the slotting
center, with accuracy improvement (Felli et al.,
2000).
In the present analysis, statistical evaluation was
performed by using 50 slots per cycle of amplitude
2=0.04 s and weighted averaging by gaussian law. In
such a way a population of 200400 samples per slot
was collected.
Section

Distance from AP (mm)

B1

4873

B2

4305

B3

3731

B4

2823

B5

2152.5

B6

1578.5

B7

861

B8

Measurements were performed along eight transversal


sections of the hull (figure 5), whose position is
described in tab.1. Measurement grids were defined
looking at the spatial resolution, which must resolve the
wake structures, especially where highest velocity
gradients are expected, as well as minimizing the
number of points in order to reduce the facility
occupancy.
At this purpose Cartesian maps of about 350
measurement points, were used. Mesh size was not
constant and was reduced where strong velocity
gradients were expected.
The rolling motion always started with the symmetry
plane of the model aligned along the vertical position
and the ship rolling toward the starboard side.

Tab.1 Measurement planes distance from AP

Test Conditions and measurement grids


Measurements were performed at a speed of 2.097 m/s,
corresponding to Froude number of 0.28, with the ship
model in forced rolling motion at the frequency of 0.5
Hz and with an oscillation amplitude of 10.
The ship model was locked at the dynamic trim and
only the roll motion was allowed. Acquisition time was
90 s at point, corresponding to 45 cycles of the rolling
motion.
This choice was a good compromise between the
opposite requirements to minimize the acquisition time
at point and to assure the convergency of the first and
second order statistical estimators (see next section).

Figure 5 Measurement grids

Measurement errors
Error sources can be classified into three different

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

classes: instrumental accuracy errors, positioning


errors, post processing errors.
Instrumental accuracy errors: the INSEAN LDV
system accuracy, for the adopted optical configuration
(465.5 mm focal length in water), is estimated to be
around 1%.
Positioning errors: LDV probe displacements have
been carried out by the traversing system with an
accuracy of 0.05 mm. A position error within 0.25 mm
is made in the LDV probe alignment on the target
setting the origin of the co-ordinate system.
The uncertainty in the alignment of the LDV
measurement volume with the target introduces a bias
error in the calculation of the transversal component
due to the position mismatch of the two separate LDV
measurements especially in the regions of strong
velocity gradients.
The setting of the ship model roll angle zero, performed
by using an inclinometer is estimated to be within 0.05
of accuracy.

(mean velocity and turbulence levels) can be performed


by evaluating the confidence interval of the statistical
estimators.
By assigning a confidence level = 99%, and
considering a statistical population of 200 400
samples on average, the confidence interval from the
mean value is in the range from 1.5 % to 2 %.
In order to asses the optimum time length for the
acquisition some test have been performed. In figures
15 and 16 the convergence of the mean and standard
deviation is plotted for a time signal length up to 120 s.
Test campaign have been performed by adopting a time
length acquisition of 90 s at a point corresponding to
45 oscillations of the rolling body at 0.5 Hz with an
expected deviation of the statistical moments equal to
0.04% and 0.9%, for the first and second order
estimators, respectively (figures 15 and 16).
Deviation has been defined as follows:

Deviation =

value(t = 90) value(t = 120)


100
value(t = 120)

(2)

where the value for t=120 s has been assumed to be the


asymptote

Experimental results
Free decay test

Figure 6 Convergency of the 1st and 2nd order


estimators
Processing errors: the processing error is mainly due to
the slotting technique during the statistical analysis of
the acquired data. The averages inside each slot
smoothes the velocity gradients acting as a low-pass
filter. The adoption of a weighted slotting technique can
improve the result reducing this effect.
The uncertainty analysis of the statistical moments

The influence of the appendages is presented by the


time evolution of the roll angle, measured in calm water
by free decay tests, at Froude 0.138 and 0.207 (figure
7). The following four configurations were tested: bare
hull, hull with rudders (rudders), hull with rudders and
shaft brackets (rudders and brackets), hull with
rudders, shaft brackets and bilge keels (full appended).
The analysis points out that the bilge keels are the
dominant contribution to the damping.
A comparison between the decay time of figure 7
confirms the velocity-sensitive nature of the damping
effect: the faster the ship forward speed, the stronger
the damping effect.
An estimation of the different contribution provided by
the four configurations can be obtained calculating the
decay coefficient, as follows (Lloyd,1989):

( j)

ln( ( j + 1)

j =1

6 =
5
5

with (j) maximum value of the oscillation after j

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

cycles. Values of the decay coefficient are summarized


in table 2 for Froude 0.138 and 0.207.

eddy effect, commonly, their hydrodynamic design is


only oriented to minimise the resistance of the
appendage when the ship is underway.
Flow field measurements
Two configurations of the DDG51 with and without
appendages (bilge keels, shaft brackets, rudders) were
tested corresponding to the extreme conditions above
described in the free decay test. The main goal was to
obtain a limited amount of flow field data suitable for
CFD validation.
Angular evolution of the flow field around the rolling
ship is described by the representation of the velocity
field during the oscillation.
Each velocity distribution, in the measurement plane, is
related to the corresponding roll angle (t) between 0
and 5,
V(x,y,z,t) = F ((t) )
(4)

Figure 7 Free decay tests: time evolution of the roll


angle at Fr=0.138 (top) and Fr=0.207 (bottom)
As the forward velocity of the ship model increases,
the contribution of the rudder and the brackets to the
roll decay is more important with respect to the bilge
keels. In fact, the best is the hydrodynamic efficiency of
the appendage the largest is the hydrodynamic
contribution to the damping moment.
bare hull
hull + rudders
hull + rudders + brackets
fully appended

Fr = 0.138
0.068
0.076
0.077
0.095

Fr = 0.207
0.080
0.101
0.110
0.142

Tab.2 Decay coefficients


This aspect can be explained as a consequence of the
better hydrodynamic effectiveness of the rudder and the
brackets. In fact, as a consequence that keels work
primarily by absorbing roll energy through the viscous

In the following, for space reasons, just few


representative angular positions of the roll motion can
be shown. In figure 8 the evolution of the longitudinal
component of the velocity is shown for four different
angular positions of the bare hull model, spaced of
=0.5 max, in correspondence of the sections B1, B2,
B3. For such case, ship induced perturbation is meanly
concentrated downstream the bulbous bow, where
strong vortical structures, oscillating at the roll
frequency, take place, causing locally a flow slowdown
and a momentum transfer from the axial to the
transversal components (figure 8).
This is pointed out by the velocity defect of the axial
component and by the streamlines trajectories, that
locally tend to roll up generating two counter rotating
vortical structures. Elsewhere, the hull shape induces an
outboard deflection of the streamlines, downward direct
probably due to the bow wake shape. The bulbous wake
is quickly faded away downstream due to the process of
turbulent diffusion and viscous dissipation.
Appendages installation exhibits a strong deformation
of the streamlines trajectory with respect to the bare
hull configuration. This effect is pointed out in figures
9, 10 and 11, 12, that show the evolution of the three
dimensional wake and the turbulent kinetic energy
respectively along the bilge keels (figures 9 and 11) and
just behind the rudder (figure 10 and 12). Turbulent
kinetic energy is represented in percentage of the
freestream velocity. Velocity field is plotted at four
different positions of the rolling motion.
The contribution of the bilge keels and the rudder to the
drag can be estimated by analysing the amplitude of the
velocity defect induced from the appendages with
respect to the bare hull configuration. In particular, the
analysis of the average field (figures 9 and 10) and the

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

turbulent kinetic energy (figures 11 and 12) points out


clearly the viscous and the vortical wakes, that take
place respectively from the boundary layers coming
from the face and the back side of the appendages as
well as the momentum transfer to the transversal
components of the velocity due to the lifting effect.
This effect is very apparent just behind the rudder. In
fact, the combined effect of the rudder and the upstream
appendages (shaft, shaft brackets) causes a strong
slowdown of the flow and locally an apparent
increasing of the turbulence levels.
The turbulence distribution in the plane downstream the
hull shows weak traces of the wake generated upstream
by the bilge keels pointing out possible interaction of
the bilge keel vortex with the rudder.
CONCLUSIONS
An experimental survey of the 3D velocity field of a
ship model in forced rolling motion was carried out at
the INSEAN Cavitation Channel. Experiments were
performed using a 2D LDV anemometer along eight
transversal sections of the ship model, in a bare hull and
fully appended configuration. Velocity field was
acquired in phase with the ship roll angle. The
following conclusions can be pointed out:
The vorticity shed from the rudder is
significantly higher then that released by the
bilge keels. However, free delay tests show
that the bilge keels are the dominant
contribution to damping. This can be explained
by the fact that the lever arm is larger than that
of the other appendages.
Comparisons of the flow field between the
bare hull and the fully appended
configurations point out strong increase of the
viscous resistance due to the bilge keels and
the appendages, as demonstrated by the
turbulence level distribution.
The roll motion induces modulation of the
strength of the rudder tip vortex, due to the
small variation of the angle of attack during
the oscillation.
The flow field in the downstream planes points
out very complex unsteadiness due to the
vortex shedding from the bilge keels. This
effect is not well resolved by the adopted
technique, that seems to require a larger
number of cycles in order to filter out such
unsteadiness. At this purpose, a significative
improvement can be achieved using Particle
Image Velocimetry.

Acknowledgements
This work was carried out in the frame of 6DOF Project,
sponsored by the Italian Navy.

REFERENCES
Broglia, R., Di Mascio, A. Unsteady RANSE calculations of
the fow around a moving ship hull. 8th NSH Busan, Korea,
2003.
Yeung, R., Liao, S., Roddier, D. On roll Hydrodynamics of
Rectangular Cylinders, Proc. Int. Off. and Polar Eng. Conf.,
Montreal, 1998.
Roddier, D., Liao, S.,Yeung, R., On Freely-Floating
Cylinders fitted with bilge keels, Proc. Int. Off. and Polar
Eng. Conf., Seattle, 2000.
Yeung, R., Cermelli, C., Liao, S. Vorticity fields due to
Rolling Bodies in a Free Surface-Experiment and Theory,
21st ONR Symposium, Trondheim, Norway, 1996.
Sclavunos, P.D, Computations of wave-ship interactions,
Advances in Marine Hydrodynaimics, Editor: M.Ohkusu,
1996.
Iwashita, H., Nechita, M., Colagrossi, A., Landrini, M.,
Bertram, V., A critical assessment of potential flow models
for ship seakeeping, Proc. Osaka Colloquium, 2000.
Colagrossi, A., Lugni, C., Landrini, M., Graziani, G.,
Numerical and Experimental Transient Tests for Ship
Seakeeping, Int. J. Off. and Polar Eng., 11, No.3, 2000.
Ikeda, Y., Himeno, Y., Tanaka, N. A prediction method for
ship roll damping, Rep. No. 00405 of Dep. of Naval Arch.,
Univ. of Osaka, 1978.
Himeno, Y. Prediction of ship roll damping-state of art",
Dept.of Naval Arch. And Marine Eng, Univ. of Michigan,
Rep. No. 239, 1981
Felli, M., Di Felice, F., Romano, G.P., Installed Propeller
wake analysis by LDV: phase sampling technique,
Proceedings of the 9th International Symposium on Flow
Visualisation, Edimburgh, 2000.
Felli, M., Di Felice, F, Analysis of the propeller-hull
interaction by LDV phase sampling techniques, Journal of
Visualization, Vol. 7, No.1, 2004.
Lloyd A.R.J.M. Seakeeping: Ship Behaviour in rough
wather John Wiley and Sons ,New York 1989.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Figure 8 Evolution of the 3D wake (planes B1, B2 and B3)

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Copyright National Academy of Sciences. All rights reserved.

Figure 9 Evolution of the 3D wake: bare hull (left column) and fully appended (right column) configurations (planes
B4 and B6)

Twenty-Fifth Symposium on Naval Hydrodynamics

National Academy of Sciences. All rights reserved.


Figure 10 Evolution of the 3D wake:Copyright
bare hull
(left column) and fully appended (right column) configurations (plane B8)

Twenty-Fifth Symposium on Naval Hydrodynamics

Copyright National Academy of Sciences. All rights reserved.

Figure 11 Evolution of the TKE: bare hull (left column) and fully appended (right column) configurations (plane B4
and B6)

Twenty-Fifth Symposium on Naval Hydrodynamics

Copyright National Academy of Sciences. All rights reserved.

Figure 12 Evolution of the TKE: bare hull (left column) and fully appended (right column) configurations (plane B8)

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Stephane Cordier
Bassin dessais des Carnes, France
Can you give the choice of axis of rotation
used and what would be the influence of a change of
axis of rotation on the damping?
How does this choice compare to totally free
(6DOF) roll axis in natural decay test?
AUTHORS REPLY
We thank you very much for your question
which allows us to clarify some physical aspects not
well described in the paper.
The axis of rotation has been chosen as the axis
parallel to the keel line of the ship and passing
through the center of gravity of the model.
During the seakeeping tests with towed
model, it is common to consider the center of gravity
as center of rotation. In our experiments, in fact, the
model was kept fixed at the trim and sinkage values
obtained from the calm water test.
Therefore, the coupling between the vertical and the
roll motions plays a more important role than the
change of axis of rotation.
The figure below gives a recent
experimental result showing the non-dimensional
damping coefficient as a function of the roll mean
angle (left plot) and of the Froude number Fn (rigth
plot) for the ship model considered in the paper.

The continuous line represents the data obtained by


Bishop with a free model at the Fn=0.069. It is
evident that fixing the roll axis position influences
the non-dimensional damping.
Moreover, in a free roll test, the axis of
rotation is changing and depends on the coupling
with the other degrees of freedom.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics


St. Johns, Newfoundland and Labrador, CANADA, 8-13 August, 2004

An Investigation of Viscous Roll Damping Through the


Application of Particle-Image Velocimetry
Richard Bishop, Paisan Atsavapranee, Scott Percival, Jerry Shan, and Allen Engle
(The Naval Surface Warfare Center, Carderock Division)
the experimental uncertainty analysis will be
presented and discussed.

ABSTRACT
Todays use of potential flow programs for the
prediction of roll motion of surface ships is severely
limited, as empirically-based formulations must be
used to account for viscous effects around the hull.
Adding to this limitation is the fact that the most
widely-used formulations do not reflect the state of
the art with respect to hull form geometries. It has
been shown that unsteady RANS is sufficiently
accurate for the forward speed diffraction problem
and that extensions for six degree of freedom (6DOF)
simulations are promising.
As a result, a
comprehensive research program has been
established for the development, verification, and
validation of a six-degree-of-freedom motion
Reynolds-Averaged-Navier-Stokes (6DOF RANS)
prediction program. In the short term, application of
a 6DOF RANS program would provide the naval
architect with a means to apply ship-specific roll
damping coefficients in support of potential flow ship
motion predictions. Longer term, the use of 6DOF
RANS would allow for a rigorous and direct
assessment of ship motions in waves problem.
Within the context of this initiative, a
systematic series of model tests are planned to
explore the viscous flow field in the region of the
bilge keel of a ship.
Using particle-image
velocimetry (PIV) measurements, two-dimensional
unsteady flow patterns around the unclassified
DTMB model #5415 have been gathered and
analyzed. This paper will discuss the application of
PIV within an overall experimental fluid dynamics
(EFD) program of research in support of a 6DOF
RANS ship motions prediction program. Technical
issues related to the use of PIV systems in support of
viscous flow measurements for a ship at speed
undergoing roll motion will be discussed. Among the
issues to be addressed are the overall system
configuration, imaging optics, and particle injection
method. In addition, results from measurements and

INTRODUCTION
The Department of the Navy has a clear and urgent
need to accelerate the integration of Computational
Fluid Dynamics (CFD) techniques into the design
cycle of its surface-ship fleet. At present, CFD based
on solving the Reynolds-Averaged Navier-Stokes
(RANS) equations has matured to the point that most
complex turbulent flows can be computed. However,
its application to the design of surface ships in
complex wave conditions under a wide range of
operating speed continues to be limited by a low level
of confidence in the solution accuracy, a lack of
robustness in the computational techniques, and a
high demand on computational memory and time.
Remedies are needed for all these difficulties, with
the establishment of confidence in the solution
accuracy being of particular importance. To achieve
this goal it is essential to have accurate and reliable
experimental data, which are especially tuned to the
needs of CFD validations. The main objective of this
research program is to perform a series of captive
roll-decay
experiments
to
obtain
detailed
hydrodynamic data suitable for the validation of CFD
techniques based on RANS, such as flow field, rolldecay coefficients, and appendage force and moment
measurements. Focus is placed upon the roll motion,
where viscous effects are dominant, with a special
emphasis on flow dynamics around the bilge keels.
This paper describes results from experiments
performed in April 2003, which includes roll-decay
measurements and two-component velocity field
measurements with PIV near the mid section of the
DTMB model #5415 bare hull with bilge keels.
Ongoing and future work at NSWCCD includes bilge
keel, rudder, and propeller force dynamometry on
fully appended and bare hull with bilge keels
configurations in calm water and waves, along with

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

three-component velocity field measurements using


stereo-PIV.
This research program is part of an
international collaborative effort among the Naval
Surface Warfare Center, Carderock Division
(NSWCCD), Instituto Nazionale per Studi ed
Esperienze di Architectura Navale (INSEAN), Italy
and the University of Iowas Iowa Institute for
Hydraulic Research (IIHR). The main objective of
this initiative is to extend currently available
Unsteady Reynolds Averaged Navier-Stokes
Equations (URANSE) to include modeling and
numerical methods for 6 degree of freedom (DOF)
simulations. The solvers to be developed will
address issues related to viscous free-surface flows,
flow separation, turbulence modeling for ship flows,
gridding (adaptive, unstructured), and efficiency of
algorithms. Under the terms of the basic agreement,
CFD development of will be performed by IIHR and
INSEAN.
A complementary effort in experimental
fluid dynamics (EFD) has also been included as part
of this initiative. This portion of the program
requires the development and application of
advanced tow-tank measurement systems among
which include the use of PIV measurement
techniques, dynamometry and strain gage systems.
IIHR, INSEAN and NSWCCD are all participating in
this portion of the program. By using geosims of the
same hull form geometry (model 5415), issues such
as facility basis and scaling effects will also be
addressed.

bilge keel and lateral force variations associated with


vortex shedding are required. It is these issues that
the EFD portion of this research program hopes to
shed light on.
To this end, a concerted effort was formed at
NSWCCD to integrate and apply state-of-the-art flow
measurement and advanced dynamometry techniques
including PIV/SPIV, appendage dynamometry and
force and moment measurement to this and other
surface ship research programs. Thus far, application
of PIV/SPIV in the naval hydrodynamics community
has been extremely limited, especially in tow-tank
organizations due to a number of challenges in
applying optical measurement techniques in largescale facilities.
The application of advanced
dynamometry techniques has thus far been
commonly practiced in the submarine community at
NSWCCD, and this research program marks a
significant milestone in terms of the integration of
this expertise into surface ship research programs.
DTMB model #5415

BACKGROUND

The surface ship model used in this study is the


DTMB model #5415, an unclassified surface ship
model that has served as a common platform for
engineering and scientific investigation in the naval
hydrodynamics community.
The DTMB 5415
represents the state of the art in modern naval
combatant hull form design and is shown in Figure 1
in its body plan view. It was constructed out of
fiberglass, fitted with external bilge keels, and
ballasted according to the hydrostatics particulars
presented in Table 1.

Current methods used to predict ship motion in a


seaway are based on potential flow methods. Since
vertical plane motions are governed by potential flow
wave damping, predictions for heave and pitch have
been found to be quite accurate. However, unless a
ships geometry is such that it produces significant
amounts of radiated waves and has little viscous
damping (typically not the case for surface ships),
roll motions will be dominated by nonlinear viscous
and rotational effects. In order to account for these
nonlinearities, it has been the practice to correct
potential flow predictions by including viscous-fluid
empirical approximations. While these empirical
methods vary somewhat, these formulations typically
rely on the findings of Kato (1966), Schmitke (1978),
Ikeda, et al (1980), or Hadarra (1992) to account for
damping effects. Unfortunately the accuracy of these
empirically-based corrections is dependent upon how
well the hull form in question mirrors the hulls
included within the test database.
A clear
understanding of the local flow dynamics around the

Figure 1: Body plan view of DTMB model #5415

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Baseline Ship Properties


Displacement
Draft
Length LPP
Beam
KMt
KG
GMT
LCG
LCB
Roll Gyradius/Beam
Scale ratio

5415 Model
Metric
552 kg
0.25m
5.72m
0.76m
0.382m
0.305 m
0.077m
2.836 m from AP
2.836 m from AP
0.39
24.824

incorporate remote focus and moisture-sensing


features. The electronics in the camera housings are
kept dry by purging the internal volume with dry
nitrogen periodically during the experiment. Images
from each camera are captured and stored on a dataacquisition computer capable of streaming images at
the full camera frame rate onto a real-time disk array
system with a capacity of 480 gigabytes. For system
redundancy, images can also be optionally captured
to 1.8 gigabytes of RAM.

Table 1: DTMB 5415 hydrostatics


PARTICLE IMAGE VELOCIMETRY (PIV)
A new submersible PIV/SPIV system was
constructed and used for the first time for
hydrodynamics measurement at NSWCCD in this
research program. PIV measurement has found wide
usage in the academic community, thus the general
concept of PIV/SPIV will not be described here and
instead the discussion will concentrate upon unique
features that allow successful application of
PIV/SPIV in large-scale hydrodynamic facilities at
NSWCCD.
The system comprises of two
independent 2D PIV systems, which can be
configured for simultaneous two-component velocity
measurement in two planes or stereo-PIV (threecomponent) measurement in a single plane. In the
work described in this paper, simultaneous
measurements in two axial planes (at the LCG plane
and one station aft LCG) were achieved using the two
cameras, each focused onto a different light sheet of
different wavelength. A custom-designed optical
filter (a sandwich of an interference filter and a spike
filter) was fitted in front of each camera lens in order
to allow each camera to record only one wavelength
of light without contamination from the other light
sheet.
Cameras
The cameras (Roper Scientific ES 4.0), with a spatial
resolution of 2048 pixels x 2048 pixels and a full
frame rate of 15 Hz, are fitted with 85 mm lens and
filters designed to collect light at 585 nm and 598
nm, respectively and placed in submersible housings
with an umbilical of 30 m. in length. The camera
housings utilize modular design with interchangeable
nose, center body, and main bulkhead sections and

Figure 2: Submersible PIV cameras


Laser sheets
Illumination is provided by two flash lamp-pumped
dye lasers, operating at 585 and 598 nm. The pulse
lasers have a maximum optical energy output of 1
J/pulse in 10 microsecond-long pulses at a maximum
rate of 15 Hz. The beams from the lasers are coupled
into 600-micron optical fibers and formed into laser
sheets at the output end, using a series of beamforming optics placed in submersible housings. The
relatively long pulse duration (three orders of
magnitude longer than the typical 10 nsec pulse for a
Nd:YAG laser) makes it possible to launch the beams
into the fibers. The usage of fiber optics greatly
facilitates the placement of laser sheets, as mirrors
and open beams are eliminated from the optical
setup, with acceptable performance in terms of
optical loss through the fiber. Roughly 700 mJ/pulse
is recorded at the output end of a 30m long fiber
when the laser is operating at full power.
Seeding
Silver-coated hollow glass spheres with a mean size
of 40 microns and a specific gravity of approximately

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

1.0 was used as flow tracers in this study. The dry


powder was premixed into a tank of water and
injected using seeding manifolds into the
measurement volume along the model tow path. For
certain experimental conditions, the tow path can be
as long as 50 m; therefore, obtaining uniformly dense
seeding for the entire tow path can be difficult.
Typically, after waiting a certain period (30-45
minutes) since the end of a forward speed run,
seeding along the tow path is performed with the
carriage going backwards very slowly (0.07 m/s). A
seeding run can take more than 10 minutes, and even
a very slow current of 1 mm/s can move the seed
cloud out of the tow path during the time it takes to
complete a seeding run.
MANEUVERING AND SEAKEEPING BASIN
(MASK)
The experiment was conducted in the Maneuvering
and Seakeeping Basin (MASK) at NSWCCD. The
basin is approximately 110 m long, 73 m wide, and
6.1 m deep. As illustrated in Figure 3, the basin is
spanned lengthwise by a 114.6 m bridge supported on
a rail system that permits the bridge to traverse onehalf the width of the basin and to rotate through
angles up to 45 degrees from the longitudinal
centerline of the basin. Ship models can be towed in
head or following seas at any angle from 0 to 90
degrees. Tracks attached to the bottom of the bridge
support the towing carriage, 6.1 m wide, 6.6 m long,
and 2 m high with a maximum speed of 7.7 m/s.
Eight pneumatic wavemaker units are located along
one 73 m side of the basin and 13 units along one 110
m side.
They can be operated in unison or
individually, permitting regular wave generation up
to 0.6 m in height and irregular wave generation
simulating the ocean up to a scaled sea state 9.

TEST SETUP
The DTMB model #5415 was held rigidly to the
towing carriage by a heave-post mechanism, which
allowed precise adjustments of hull pitch angle and
sinkage. The heave-post mechanism was designed to
give the model the complete freedom to roll about a
pivot axis located longitudinally along the hull VCG,
while restricting all other degree of freedom motions.
Camera and light sheet strut assemblies were attached
to the main support beam, which was suspended
below the towing carriage at the forward and aft
ends. The strut assemblies were designed so that
they could be rotated out of the water for accessibility
to the instruments, and in the lowered position, be far
enough away from the model as not to introduce any
undesirable hydrodynamic interference. The camera
strut supported two camera housings placed side-byside, one focusing at the LCG plane, and the other at
one station aft LCG. The light sheet strut supported
two light sheet housings, which illuminated the two
planes of interest. The data acquisition computers
and dye lasers were placed on the carriage deck,
along with all other accompanying instruments.
Figure 4 illustrates a CAD rendering of the test setup,
while Figure 5 is a photograph of the setup with the
camera and light sheet struts in the raised position.

Figure 4: CAD rendering of test setup in the MASK

Figure 3: Elevation view of Maneuvering and


Seakeeping Basin (NSWCCD) at NSWCCD

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

generated using the remaining good runs and


presented in Figure 7, which compares the time
history of runs at various Froude numbers, in all
cases with an initial roll angle of 20 degrees.

20

PIV 45

15

PIV 46
PIV 47
PIV 48

10

PIV 49

Roll angle, degrees

PIV 50
PIV 51

PIV 52
PIV 53
0

PIV 107
0

10

15

20

25

30

PIV 108
PIV 109

-5

PIV 110
PIV 111
PIV 112

-10

PIV 26
PIV 27
PIV 37

-15

PIV 41
PIV 43
PIV 44

Figure 5: Photograph of test setup in the MASK

-20

TTL

-25

Time, sec

TEST PROCEDURES
Figure 6: Overlaid roll angle vs. time for 21 runs at
Fr = 0.138 with an initial roll of 20 degrees

20
15
10

Roll angle, degrees

For each test condition, the model was adjusted to the


desired sinkage and trim by lowering or raising the
forward and aft heave posts. At the start of each test
run, the model was manually rolled, and an electromagnet was used to hold the model to the desired
initial roll angle. The carriage was then brought to
the desired test speed, and after the initial transients
had subsided, the magnet was released, causing an
electronic trigger to start the PIV data acquisition.
PIV images were collected at the rate of 5.88 image
pairs/sec, while the time histories of the resulting roll
decay motion were collected at 500 Hz. At the end
of the test run, the towing carriage was parked at the
end of the tow path for a certain waiting period (3045 minutes) to allow any fluid motion in the basin to
dissipate. The carriage was then brought back to the
starting position at the seeding speed of about 0.07
m/s, while a trail of seeding particles was laid along
the tow path in the volume of flow measurements.
For each condition, at least 15 repeat runs were
performed.

5
0
0

10

15

20

25

30

35

40

-5

Fr = 0, 20 deg
Fr = 0.069, 20 deg
Fr = 0.138, 20 deg
Fr = 0.28, 20 deg

-10
-15
-20
-25

Time, sec

Figure 7: Roll decay comparison at various Froude


numbers with a 20-degree initial roll angle
Damping coefficients Nj for each roll cycle were then
calculated for using subsequent peak-to-peak
information using formula (1) and presented in
Figure 8.

ROLL DECAY RESULTS

Nj = 1/ * ln (j / j+1),

As Figure 6 shows, high run-to-run repeatability in


each experimental realization is obtained. Even
though the overlay of the time history of the roll
angle shows no appreciable scatter, a systematic
uncertainty analysis was performed on the data to
identify the outlying runs. The determination of any
outlying runs was performed by computing the tvalue or standard error for each condition.
Chauvenets criterion was then applied to identify the
outliers. An average of the time history is then

Where 1 is the absolute value of the first local


maximum of roll angle after the release of the
electromagnet and 2 is the absolute value of the
second local maximum of roll angle one half roll
cycle later, and so on. Mean roll angle for each data
point is the average of the absolute value of the two
adjacent local maxima:
Mj = (j + j+1) / 2,

Copyright National Academy of Sciences. All rights reserved.

for j = 1, 2, 3,

for j = 1, 2, 3,

(1)

(2)

Twenty-Fifth Symposium on Naval Hydrodynamics

Figure 8 shows a clear trend of damping coefficients


increasing with higher forward speed. Even though
this phenomenon, called lift effect, has been
documented in the literature, no clear explanation of
the dominant physics leading to this effect has been
offered thus far. The vortex shedding off the bilge
keels during the roll motion may very well play a
dominant role in this effect. PIV data show that the
opposite sign vortices formed during each half cycle
interact differently at different forward speeds
depending on the rate of downstream advection of the
system of vortices. Measurement of roll period for
each full cycle, presented in Figure 9 reinforces this
conjecture. The roll period shows a decreasing trend
with each subsequent cycle until it asymptotes to a
mean natural roll period after approximately 6 cycles.
It is reasonable to conclude that during the early roll
cycle where vortex shedding is strongest,
flow/structure interaction causes a nonlinear coupling
between vortex shedding and roll motion, which
results in the increased roll period. It is hoped that
with additional ongoing study at NSWCCD, which
includes the measurement of the lateral force on the
bilge keels, we can gain new and important insights
into this phenomenon.

A sample of PIV results for approximately the first


1.5 cycles of oscillation at Fr = 0 with an initial roll
angle 20 degrees are presented in Figure 10. The
vectors represent in-plane velocity, normalized by the
characteristic velocity 2ro/o , where r is the
distance from the roll center to the tip of the bilge
keel, o is the initial roll angle in radians, and o is
the natural roll period of the model in water. The
characteristic velocity 2ro/o can also be thought
of as the maximum linear velocity at the tip of the
bilge keel if viscous damping is zero. The color
contours represent the out-of-plane vorticity,
normalized by the characteristic frequency 1/o. A
complex system of counter-rotating vortices is
observed to be generated during each half cycle of
oscillation. Due to the induced velocity of a largescale coherent vertical structure to an opposite-signed
vortex, the vortex pairs have a tendency to be
convected away from the bilge keel, either towards
the water surface in one half cycle or towards the
centerline of the model in the other. Figure 10
includes only 1/64 of the numbers of velocity vectors
obtained with the measurement and only every other
PIV velocity vector field is shown.

0.18

CONCLUSIONS AND DISCUSSION

0.16
Damping coefficient

0.14
0.12

Fr =
Fr =
Fr =
Fr =
Fr =

0.1
0.08
0.06

0, Free Roll
0, 20 deg
0.069, 20 deg
0.138, 20 deg
0.28, 20 deg

0.04
0.02
0
0

10

15

20

Mean Roll Angle, degrees

Figure 8: Damping coefficients at various Froude


numbers with a 20-degree initial roll angle
2.4
2.38
2.36
Roll period, sec

2.34
2.32

Fr
Fr
Fr
Fr

2.3
2.28
2.26

=
=
=
=

0, 20 deg
0.069, 20 deg
0.138, 20 deg
0.28, 20 deg

2.24
2.22
2.2
2.18
0

10

15

20

Cycle

Figure 9: Roll periods at various Froude numbers


with a 20-degree initial roll angle
PIV RESULTS

A systematic series of model tests to explore the


viscous flow field in the region of the bilge keel of a
ship has been performed.
Using particle-image
velocimetry (PIV) measurements, two-dimensional
unsteady flow patterns around the unclassified
DTMB model #5415 at various Froude numbers have
been gathered and analyzed. Recently obtained data
has elucidated the flow dynamics of vortex shedding
around the bilge keel of a rolling ship and will lead to
new insights regarding physical mechanisms
dominant in viscous roll damping around a surface
ship. While these data provide valuable validation
and verification to 6DOF RANS development,
additional ongoing and future work will explore these
issues in even more details utilizing more advanced
flow-measurement tools (SPIV) and state-of-the-art
dynamometry techniques.
By performing
experiments using different configurations (fullyappended, fully-appended without rudders, bare hull
with bilge keels, etc) and at the same time
instrumenting these appendages with dynamometers,
we hope to quantify the component contribution of
various ship appendages to roll damping. These data
would not only serve to correlate to the quantitative
flow field measurement, hopefully yielding a better
physical understanding of the complex flow physics

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

but would also serve as valuable database for CFD


validation and ship designers as well.
REFERENCES
Kato, H., Effect of Bilge Keels on Rolling of Ships,
Memories of Defense Academy, Japan, Vol 4, No 3,
1966.
Schmitke,R., Ship Sway, Roll and yaw Motions in
Oblique Seas, SNAME Transactions, 1978.
Ikeda, Y., Masafumi, I., Tanaka, N., Visocus Effect
on Damping Forces of a Ship in Sway and Roll
Coupled Motion, J.S.N.A. No 176, 1980.
Haddara, M.R., Cumming, D., Experimental
Investigation into the Physics of Roll Damping of a
Long Slender Hull Form, International Shipbuilding
Progress, 1992.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Design Waterline

X/L = 0.504
Vorticity, xTo
96

0.85 secs
-14.16 Roll

Fr = 0.0
Initial Roll angle = 20

Hull Centerline

64
32
0
-32
-64
-96

Figure 10 (a): Fr = 0, 20 degrees

(d)

0.17 secs
13.64 Roll

(b)

1.19 secs
-14.59 Roll

(e)

0.51 secs
-1.71 Roll

(c)

1.53 secs
-4.27 Roll

(f)

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

1.87 secs
8.55 Roll

(g)

2.89 secs
-1.69 Roll

(j)

2.21 secs
14.23 Roll

(h)

3.23 secs
-10.45 Roll

(k)

2.55 secs
9.26 Roll

(i)

3.57 secs
-10.61 Roll

(l)

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Joseph Gorski
Naval Surface Warfare Center, Carderock Division,
USA
The authors have done a nice job of creating
a data set that will provide much needed time
dependent data for the very relevant problem of ship
roll motion that can be used to validate RANS and
other computational techniques.
It appears great
pains have been taken to ensure repeatability of the
data, which is demonstrated for roll angle vs. time.
Would the authors please comment on the
repeatability of the PIV data and specifically what
has been done to differentiate between mean data at a
particular instant in time of the roll decay and the
instantaneous measurements, particularly at the
higher Froude numbers where the turbulent boundary
layer may be impacting the instantaneous field. Also,
could the authors comment on how their data
compares to that obtained at IIHR and INSEAN.

Figure 1: u component of the velocity residual for Fr


= 0.069 and 0.138 and initial roll angle of 20 degrees

AUTHORS REPLY
PIV measurements were conducted at Fr =
0, 0.069, 0.138, and 0.28 for an initial roll angle of 20
degrees and at Fr = 0 and 0.138 at 10 degrees. For
each of the six conditions, 15-20 experimental
realizations were performed. In order to gauge the
consistency of the data, the velocity residual for each
of the velocity components was calculated. The
velocity residual for u for N averages is defined as
the following:

Ru , N =

(u
u

avg , N

uavg , N 1 )

AllFrames AllVectors

avg , N =10
AllFrames AllVectors

The residual for v is defined in a similar


manner. Examples of the behavior of the velocity
residual with increasing N are shown in the plot
below. Figure 1 is a plot of the u velocity residual as
a function of N for Fr = 0.069 and 0.138 and initial
roll angle of 20 degrees. As can be seen for both
cases, the velocity residual decreases to very low
values as N approaches 15, indicating that the total
average represents a good comparison for CFD
calculations. At N = 15, the u component of the
residual for Fr = 0.069 and 0.138 and initial roll angle
of 20 degrees are 0.0030 and 0.0035, respectively.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Arthur M. Reed
Naval Surface Warfare Center, Carderock Division,
USA
A set of very impressive experiments.
have several comments:

1.

The roll center is generally at the center of


gravity this is important to motion predictions,
and does not impact the significance of these
experiments.

2.

The bilge keel forces are stated to be relatively


constant with varying Froude No., but the
damping varies as does the roll period this may
be due to changes of the force phase, moving
bilge keel forces from damping to added mass,
decreasing roll damping and increasing roll
periods (or vice versa).

3.

20 of initial roll is not large roll, thus the use


of a nonlinear code such as LAMP is
unnecessary though it is interesting to see that
the empirical roll damping in LAMP is not
adequate

4.

There is a need for roll damping force/moment


data at truly large roll angles, angles or 60, 70,
or more these are going to be the damping
models.

AUTHORS REPLY
The discussers comments added valuable
insights to the discussion already presented in the
paper. The authors strongly agree that similar
experiments should be performed in at very large roll
angles. Such data do not currently exist, and all
previous measurements have been at relatively mild
roll angles. In order to improve the prediction of
capsizing events, it is crucial to understand the
physics of roll damping at such large angles and
obtain experimental data for motion predictive tools
such as FREDYN.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
William L. Thomas III
U. S. Coast Guard Engineering Logistics Center,
USA
I would like to congratulate the authors for
presenting a fine paper that examines a fundamental
measurement typically associated with seakeeping
model tests: This is the use of roll extinction tests in
the measurement of roll damping. My observation
has been that in recent years, most model tests in the
seakeeping arena have been dedicated to
developmental or applied aspects of R & D. It is very
nice to read about seakeeping model tests oriented
toward more fundamental research. The Maneuvering
and Seakeeping Basin at NSWCCD needs to perform
more testing at this level.
Having assisted in a number of roll
extinction tests, I was very pleased to view concrete
and irrefutable evidence that ship motion data, in this
case roll extinction tests, are very repeatable when
steps are taken to sufficiently control the
experiments, as illustrated in Figure 6 of this paper. I
have no doubt that most people performing
seakeeping tests believe that the model performance
during any run is repeatable, provided that all initial
conditions and external forces are precisely the same.
The frustrating fact that seakeeping model testers
face is that budgets for experiments are very seldom
sufficient to allow the elaborate set-up required to
prove this. Thank you for providing an excellent
example.
Close examination of Figures 7, 8, and 9
deserve comment. In Figure 7, one observes the
classic increase in roll damping as speed increases.
One often attributes this to lift effects, for example an
increase in lift experienced by the bare hull. Your
paper also implies that the vortex shedding at the
bilge keep might also play a significant role,
exhibiting non-linear speed dependant behavior.
Would the authors like to comment on this?
An inspection of Figure 8, starting from the
right hand side of the plot, working left, one observes
expected trends in roll damping with the exception of
very small mean roll angles. The expected trend is to
observe larger damping at higher speeds due to lift,
and larger damping at higher mean roll angle angles
due to higher roll velocities experienced during larger
roll cycles. The sudden increase in roll damping
displayed at small roll angles (perhaps 1 degree or
less) during the runs at speed differ significantly from
measurements that I have observed in the past.
Would the authors like to comment on this? Could it
be that the measured data at such small angles has
fallen in the uncertainty region of the test

measurement or could noise in the test basin have


disturbed the roll data at very small angles?
In Figure 9, expected trends are observed at
all speeds, except the highest speed, Froude Number
0.28. During the early cycles, I would expect the roll
period to be slightly longer due to increased damping
present with the higher roll velocities. The plotted
results for Froude Number 0.28 do not follow a trend
consistent with data at other speeds. This is very
puzzling. Can the authors explain why the data at
Froude Number 0.28 plots significant different from
the other data?
I am very glad to observe efforts undertaken
to examine bilge keel interaction with the hull as well
as the collection of data for validation of CFD
predictions. Some issues that might be examined in
future tests might include:
Variation of the flow field at different
points along the bilge keel
Changes in vortex shedding at moderate to
high roll angles
Differences in the flow field between calm
water and the presence of waves
Once CFD is deemed suitable for the
prediction of roll damping, it will be necessary to
plan how this technology might be implemented into
seakeeping codes. Would seakeeping simulation
programs be expected to run CFD within the
simulation? Should this be done in 2D? 3D? As an
alternative, should CFD be used to upgrade the
existing empirical roll damping formulations in the
seakeeping codes as a means to have minimal impact
on run time?
In order to be effective, the implementation
strategy for CFD needs to be practical in terms to
ship acquisition schedules such that trade-off studies
can be performed early in the process.
AUTHORS REPLY
For a 2-D problem of a cylinder oscillating
in a viscous medium, the force on the cylinder is
usually represented by Morisons equation:

F=

1
1
dU
DCd U U + D 2 Cm
2
4
dt

At least for the case with no forward speed,


one may think of the localized flow field and
sectional bilge keel force as being idealized by a 2D
wedge-shaped cylinder oscillating along an arc in
viscous medium. For the low range of KeuleganCarpenter number, UmT/D, Um being the maximum
cylinder velocity during each cycle, T period of
oscillation and D bilge keel span, such as in this

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

experiment, the drag force is dominated by the force


induced by the vortices generated during each half
cycle of oscillation. As the model speed increases,
the forward speed would induce additional normal
flow velocity to the bilge keel, depending on the local
angle of attack at each longitudinal location. The
nature of the flow physics is inherently complex and
three-dimensional, with the bilge keels now acting
partly as a lifting body (with its own set of vortices,
such as tip vortices generated by a lifting body in a
steady flow) and partly as eddy-making fins from roll
oscillation. In addition, the vortices generated during
each half cycle will be convected downstream,
further complicating the flow field.
The unexpected increase in damping
coefficient at small mean roll angle, seen in Figure 8,
is due to the mechanical friction in the roll
mechanism used to constrain the model. In order to
isolate the effects of roll motion, the model was
constrained in pitch, heave, yaw, sway, and surge but
allowed to roll freely around a mechanical bearing.
At small mean roll angles, fluid forces acting to damp
the roll motion are very small, while the friction in
the bearing dominates. This point was not made
clear in the paper. It appears that below 2 degrees,
the damping coefficients presented in figure 8 are
dominated by the friction in the bearing and not by
fluid forces on the hull and bilge keels.
The large scatter in roll period at Fr = 0.28
may be due to the effect of increasingly unsteady
turbulent flow around the hull. At very low forward
speed, the flow around the hull is dominated by the
damped sinusoidal roll motion and its associated flow
structures, such as vortex shedding from the bilge
keels. The higher roll period during the first few roll
cycles is likely due to a combination of added mass
effect and increased damping from stronger bilgekeel vortex shedding. During subsequent roll cycles,
both of these effects monotonically decrease, causing
the roll period to decrease towards an asymptotic
value. However, as the forward speed increases,
other effects come into play. The strength of the
bow-dome vortices increases with forward speed and
may interact with the hull and bilge-keel vortices in
an unsteady manner. The boundary layer around the
hull also becomes increasingly more turbulent at
higher forward speeds. Both these effects may
contribute to the larger scatter in the roll period on
the order of 1.5% of mean roll period. It is
interesting to note that roll period for Fr = 0.138 also
exhibits larger scatter towards the later cycles as roll
mean angle decreases. Roll period decreases in a
well behaved manner until about the 7th cycle, when
larger scatter becomes apparent.
The authors strongly agree with the
discussers opinion on the direction for future work.

We plan to investigate the effect of beam sea and


stern quartering sea on roll damping in Summer
2005. Also, plans exist for a study of very large roll
motion on an advanced hull form in 2006.
Ultimately, these experimental data will be used to
validate CFD, which itself can be used in many
different ways. In the least resource intensive
fashion, steady CFD calculations (and also
experimental data) can be used to determine damping
coefficient on a wide range of conditions to improve
empirical roll damping formulations in seakeeping
codes. More sophisticated strategies may include
integrating CFD in design optimization or using CFD
such as 6DOF RANS to narrow the field of candidate
designs to reduce seakeeping testing costs. To echo
the discussers last comment, the authors agree that
in order to be effective, the implementation strategy
for CFD needs to be practical in terms of ship
acquisition schedules such that trade-off studies can
be performed early in the process.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics


St. Johns and Labrador, Newfoundland, Canada, 8-13 August 2004

Towing-Tank Tests for Surface Combatant for Free Roll


Decay and Coupled Pitch and Heave Motions
M. Irvine, J. Longo, and F. Stern
(IIHR-Hydroscience & Engineering,
The University of Iowa, Iowa City, IA, USA)
Reynolds-averaged Navier Stokes (RANS) simulations
of ships.

ABSTRACT
Towing-tank experiments are performed for an
advancing surface combatant in free roll decay and
coupled pitch and heave motions. For free roll decay
experiments, results are presented for motions (surge,
sway, heave, roll, pitch and yaw), forces (resistance,
sway and heave), moments (pitch and yaw), phaseaveraged velocities (U, V and W) for measurement
region near bilge keel and free surface elevations. For
coupled pitch and heave experiments, results are
presented for pitch and heave transfer functions, and
pitch and heave phase angles. The geometry of interest
is DTMB model 5512, which is a 1/46.6 scale geosym
of DTMB model 5415 (DDG-51), with L=3.048 m. The
experiments are performed in a 3x3x100m towing tank
equipped with a plunger-type wavemaker.
The
measurement systems include Krypton contactless
motion tracker, 4-component load cell, towed 2-D
particle image velocimetry (PIV) system, and servo
wave probes with 2-D traverse. Uncertainty assessment
following standard procedures is used to evaluate the
quality of the data.
Pitch and heave transfer functions and phase angles
collapse to a single value independent of wave
steepness. Free roll decay results show the addition of
bilge keels to a ship model increases roll period and roll
damping. Results show non-linear roll decay for Fr
0.138 and linear roll decay for Fr 0.190 for both
without and with bilge keels. Phase-averaged flow-field
velocity results show the evolution and subsequent
decay of the bilge keel vortex. The vortex trails the
motion of the bilge keel rotating clockwise for counterclockwise model rotation (rolling to port) and rotates
counter-clockwise for clockwise model rotation (rolling
to starboard).
The phase-averaged wave-field
resembles the steady wave pattern (Kelvin wave
pattern) with a superimposed oscillation due to the
rolling motion of the model. As the model rolls,
alternating crests and troughs radiate from the hull
shoulder and dissipate forward, toward the bow with
time. The data set obtained in this research is extensive
and will be useful for validation and development of

INTRODUCTION

As the evolution of computational fluid dynamics


(CFD) codes for simulation-based design (SBD) in ship
hydrodynamics continue, more complex and detailed
experimental fluid dynamics (EFD) benchmark data are
necessary to validate and develop the Reynoldsaveraged Navier Stokes (RANS) CFD codes used in
SBD. RANS-CFD simulations are advancing from
simpler problems such as resistance and powering to
much more complex problems such as ship motions and
maneuvering. RANS codes have already been used
with success to predict ship resistance at the
Gothenberg 2000 Workshop on CFD in Ship
Hydrodynamics (Larsson et al., 2000) using benchmark
experimental data. The experimental data required for
validation for resistance and powering and steady flow
predictions can be obtained with load cells (for model
forces), pitot probes and PIV (for flow-field velocities).
However, maneuvering and seakeeping experiments are
much more complex and require innovative and
evolving technology and techniques to obtain the
necessary data for a moving model. A complementary
approach can use the RANS-CFD simulations to guide
the EFD measurements to reduce the time and scope of
data necessary for validation.
The first step in the evolution from resistance and
powering to motions and maneuvering is the unsteady
forward speed-diffraction problem (an advancing ship
in head waves). Towed 2-D PIV has been already been
used by Gui et al. (2001a) for steady flow
measurements of the mean and turbulent nominal wake
of DTMB model 5512. Longo et al. (2002) are the first
experiments to measure the flow-field (with 2-D towed
PIV) and Gui et al (2002) measure the wave-field (with
servo wave probe) and obtain resistance data (with load
cell) for the forward speed-diffraction problem. The
use of towed PIV allows for the acquisition of critical
turbulence data that is difficult to obtain for steady and
unsteady flows in water. These benchmark datasets are
used for forward speed-diffraction predictions by

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Wilson and Stern (2000) for a RANS solution using a


surface tracking method and by Wilson et al. (2004b) to
validate a RANS solution using a single-phase level set
method. The experimental datasets will be a test case
for Tokyo 2005 Workshop on CFD in Ship
Hydrodynamics. The next step in the evolution is ship
motions problems.
Vertical plane motion (pitch and heave)
experimental data from Journee (1992) is used for
validation and development of RANS-CFD by
Weymouth (2002). While vertical plane motions
measurements and predictions have been successful,
lateral plane motions (roll, sway and yaw) have proven
more difficult to measure and predict accurately.
Measurement of lateral plane motions usually requires
complex model rigging to accurately simulate ship
motions. Roll motions require either forced rolling gear
or a low friction roll mount to simulate roll. Sway and
yaw motions require a planar motion mechanism
(PMM) to move the model in the desired motion. In
addition for oscillatory/periodic motions like roll, sway
and yaw, a large number of measurement samples are
needed to produce adequate phase-averaged data.
Wilson and Stern (2002) present RANS-CFD results for
forced rolling and free roll decay without bilge keels
that compare favorably with preliminary results of this
papers research. Solving for the roll motion of a ship
has been a difficult problem for engineers. The
problem arises from non-linear, viscous roll damping
(the roll damping is proportional to the square of the
roll velocity) of rolling ships. Viscous effects such as
skin friction, eddy shedding (from the hull and
appendages) and appendage forces contribute
significantly to the roll damping. Since viscous effects
are neglected in traditional strip theory, other methods
must be developed to solve for roll motions (i.e.
equivalent linear damping, RANS-CFD, etc.). Since
vertical plane motions can be measured and predicted
satisfactorily with existing methods, the next step will
be roll motions. Roll motions are extremely important
in modern naval combatant designs, since current
designs exhibit capsizing tendencies.
In ship design, a static free roll decay test is
typically performed on a ship model to evaluate its
natural roll period and roll damping characteristics. As
an initial problem for roll motions, this paper will focus
on obtaining detailed experimental data for an
advancing surface combatant in free roll decay. There
are two objectives for this research. The first objective
is to gain insight into physics of the flow-field for a
rolling ship hull due to the presence of the bilge keels.
The second objective is to obtain a detailed and
extensive benchmark data set of motions, forces and
moments, wave-field and flow-field measurements that
can be used in development and validation of a RANS
CFD code for ship motions calculations. These

objectives will be met by measuring global and local


data sets. Global data include ship motions in all 6
degrees-of-freedom, forces and moments on the model
and model speed. Local data includes wave-field
elevations and flow-field velocities.
Detailed
Uncertainty Analysis (UA) will be performed to
quantify errors and validate CFD results. The results
obtained in this research are extensive and cannot be
presented in their entirety in this paper. Results from
the complete dataset will be presented in Irvine (2004).
This research is part of an international
collaboration of IIHR Hydroscience & Engineering at
The University of Iowa, the Naval Surface Warfare
Center Carderock Division (NSWCCD, formerly David
Taylor Model Basin) and the Italian Ship Model Basin
(INSEAN) in Rome, Italy. Data will be compared
between facilities to quantify facility biases.
2
2.1

TEST DESIGN
Facility

The experiments are performed in the IIHR towing


tank. The tank is 100 m long, 3.048 m wide and 3.048
m deep, and is equipped with a drive carriage, plunger
wavemaker and wave dampener system. There are two,
right-hand, Cartesian (x, y, z) coordinate systems for the
present experiments. The global coordinate system is
used for the ship motions and forces and moments
measurements. Its origin is at the intersection of the
LCG, centerplane, and VCG. The x, y, and z axes are
directed aft, to starboard, and upward, respectively.
The local coordinate system is used for the flow-field
and wave-field measurements. Its origin is at the
intersection of the waterplane and the model FP. The x,
y, and z axes are also directed aft, to starboard, and
upward, respectively.
2.2

Model

The geometry of interest is DTMB model 5512


(Figure 1), which is a 1/46.6 scale geosym of DTMB
model 5415 (DDG-51), with L=3.048 m and block
coefficient, CB = 0.506. The model is unappended (no
shafts, struts, rudders or propulsors) except for the use
of bilge keels as noted. The DTMB model 5512 and
full-scale particulars are summarized in Table 1.
2.3

Data-Reduction Equations

Data-reduction
equations
are
used
with
individually measured results to compute additional
variables of interest.
2.3.1

Coupled Pitch and Heave Tests

Measured global variables of interest for


pitch and heave tests include model speed,
(surge, sway, heave, roll, pitch and yaw), and
wave elevations. Ship motion responses to

Copyright National Academy of Sciences. All rights reserved.

coupled
motions
incident
incident

Twenty-Fifth Symposium on Naval Hydrodynamics

wave spectra are typically presented as transfer


functions which relate the ship motion response to the
incident wave amplitude 0, in a non-dimensional form.
The model speed, Uc is determined by the speed of
the carriage. The carriage speed is determined by the
following DRE:
Uc =

nD
(1 +UB )
8000t

e = 2f e

2.3.2

(1)

34

1 Frh2

(2)

where Am is the midship sectional area of the model, AT


is the cross-sectional area of the towing tank, Lm is the
length of the ship model, BT is the towing tank width
and Frh is the Froude number based on the towing tank
depth, h.
The data-reduction equation for pitch transfer
function, TFx5 is defined as:
x5
TFx 5 = 1
(3)
k 01

n=

x4 m =

(4)

01
where x31 is the 1st-harmonic amplitude of the heave
response.
The encounter frequency, fe in hertz is calculated
using the incident wave frequency, fw in hertz. The
incident wave frequency is calculated as:
fw =

fd =

n =

Uc

(7)

(10)

d
2

(11)

(12)

1 2

where is determined from the roll time history by:

The relationship between the encounter and incident


wave frequencies is:
fe = fw +

x4 , j + x4 , j 1

The natural roll frequency n (rad/s) is determined from


the damped roll frequency d and the damping ratio :

(6)

(9)

where x4,j is the roll amplitude at the jth crest or trough


of the roll time history, and x4,j-1 is the roll amplitude at
the jth 1 crest or trough of the roll time history.
The damped roll frequency fd is determined by
locating successive zero crossings in the x4 time
histories. fd (Hz) is related to d (rad/s) by:

where is the wavelength of the incident wave.


The data-reduction equation for heave transfer
function, TFx3 is defined as:
x3
TFx 3 = 1
(5)

x
1
ln 4n
2 x4 n+1

where x4-n is the roll displacement at the nth roll cycle


crest or trough, and x4-n+1 is the roll displacement at the
nth+1 roll cycle crest or trough. In this study, roll decay
coefficient is evaluated from an average of five or six
successive peaks and troughs.
The mean roll angle, x4m is defined as:

where x51 and 01 are the 1st-harmonic amplitude of the


pitch response and incident wave, respectively, and k is
the wave number. The wave number is defined as:
k=

Roll Decay Tests

The measured global variables of interest for roll


decay tests include model speed, motions (surge, sway,
heave, roll, pitch, and yaw), resistance, heave force, and
pitch moment. The measured local variables of interest
are flow-field velocities and wave-field elevations,
however, for these tests, it is necessary to measure and
monitor the motion of the model through measurement
of surge, sway, heave, roll, pitch, and yaw.
Additionally for these tests, the loadcell is tipped 90
(compared with global tests) and integrated into the roll
mount which enables measurement of resistance, sway
force, and yaw moment.
Roll motion is analyzed to examine the
decay/damping of the model in free roll decay. The
decay of the model is quantified by the roll decay
coefficient, n and is defined as:

where n is the pulse count, D is the wheel diameter, t


is the time interval of the pulse and UB is a blockage
correction factor to account for the presence of the
towing tank walls. UB is determined from Tamuras
formula as:
A L
U B = 0.67 m m
AT BT

(8)

x
1
ln 4 1
n 1 x4 n
1
x
4 2 +
ln 4 1
n 1 x4 n

The encounter frequency, e in rad/s is calculated as:

Copyright National Academy of Sciences. All rights reserved.

(13)
2

Twenty-Fifth Symposium on Naval Hydrodynamics

and n is the nth cycle number, x4-1 is the roll


displacement at the 1st roll cycle crest or trough, and x4-n
is the roll displacement at the nth roll cycle crest or
trough.
Experimentally, the total roll damping, b44 can be
evaluated from roll decay experiments by:
b44 = 2 c44 a44

for the coordinates x, y, z, respectively, and i=1, 2, ,


N is the number of PIV recordings in one carriage run.
Ck,i is written as:
Ck ,i =

(15)

The natural roll frequency, fn in hertz is calculated


using:
n
2

(16)

The damped roll period, Td and the natural roll


period, Tn are determined from their respective roll
frequencies as:
Td =

1
fd

(17)

Tn =

1
fn

(18)

C k ,i

Fx g
0.5 ( T )U c2 S

CS =

Fy g

0.5 ( T )U c2 S
Fz g
CH =
0.5 ( T )U c2 S
CM =

pa

x =

0.5 ( T )U SL
Mzg
CY =
0.5 ( T )U c2 SL
2
c

Lobj

Lobj
1 M
S k ,i
S k ,i ( t j ) =
Limg tU c M j =1
Limg tU c

(25)

W V

y
z

(26)

(20)

The phase-averaged wave-field elevations (znf, zff)


in the (x, y) coordinates are computed from multiple
datasets of wave-field measurements where znf and zff
are dimensional wave elevations in the nearfield and
farfield regions of the wave field, respectively. The
phase-averaged wave elevation at any time step from
i=1, 2, , N can be determined with:

(21)

(19)

pa

2.4

M yg

After the phase-averaged 3-D flow-field velocities are


computed, the axial vorticity is determined with the
following equation:

Resistance, sway force, heave force, pitch moment,


and yaw moment are converted to non-dimensional
coefficients with the following equations, respectively:
CT =

(24)

where Lobj is the width of the camera view in the object


plane, Limg is the width of the digital image, t is the
time between the first and second image of a PIV
recording, Sk,i(tj) is the time-varying component of the
particle image displacement obtained by evaluating the
digital PIV recording, and j=1, 2, , M is the number
of carriage runs.
The phase-averaged velocity
component at any time step from i=1, 2, , N can be
determined with:

and the contribution of the bilge keels to the total roll


damping, bBK can be evaluated by:

fn =

Limg tU c

k = 1, 2, 3; i = 1,2,..., N; and j = 1,2,...,M

(14)

bBK = b44( withBK ) b44( withoutBK )

Lobj S k ,i ( t j )

(22)

2.4.1

1 M z nf ,i
1 M z ff ,i
( t j ) or =
(tj )

M j =1 L
M j =1 L

(27)

Measurement Systems and Procedures


Krypton Motion Tracker

Ship Motions are measured using a Krypton


Motion Tracker. The motion tracker is a contactless
measurement system capable of measuring 6DOF
motions in real time. Contactless measurement is
achieved through the use of three CCD cameras that
measure the motion of three light emitting diodes
(LED). The LEDs are fixed to a target that is attached
to and moves with the ship model. Motions can be
measured at any point on the ship model by
configuring, in the data acquisition software, a shift
from the center of the target to the desired measurement
location. Data can be acquired real time at rates up to
100 Hz. Translations and rotations can be measured
with accuracies of 0.1 mm and 0.04 degrees,
respectively. The motion tracker camera is factory

(23)

Where Fx, Fy, Fz are the dimensional resistance, sway


force, and heave force in mass units; My and Mz are the
pitch and yaw moments in mass-length units; g is the
gravity constant (9.8031 m/s2); (T) is the density of the
towing tank water at the measured towing tank water
temperature; and S is the wetted surface area of the ship
model at the design waterline.
The phase-averaged flow-field velocities (U, V, W)
in the (x, y, z) coordinates, respectively, are computed
from multiple datasets of instantaneous PIV
measurements. For convenience the instantaneous
velocity components are defined as Ck,i where k=1, 2, 3

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

calibrated but the measurement system including


camera, target, and signal conditioner require a multistep setup procedure prior to making measurements.
Data acquisition begins by measuring reference
values for all 6DOF motions. The ship model is then
heeled to the desired initial roll angle. The carriage is
brought to a steady speed and data acquisition begins.
Approximately 2 seconds after data acquisition begins,
the model is released and data is acquired at 100 Hz.
Data is post processed with a FORTRAN 90
program. The data files are read by the code, voltages
are scaled using the correct calibration factors to
convert the measurements into engineering units. The
scaled reference value measurements are then
subtracted from the scaled data. The motions and force
and moment data are passed 4 times through an average
filter to remove high-frequency content. The damped
roll frequency fd is determined by locating zero
crossings of the roll time history and an average roll
period over 5 cycles is determined.
For each Froude number condition, a zero roll
angle run is made and is considered a bias that is
subtracted from each run made at a non-zero angle.
The crest and trough roll amplitudes of the first 5 roll
cycles are used to calculate a roll decay coefficient,
damping ratio, roll damped frequency and damping
coefficient for each cycle as well as average values over
the first 5 roll cycles. Time histories of each measured
variable and roll damping parameters summary are
output to Tecplot formatted files.
2.4.2

Approximately 2 seconds after data acquisition begins,


the model is released and data is acquired at 100 Hz.
Data is post processed with a FORTRAN 90
program. The data files are read by the code, voltages
are scaled using the correct calibration factors to
convert the measurements into engineering units. The
scaled reference value measurements are then
subtracted from the scaled data. The wave elevation
data is passed 4 times through an average filter to
remove high-frequency content. Time histories of each
measured wave elevation are output to a Tecplot
formatted file. Wave elevation time histories are
assembled to produce a time history animation of the
wave-field for free roll decay ship motion.
2.4.3

2-D Dantec Particle Image Velocimetry

The towed DANTEC PIV system combines


hardware and software that are integrated into a single
measurement system illustrated in Figure 2. The PIV
hardware components (hydrodynamic strut, laser, lightguiding arm, light-sheet optics, digital camera) are
assembled with a 2D, computer-controlled traversing
system capable of automated movement along the
transverse (y) and vertical (z) coordinates. Movement
in the x-coordinate is done manually. The strut is
pressurized, partly submerged, and contains a 20 mJ,
dual cavity Nd:Yag laser and light-guide arm for
steering 532 nm beams through the light-sheet optics,
which are housed in a submerged, streamlined torpedo.
The digital camera is a 1K1K (10081018 pixels)
cross-correlation camera fitted with a f/1.4 50 mm lens
that views the light sheet from a distance of 50 cm
through a 90 mirror. Vector map results are displayed
virtually in real-time at a rate of 7.5 Hz. Instantaneous
velocity data is phase-locked to the roll angle of the
ship model by sampling the analog signal for roll from
the Krypton Motion Tracker. Silver-coated hollow
glass spheres with a density of 1600 kg/m3 and an
average diameter of 15 m are used as seed particles.
Two calibrations are required before making PIV
measurements.
The first calibration requires
determination of the PIV scale factor or Lobj/Limg. The
scale factor is determined by recording an image of a
ruler at the desired focal length. The image of the ruler
is then compared to the image length and a scale factor
is determined. The second calibration requires towing
the PIV at forward speed, Uc in a uniform flow (no ship
model). PIV measurements are compared to the towing
speed. Any deviations from uniform flow (U =Uc, V=0
and W=0), can be used as measured biases and
compensated for in test measurements.
The ship model is heeled to the desired initial roll
angle. The carriage is brought to a steady speed. When
the model is at steady speed, the data acquisition is
initiated by the user selecting the data acquisition
virtual button in the LabVIEW data acquisition virtual

Kenek Wave Probe

Two short and two long servo wave probes are


used to measure the free surface on the starboard side of
the model. The short (5 cm range) and long (15 cm
range) servo probes are pre-calibrated Kenek wave
probes with resolutions of 0.1 mm and 0.3 mm,
respectively, and maximum probe velocity of 700
mm/s. The probes are mounted in pairs on Velmex
BiSlide precision linear slides. The slides allow
automated movement of the wave probes in both the yand z-coordinates with an accuracy of 0.0762 mm.
The slides and wave probes are mounted to a rail
system under the carriage that enables manual
adjustment of the probe positions in the x-coordinate.
The wave probes are statically calibrated in pairs by
incremental movement up and down about a reference
position while recording the output at each elevation. A
1st-order linear regression curvefit is computed for the
calibration datasets to determine the scaling factors for
the wave elevation measurements.
Data acquisition begins by measuring reference
values for all wave probes. The ship model is then
heeled to the desired initial roll angle. The carriage is
brought to a steady speed and data acquisition begins.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

A 1st-order linear regression curvefit is computed for


the calibration datasets to determine the scaling factors
for the forces and moment measurements.
Data acquisition begins by measuring reference
values for the measured forces and moment. The ship
model is then heeled to the desired initial roll angle.
The carriage is brought to a steady speed and data
acquisition begins. Approximately 2 seconds after data
acquisition begins, the model is released and data is
acquired at 100 Hz.
Data is post processed with a FORTRAN 90
program. The data files are input and the measured
voltages are scaled with the calibration factors to
convert the measurements into engineering units. The
scaled reference value measurements are then
subtracted from the scaled data. The force and moment
data is passed 4 times through an average filter to
remove high-frequency content. Force and moment
data is analyzed over the first 5 roll cycles for average
values. Time histories of the forces and moments are
output to Tecplot formatted files.

instrument (VI). Initially only motions, force and


moment, carriage speed and wave elevation data are
acquired.
After 2 seconds, a TTL (TransistorTransistor Logic) signal is sent from the data
acquisition synchronously to the solenoid release and
the PIV processor. The TTL initiates data collection by
the PIV processor. The laser enters a free-running
mode at 15 Hz and begins acquiring double images.
For each double image, the PIV processor makes one
sweep across its analog inputs which include the output
from the Krypton Motion Tracker roll signal. The roll
signal sampled by the processor is used to assign a roll
phase angle to each vector map.
PIV measurement area dimensions are 1921018
pixels (14.374.9 mm) or 18% of the total field of
view. Advantages of above area dimensions include
previous use by Gui et al. (2001a), higher data
throughput, and reduction of amplitude and phase errors
for unsteady tests (Longo et al., 2002). Interrogation
areas are 32x32 pixels, 50% overlap in both
coordinates, 8 pixels of offset in the axial coordinate,
and a Gaussian window function is used in the
correlations. With the above settings, the measurement
grid is 1162. Digital images are correlated using a
cross correlation algorithm with a data buffer for
essentially real-time data processing. Vector-map data
is exported after each carriage run into ASCII files
containing a matrix of vector positions (X,Y or X,Z),
instantaneous velocities (U,V or U,W), and ship model
roll phase angle.
PIV data is batch processed for each lightsheet
position. The vector positions, instantaneous velocities,
and roll phase angles are read into a FORTRAN
program. The velocities data are then filtered using a
three stage filter. The data are first filtered one time
using a range filter to remove outliers. The data are
filtered in the second stage using a median filter. The
data are filtered a third and final time with a range
filter. At each grid point in the data, a mean and RMS
value is calculated for n, number of runs. Convergence
is monitored by calculating a residual for U and V(or
W) velocities at each grid point. The average residual
over the whole time history is monitored. The 3-D
flow-field can then be reconstructed and presented as a
time history animation of the flow-field near the bilge
keel for free roll decay.
2.4.4

2.5
2.5.1

Measurement Conditions and Locations


Coupled Pitch and Heave Tests

Coupled pitch and heave motions tests with


forward speed and head wave for the condition without
bilge keels are completed to commission the Krypton
Motion Tracker contactless measurement system and
compare vertical plane motions from experimental
results with SMP. Motions in all 6-DOF (at the model
VCG) and incident wave elevation (forward of the
model at x = -0.2) are measured. The test conditions
are summarized in Table 2. Wave steepness Ak is the
product of the wave amplitude of 0 and the wave
number, k. For each Fr, each wave steepness is tested.
2.5.2

Roll Decay Tests

Roll Decay tests without and with bilge keels are


completed to evaluate the effect of bilge keels and
model speed on roll decay. The testing conditions are
summarized in Table 3. For each Fr, each initial roll
angle is tested.
For roll decay tests, global data measured are
model speed, motions (surge, sway, heave, roll, pitch
and yaw), and forces and moments (resistance
coefficient, heave force coefficient, pitch moment
coefficient, sway force coefficient and yaw moment
coefficient). Motions data are measured at the model
VCG.
For roll decay tests, local data measured are wave
field elevations and flow-field velocities (U, V, and W).
Local data are measured for Fr = 0.138 and x4i = -10.
The wave-field measurement grid (Figure 3) is
developed using a RANS simulation, of the wave-field
for DTMB 5512 in roll decay, to guide grid spacing and

4-component load cell

The forces and moments on the model are


measured using a 4-component (2 force, 2 moment)
strain gage load cell. The load cell is capable of
measuring forces of 20 kg, and moments of 20kg-m.
The loadcell is statically calibrated by incrementally
hanging standard weights of increasing and decreasing
mass from the center axis (forces) or moment arm
(moment) while recording the output at each load state.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

releasing the model to roll freely. The model is


accelerated manually and after approximately 2 model
ship lengths, the model is at steady speed. Data is
acquired over 14 model ship lengths. All ship model
motions are constrained except for roll.

placement to ensure coverage of significant wave


perturbations and optimize data collection. The grid is
divided into 6 zones of varying spacing and is
summarized in Table 4.
The flow-field velocities measurement area
(Figure 4) is developed using a RANS simulation, of
flow-field near the bilge keel for DTMB 5512 in roll
decay, to guide PIV cut placement to ensure coverage
of important flow features and optimize data collection.
The measurement plane is at x = 0.675 and consists of
17 horizontal (x-y) cuts and 14 vertical (x-z) cuts to
cover the region of interest near the bilge keel.
2.6
2.6.1

The uncertainty assessment (UA) of the


measurement results follows the ASME PTC 19.11998Standard (ASME, 1998). The UA procedures are
based on separation and identification of systematic
(bias) and random (precision) error sources, and
combination with a root-sum-square (RSS) procedure to
determine total uncertainty. 95% confidence levels are
maintained for both bias and precision limits through
careful selection of individual bias error sources and
small-sample multiple test approach for precision
errors.

Test Procedures
Ballasting

For dynamically moving models, it is critical that


the model is ballasted correctly such that the models
dynamic behavior is the same as the full-scale ship.
DTMB model 5512 is ballasted to achieve the desired
static sinkage and trim, vertical center-of-gravity
(VCG), pitch gyradius (k5) and roll gyradius (k4).
Ballasting is performed using standard procedures as
described in Stahl (1995). First, the model is ballasted
to the correct static waterline by adding ballast until the
model sits at the design waterline with zero pitch (trim).
Second, the model is ballasted to the desired VCG
(VCG = 0.163 m) using the added ballast method.
Ballast is shifted vertically as necessary to achieve the
desired VCG. Finally, the pitch and roll gyradii are
determined using the pendulum method. Ballast is
shifted longitudinally (pitch gyradius) and transversely
(roll gyradius) to achieve the desired gyradii. The
model is ballasted to the desired pitch gyradius, k5 =
0.762 m or 0.25*L. The desired roll gyradius, k4 =
0.150 m or 0.39*B, is not possible due to constraints in
moving sufficient ballast outboard on a small model
such as DTMB 5512. The model maximum, attainable
roll gyradius, k4 = 0.149 m or 0.385*B is the condition
used in these experiments. The ballasted properties of
the model are summarized in Table 5.
2.6.2

3.1

Background

UA procedures have been previously developed for


typical towing tank tests by Longo and Stern (2004) for
resistance, carriage speed and wave elevations
measurements and Gui et al. (2001b) for resistance,
wave elevations, heave force and pitch moment
measurements. The UA procedures for resistance,
heave force and pitch moment can be easily extended to
sway force and yaw moment. Using the previously
developed procedures, UA results for the current tests
will be presented for carriage speed, resistance, sway
and heave forces, pitch and yaw moments and wave
elevations.
The current measurements require the development
of new UA procedures for the ship motions
measurements. UA results are only presented for
heave, roll and pitch motions as all other motions are
constrained in the current tests. UA for flow-field
velocities is in progress and will be presented in Irvine
(2004).
3.2

Carriage Speed

Carriage speed uncertainties are determined using


the procedure outlined in Longo and Stern (2004). The
biases associated with carriage speed measurements are
pulse count, wheel diameter and time base. The UA
results for carriage speed are presented in Table 6.
The carriage speed bias and precision are expressed
as a percentage of the total uncertainty. The uncertainty
in carriage speed is expressed as a percentage of the
dynamic range of the carriage speed. The carriage
speed uncertainty is dominated by the bias errors. The
total uncertainty of the carriage speed is very small, less
than 1%.

Coupled Pitch and Heave Tests

Coupled pitch and heave motion tests consist of


generating a wave train of the desired steepness and
accelerating the model to the desired, steady
speed/Froude number. The model is accelerated
manually and after approximately 2 model ship lengths,
the model is at steady speed. Data is acquired over 14
model ship lengths. All ship model motions are
constrained, except for pitch and heave.
2.6.3

UNCERTAINTY ANALYSIS

Roll Decay Tests

Free roll decay tests with forward speed consist of


holding the ship model to the desired initial roll angle,
x4i, accelerating to the desired Froude Number, Fr and

3.3

Motions

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

The uncertainty assessment procedure for the ship


motions measurements is developed in this section.
Bias errors are identified and combined to determine
the total bias. The only bias error associated with using
the Krypton Motion Tracker is the bias of the
measurement system itself. The bias of the Krypton
system is determined by the placement of the LEDs on
the measurement target. From the LED placement and
factory specifications, it is determined that the bias for
linear motions is 0.1 mm and for rotations is 0.04
degrees. Since the only DOF in this research are heave,
roll and pitch, the UA will focus on these
measurements. The precision limit for roll motions is
measured 10 times for Fr = 0.138 and x4i = -10. The
UA results for pitch, heave and roll motions are
presented in Table 7. Uncertainties in all three motions
are small and dominated by bias errors.
3.4

biases associated with wave elevation measurements


are traverse calibration, calibration scatter and probe
misalignment.
The precision limit locations are
selected near the ship model in a region of significant
wave response. The precision limit locations are
position 1 (x = 0.68, y = 0.08) and position 2 (x = 0.68,
y = 0.13). The UA results are presented as averages
over the whole carriage run for the measured wave
elevations at positions 1 and 2 in Table 9. Uncertainties
for both positions are less than 8%. Position 2 is
located in a region of greater free surface fluctuation
than position 1, resulting in larger precision errors and
subsequently larger uncertainties for position 2
measurements.
4

The results of these experiments are extensive and


consequently all aspects of the data set cannot be
discussed in this paper. The results are divided into two
sections, coupled pitch and heave, and roll decay
experiments. The coupled pitch and heave test matrix
(Table 2) covers many Froude numbers and wave
steepnesses. All results for pitch and heave transfer
function collapse for all wave steepnesses. Results are
presented for coupled pitch and heave tests for a Froude
number, Fr = 0.28 and three wave steepnesses. Pitch
and heave transfer functions and phase angles from
experiments are compared with results from SMP.
The free roll decay test matrix (Table 3) covers
many Froude numbers and initial roll angles. For space
consideration, the majority of results are presented for
Froude number, Fr = 0.138 and initial roll angle, x4i = 10 as this is the condition for flow-field and wave-field
measurements. Results are presented for roll decay
tests without and with bilge keels. Motions, forces and
moments results are presented and evaluated for a range
of Froude numbers and initial roll angles, without and
with bilge keels. Roll decay results are then compared
to previous experimental data and SMP predictions.
Results for local data (flow-field and wave-field) are
presented for Froude number, Fr = 0.138 and initial roll
angle, x4i = -10. Flow-field results are presented for
phase-averaged velocities (U, V, W) with analysis
focusing on the physics of the flow near the bilge keel.
Wave-field elevations are examined for convergence.
Finally, wave-field elevations results are presented with
analysis focusing on the physics of the wave-field due
to the ship model in roll decay.

Forces and Moments

As previously mentioned, the UA procedures for


the force and moment measurements will follow those
outlined in Gui et al. (2001b) for resistance, heave force
and pitch moment. Those procedures are extended to
sway force and yaw moment measurements as they are
measured using the same instrument in a different
orientation. The biases associated with the force and
moment measurements are model wetted surface,
carriage speed, measured force or moment, and water
temperature and density. The uncertainty results are
presented as averages over the whole carriage run, for
resistance coefficient, CT, sway force coefficient, CS,
heave force coefficient, CH, pitch moment coefficient,
CM, and yaw moment coefficient, CY in Table 8.
The force and moment bias and precision are
expressed as a percentage of the total uncertainty. The
uncertainty in each force or moment is expressed as a
percentage of the dynamic range of its respective force
or moment. Uncertainties are low (U 2.1%) for
resistance and pitch moment, but significantly higher
(U 12.4%) for sway and heave force and yaw
moment. The higher uncertainties for sway and heave
force, and yaw moment are attributed to the motions
being restrained. Since those motions are restrained,
the measured motions are small and result in a small
dynamic range that is used to normalize the uncertainty.
As a result, the small dynamic range exaggerates the
uncertainty. Precision errors dominate the uncertainty
results due to the model having one DOF and motion in
that DOF yields vibrations in the restrained DOF.
These vibrations result in higher standard deviations
that lead to the increased precision errors.
3.5

RESULTS AND DISCUSSION

4.1

Coupled Pitch and Heave Tests

Figures 5a and 5b display pitch transfer function,


TFx5 and phase angle, 5, respectively, as a function of
encounter frequency, e for Ak = 0.025, 0.050, 0.075
and Fr = 0.28. Also included for comparison are SMP
predictions. The figures show that TFx5 and 5 are
independent of Ak as the data points mostly collapse to

Wave Elevations

The UA procedures for wave elevation


measurements follow those outlined in Gui et al.
(2001b) for near-field free-surface elevations. The

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

single values for each e. This trend is evident because


x5 is linear for the range of Ak in the tests.
Experimental data and SMP predictions exhibit similar
trends and magnitudes for the x5 magnitude and
frequency response. TFx5 and 5 magnitude differences
with SMP are likely due to use of a coupled pitch and
heave mount. Using a coupled pitch and heave mount
allows neither pure pitch nor heave such that either
motion may be slightly damped due to motion in the
other DOF. TFx5 and 5 frequency differences with
SMP are roughly 7% with a shift toward increased e
apparent for the measurements. These discrepancies
could be related to the combined effects of uncertainties
in the variables (fw, Uc, ) of equation (7) and
uncertainties in the phase calculation for x5. More
analysis is needed for a full understanding of the
observed differences.
Figures 5c and 5d display heave transfer function,
TFx3 and phase angle, 3, respectively, as a function of
encounter frequency, e for Ak = 0.025, 0.050, 0.075
and Fr = 0.28. Also included for comparison are SMP
predictions. Similarly as for pitch, the figures show
that TFx3 and 3 are independent of Ak as the data points
mostly collapse to single values for each e. This trend
is evident because x3 is linear for the range of Ak in the
tests. Experimental data and SMP predictions exhibit
similar trends and magnitudes for the x3 magnitude and
frequency response.
The level of TFx3 and 3
differences between measurements and SMP are very
similar to those for pitch. The source of the magnitude
and frequency differences is thought to be the same as
those for pitch and are not repeated here. More analysis
is needed for a full understanding of the observed
differences.
Results for the other Froude numbers are similar as
those presented for Fr = 0.28. Results from the coupled
pitch and heave tests confirm that the Krypton Motion
Tracker can be used to measure ship motions
accurately.
4.2
4.2.1

Figure 6 shows the roll time histories for Fr =


0.138 without and with bilge keels for initial roll angle
x4i = -10. As we would expect, results show increased
damping of the roll amplitude due to the presence of
bilge keels. An increased roll period is evident for the
condition with bilge keels. On average, the addition of
bilge keels lengthens the roll period by 3%.
Figure 7 shows the average damping ratio, versus
Froude number, Fr for seven Fr without and with bilge
keels, for all initial roll angles, x4i. Without bilge keel
data are denoted by circles and with bilge keels data are
denoted by squares. Results show that for all tests, the
damping ratio, < 1, verifying that the ship model
system in roll is underdamped. Since the system is
underdamped, the assumptions of a 1DOF
underdamped, spring-mass-damper are valid for
evaluating the problem of a ship model that is only free
to roll. The damping ratio ranges from 0.02 to 0.10
showing that the ship model is very lightly damped in
roll. The damping ratio increases with increasing
Froude number.
Figure 8 shows the average damping ratio, versus
initial roll angle, x4i. for all seven Fr, without and with
bilge keels. The damping ratio shows no apparent trend
or influence due to the variation of initial roll angle, x4i.
The damping ratio is typically higher for the condition
with bilge keels.
Figure 9 shows the roll damping coefficient for the
unappended model (bare hull) b44-BH, the total roll
damping coefficient (bare hull + bilge keels) b44, and
the bilge keel roll damping coefficient b44-BK versus
mean roll angle, x4m for Fr = 0.138 with both the
experimental (EFD) and SMP results. Results show the
total roll damping, b44 increases with increasing mean
roll angle. EFD results for the bilge keel roll damping
are nearly constant over the range of mean roll angles.
EFD results for the total roll damping, b44 compare well
with total roll damping predictions from SMP.
Comparisons of the bare hull damping b44-BH and bilge
keel damping b44-BK for EFD and SMP show significant
differences.
The contributions of the bare hull
damping, b44-BH and the bilge keel damping b44-BK to the
total roll damping are summarized in Table 11.
The EFD shows a larger contribution of the bare
hull (78.1%) to the total damping than SMP (16%).
The contribution due to the bilge keels for the EFD
(21.9%) is much smaller than SMP (84%). These
differences can be attributed to two factors. The first
factor is that the roll mount for the EFD may have
significant static friction (stiction) and overall friction
that is not evident upon visual inspection, but becomes
evident when examining the roll damping data. The
final factor is error in the theory/SMP in predicting the
bare hull damping of the DTMB 5512/DDG-51 hullform. This is likely a result of the near cylindrical
cross-section of the hull-form for much of the mid-

Roll Decay Tests


Motions

Ideally the roll time history should be 0.0 for x4i =


0, however it is not. An average roll angle is
calculated over the time history for x4i = 0 and this
average is used as a correction bias to correct carriage
runs at Fr = 0.280 where x4i 0. A correction bias is
similarly determined for each Froude number tested and
roll correction biases for all Fr are shown in Table 10.
Results show the correction bias increases with
increasing forward speed and must be used to correct
roll time histories. This bias is attributed to small
asymmetries of the ship model that yield a non-zero roll
angle when towed.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

body. However, the total roll damping, b44 compares


well since viscous effects dominate with the addition of
the bilge keels. For the EFD, the addition of bilge keels
overcomes any friction in the mount. Similarly for the
theory/SMP predictions, the total roll damping is
dominated by the damping due to the bilge keels, which
overcomes any deficiency in prediction of the bare hull
damping. The true results for the bare hull roll
damping, b44-BH and bilge keel damping, b44-BK most
likely lie in between the EFD results and the SMP
predictions.
Figure 10 shows the roll decay coefficient, n versus
mean roll angle, x4m for seven Fr without and with bilge
keels. Also shown on this figure is historical data from
Bishop (1983) of roll decay coefficient results for a 7-m
model of the DDG51. Results verify the observation of
increased decay/damping due to the presence of bilge
keels shown in the roll time histories. The increased
damping due to presence of bilge keels varies from 10
% at low Fr to 20% at high Fr. Roll decay increases
with increasing Fr due to the increase in lift damping
with increasing forward speed. Comparison of current
results with data from Bishop (1983) shows good
agreement for trend and slope of data. Results from
Bishop (1983) show larger magnitude (25% for without
bilge keels, 50% with bilge keels) for roll decay
coefficient, n than current results. This is attributed to
differences in the model scale and experimental setup.
Bishops results are obtained using a free running
model that allows motions in all 6 DOF. Current
results are restrained in all motions except roll and do
not allow for additional dissipation of energy and hence
damping due to sway and yaw motions.
Figure 11 shows roll decay coefficient, n versus Fr
for seven Fr, without and with bilge keels for all initial
roll angles, x4i. The trend of n with Fr shows that n
increases with increasing Fr with a plateau for 0.190
Fr 0.340 before increasing again. The roll decay
coefficient at Fr = 0.41 is 80% greater than the roll
decay at Fr = 0.069. The figure shows that in general,
n increases with increasing initial roll angle.
Figures 12a and 12b show damped roll frequency,
fd versus mean roll angle, x4m and damped roll
frequency, fd versus Fr respectively. As the ship model
is a lightly damped system, fd and natural frequency fn
are very close in magnitude and exhibit the same trends.
The damped and natural frequencies tend to decrease
slightly with increasing initial roll angle.
The
difference in frequency magnitude from lower to higher
initial roll angles increases with increasing Fr. Both fd
and fn increase with increasing forward speed. For the
figure showing frequency versus Fr, the data show
increasing variation of frequency as a function of initial
roll angle with increasing Fr. As expected, the addition
of bilge keels reduces the roll frequency magnitude
(and consequently increases roll period) for all Fr. The

frequency data is clustered near the natural period of


roll for zero forward speed, Tn = 1.54 seconds or a
natural frequency, fn = 0.649 Hz. For Fr = 0.069, the
frequency is 4% lower than the natural frequency. For
Fr = 0.410, the frequency is 12% higher than the
natural frequency.
Using the log-normal plots of the absolute value of
the crest and trough peak amplitudes, allows insight
into the linearity of the decaying motion. Motion is said
to be linear on a log-normal plot if the curve appears as
a straight line. Motion is said to be non-linear on a lognormal plot if the curve is concave upward. Figure 13
shows log-normal plots of the absolute value of the
crest and trough roll amplitudes versus time for all
seven Fr, x4i = -10 for without (Figure 13a) and with
(Figure 13b) bilge keels. From these figures we see
non-linear roll decay for Fr 0.138 for both without
and with bilge keels. Linear roll decay is evident for Fr
0.190 for both without and with bilge keels.
4.2.2

Forces and Moments

Figure 14 shows resistance coefficient CT versus Fr


for seven Fr without and with bilge keels and all initial
roll angles. Also included on this figure is archival CT
data from previous DTMB 5512 model tests for zero
roll angle at the dynamic sinkage and trim. Current
results without and with bilge keels for all Fr at x4i = 0
compare well with archival data. In general for all Fr,
results for CT with bilge keels is higher than the without
bilge keel condition. As Fr increases, CT tends to
collapse to zero roll angle value regardless of initial roll
angle, x4i. For Fr 0.280, CT tends to increase with
increasing x4i.
Figure 15 shows heave force coefficient CH versus
Fr for seven Fr without and with bilge keels and all
initial roll angles. CH varies with initial roll angle for
Fr 0.190 which is the non-linear roll region. CH
increases with increasing initial roll angle in this region.
For Fr 0.280, the heave force coefficient data
collapses and becomes independent of initial roll angle.
In general as Fr increases, the model sinks deeper into
the water (i.e. increased sinkage).
CH becomes
increasingly more negative with increasing Fr, which
represents the increased force required to hold the ship
model at the required sinkage. Heave force coefficient
is larger for conditions without bilge keels compared to
with bilge keels. This is due to that appendage lift
created by the bilge keels resulting in additional
negative (downward) heave force.
Figure 16 shows pitch moment coefficient CM
versus Fr for seven Fr without and with bilge keels and
all initial roll angles. CM varies with initial roll angle
for Fr 0.190 which is the non-linear roll region. CM
decreases with increasing initial roll angle in this
region. For Fr 0.280, the pitch moment coefficient

10

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

data collapse and become independent of initial roll


angle. Pitch moment coefficient is more negative for
conditions without bilge keels compared to with bilge
keels. That appendage lift due to the bilge keels results
in additional bow down pitch moment and the measured
force necessary to sustain the required pitch angle is
consequently larger (more positive).
4.2.3

third columns, lower velocity boundary layer fluid is


pulled away from the model and higher velocity fluid is
pulled in toward the model. With increasing cycle
number, the U velocity of the vortex/circular region
increases and the vortex/circular region size decreases.
This vortex/circular region is due to the presence of the
bilge keel and indicates a momentum deficit that results
in an increased drag or resistance on the model as
shown in Figure 14.
Figure 20 shows the contours of phase-averaged
velocity V for five roll cycles at quarter phases for Fr =
0.138 and x4i = -10 presented as described for Figure
19. As expected, Figure 20a shows a steady flow
pattern for V velocity before the model is released. In
the region trailing the bilge keel, in the second column
shows a region of positive V velocity near the model
and negative V velocity just below the positive region.
This indicates that the bilge keel vortex is rotating
clockwise when the model is rolling to port. In the
fourth column the model is rolling to starboard, the
bilge keel vortex is now on the inboard side of the bilge
keel. The negative V velocity region and the positive V
velocity region have switched positions such that the
negative velocity region is above the positive velocity
region indicating that the bilge keel vortex is now
rotating counter-clockwise as the model rolls to
starboard. With increasing roll cycle, the V velocities
decrease toward zero, indicating the decay of the bilge
keel vortex with time.
Figure 21 shows the contours of phase-averaged
velocity W for five roll cycles at quarter phases for Fr =
0.138 and x4i = -10 presented as described for Figure
19. As expected, Figure 21a shows a steady flow
pattern for W velocity before the model is released.
After the model is released (figure 18b), downflow
(negative W velocity) is evident near the bilge keel tip
and upflow (positive W velocity) is evident outboard of
the downflow. This indicates clockwise rotation of the
bilge keel vortex for the model rolling to port,
confirming the results shown in the V velocity contours
of Figure 20. For the third column, we see the
inception of another region of upflow interior of the
downflow just before the model begins rolling to
starboard. In the fourth column we see the outboard
upflow has decayed and moved toward the model while
the inboard upflow has increased in size and magnitude.
With the upflow now inboard of the downflow, the
bilge keel vortex is now rotating counter-clockwise as
the model rolls to starboard, confirming the results seen
in Figure 20 for the model rolling to starboard. With
increasing roll cycle, the W velocity upflow and
downflow regions decrease in size and magnitude,
indicating the decay of the bilge keel vortex with time.
Figure 22 shows the contours of axial vorticity, x
for five roll cycles at quarter phases for Fr = 0.138 and
x4i = -10 presented as described for Figure 19. The

Flow field

Figures 17a shows the complete 55 second roll


time history for Fr = 0.138 and x4i = -10. The roll
oscillations are seen to decay to essentially zero after 35
seconds and the flow-field returns to steady flow.
Figure 17b shows just the first 5 roll cycles of the time
history from Figure 17a with markers at the quarter
phase angles of the roll cycles. These markers denote
the phase angles used in both the flow-field and wavefield analyses.
Figures 18a, 18b and 18c show the steady flowfield velocity contours at x = 0.675 for U, V and W
velocity components respectively. U velocity contours
show a circular region of decreased velocity due to the
presence of the bilge keel and bilge keel boundary layer
growth. Contours also show slight boundary layer
thickening closer to the model centerline. Although
current measurements do not reach the centerline, the
boundary layer thickening is likely due to the sonar
dome vortices that are convected downstream near the
model centerline. V velocity contours show almost no
flow in the y-direction for steady flow. W velocity
contours show almost no flow in the z-direction except
for a small region of upflow interior of the bilge keel.
Figure 19 shows the contours of phase-averaged
axial velocity U for five roll cycles at quarter phases for
Fr = 0.138 and x4i = -10. The figure is laid out such
that figures in the same column are at the same phase
angle and figures in the same row are in the same roll
cycle. The roll phase angles are denoted by the markers
in Figure 17b. Figures in the first column are at
negative peaks of the roll cycle or maximum roll to
starboard. Figures in the second column are roughly at
x4 = 0 for the model rolling to port. Figures in the
third column are at positive peaks of the roll cycle or
maximum roll to port. Figures in the fourth column are
roughly at x4 = 0 for the model rolling to starboard.
Rows 1-5 correspond to the 1st-5th roll cycle
respectively. Contours show the evolution of the bilge
keel vortex, shown as a circular region of decreased
velocity that is developed at the bilge keel tip. The
vortex trails the bilge keel tip as the model rolls. In the
first column, the model is rolling to port and the bilge
keel vortex begins to develop on the outboard side of
the bilge keel. In the third column, the model begins to
roll back to starboard and the vortex moves to the
inboard and now trailing side of the bilge keel. As the
model rolls from starboard to port in the second and

11

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

bilge keel vortices indicated in the results for V (Figure


20) and W (Figure 21) velocity contours are more
clearly evident in the axial vorticity results. Overall,
the results show the roll up of a pair of counter-rotating
vortices at the bilge keel. The vortices initially, are
roughly the scale of the bilge keel chord. As the model
rolls to port, a clockwise (negative vorticity) rotating
vortex is created at the bilge keel tip. When the model
has fully rolled to port, the clockwise rotating vortex is
shed from the bilge keel, and a counter-rotating,
counter-clockwise (positive vorticity) rotating vortex is
created at the bilge keel tip and rolls up as the model
rolls to starboard. When the model is fully rolled to
starboard, the counter-clockwise rotating vortex is shed,
the clockwise rotating vortex is created at the bilge keel
tip and the cycle repeats. The vortices decay in both
size and strength with each successive roll cycle, such
that after 5 roll cycles, the vorticity is small and
concentrated in a small region near the bilge keel tip.
4.2.4

Figure 25 shows the phase-averaged wave


elevation perturbations of the first five roll cycles at
quarter phase angles for all zones, Fr = 0.138 and x4i =
-10. The figure is presented exactly the same as the
Figure 24. The wave perturbation is determined by
subtracting the steady flow wave pattern from the total
unsteady instantaneous phase-averaged wave pattern.
The wave elevation perturbations show more clearly the
forward moving waves at the shoulder of the model.
5

CONCLUSIONS AND FUTURE WORK

Results are presented for towing-tank experiments


performed for an advancing surface combatant in free
roll decay and coupled pitch and heave motions. For
free roll decay experiments, results are presented for
motions (surge, sway, heave, roll, pitch and yaw),
forces (resistance, sway and heave), moments (pitch
and yaw), phase-averaged velocities (U, V and W) for
measurement region near bilge keel and free surface
elevations. For coupled pitch and heave experiments,
results are presented for pitch and heave transfer
functions, and pitch and heave phase angles. The
geometry of interest is DTMB model 5512, which is a
1/46.6 scale geosym of DTMB model 5415 (DDG-51),
with L=3.048 m. The experiments are performed in a
3x3x100m towing tank equipped with a plunger-type
wavemaker.
The measurement systems include
Krypton contactless motion tracker, 4-component load
cell, towed 2-D particle image velocimetry (PIV)
system, and servo wave probes with 2-D traverse.
Uncertainty assessment following standard procedures
is used to evaluate the quality of the data.
Coupled pitch and heave results show that pitch
and heave transfer functions and phase angles collapse
to a single value independent of wave steepness due to
linear response of pitch and heave motions. Transfer
functions and phase angle results show 7% difference in
encounter frequency compared to results from SMP.
Free roll decay results show that the addition of
bilge keels to a ship model increases the roll period and
the roll damping. Roll decay coefficient results
compare well with previous experimental results for
trend and slope. Roll decay coefficient increases with
increasing Fr with a plateau for 0.190 Fr 0.340
before increasing again.
Roll decay coefficient
increases with increasing initial roll angle. The
addition of bilge keels reduces the roll frequency
magnitude (and consequently increases roll period) for
all Fr. The roll frequency data is clustered near the
natural period of roll for zero forward speed, Tn = 1.54
seconds or a natural frequency, fn = 0.649 Hz. Lognormal plots of the absolute value of the crest and
trough peak amplitudes show non-linear roll decay for
Fr 0.138 for both without and with bilge keels and
linear roll decay for Fr 0.190 for both without and
with bilge keels.

Wave-field

Figure 23a and 23b show the steady wave-field


and phase-averaged residuals (at t = 0) respectively for
zone 1, Fr = 0.138 and x4i = -10. The steady wavefield is reproduced using the wave elevation
measurements for time greater than 35 seconds (see
Figure 17a). The steady wave-field (Figure 23a) shows
the expected Kelvin wave pattern generated from an
advancing ship hull. Alternating peaks and troughs
radiating from the hull shoulder are evident. The
unsteady phase-averaged residuals in Figure 23b show
that the residuals for zone 1 are typically of the order of
<1%. Residuals locally are as high as 5% for the
measurement locations corresponding to the diverging
wave pattern. Residuals can be higher in this region
due to larger wave elevation gradients, which affect the
repeatability of the measurement.
Figure 24 shows the phase-averaged wave
elevations of the first five roll cycles at quarter phase
angles for all zones, Fr = 0.138 and x4i = -10. The
figure is presented exactly the same as the velocity
contours (see description for Figure 19). The wavefield resembles the steady wave pattern with a
superimposed oscillation due to the rolling motion of
the model. Focusing on the region near the shoulder of
the model for all roll cycles, we see in the first column,
with the model rolled to starboard, a wave trough is
developed at the shoulder followed aft by a wave crest.
As the model rolls to port, the trough and crest move
forward and the crest is now stationed at the model
shoulder when the model is rolled fully to port. The
wave crests and troughs dissipate forward of the model
as the model rolls. As model roll motion decays, the
amplitude of the generated crests and trough at the
shoulder decay.

12

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Force and moment data show that resistance


coefficient, heave force coefficient and pitch moment
coefficient data collapse to a single value independent
of initial roll angle for increasing forward speed. The
addition of bilge keels increases resistance and heave
coefficients while decreasing the pitch moment.
Phase-averaged flow-field velocity contours show
the evolution and subsequent decay of the bilge keel
vortex. The vortex trails the motion of the bilge keel.
The vortex rotates clockwise for counter-clockwise
model rotation (rolling to port) and rotates counterclockwise for clockwise model rotation (rolling to
starboard).
The bilge keel vortex indicates a
momentum deficit in the model boundary layer that
results in an increased drag or resistance on the model.
The bilge keel vortex decays with the decaying roll
motion.
The phase-averaged wave-field resembles the
steady wave pattern (Kelvin wave pattern) with a
superimposed oscillation due to the rolling motion of
the model. For the model rolled to port, a crest is
developed at the hull shoulder. As the model rolls to
starboard, the crest moves forward of the shoulder and a
trough is then developed at the shoulder. The crests
and troughs dissipate forward, toward the bow with
time. As model roll motion decays, the amplitude of
the generated crests and trough at the shoulder decay.
Future work includes continued analysis of the
present data. The final dataset will include complete
results for all test conditions for coupled pitch and
heave and free roll decay tests. The complete dataset
will be presented in Irvine (2004) and archived at
http://www.iihr.uiowa.edu/~towtank/ for the purpose of
CFD validation. In fact, validation and verification of
CFDSHIP-IOWA using some of the roll decay time
history data from the present study has been completed
Wilson et al. (2004a). Future testing at the IIHR
Hydroscience and Engineering towing tank includes
maneuvering tests with stereoscopic 3-D PIV to obtain
benchmark data for RANS-CFD validation and model
development.

REFERENCES
ASME, Test Uncertainty, ASME PTC 19.1-1998,
The American Society of Mechanical Engineers,
1998, 112 pp.
Gui, L., Longo, J., and Stern, F., Towing Tank PIV
Measurement System, Data and Uncertainty
Assessment for DTMB Model 5512, Experiments
in Fluids, Vol. 31, 2001a, pp. 336-346.
Gui., L., Longo, L., Metcalf, B., Shao, J., and Stern, F.,
Forces, Moment, and Wave Pattern for Surface
Combatant in Regular Head Waves-Part 1:
Measurement
Systems
and
Uncertainty
Assessment, Experiments in Fluids, Vol. 31,
2001b, pp. 674-680.
Gui., L., Longo, L., Metcalf, B., Shao, J., and Stern, F.,
Forces, Moment, and Wave Pattern for Naval
Combatant in Regular Head waves-Part 2:
Measurement
Results
and
Discussions,
Experiments in Fluids, Vol. 32, 2002, pp. 27-36.
Irvine, M., Towing-Tank Tests for Surface Combatant
for Free Roll Decay and Coupled Pitch and Heave
Motions, Ph.D. Dissertation, University of Iowa,
IIHR-Hydroscience and Engineering, expected
August 2004.
Journee, J. M. J., Experiments and Calculations on
Four Wigley Hull Forms. Technical University
Delft, Report No. 909, Delft, 1992.
Larsson, L., Stern, F., Bertram, V., editors, Proceedings
of the Gothenburg 2000: A Workshop on
Numerical Ship Hydrodynamics, Chalmers
University of Technology, Gothenburg, Sweden,
Sept. 2000.
Longo, J., Rhee, S.-H., Kuhl, D., Metcalf, B., Rose, R.,
and Stern, F., IIHR Towing-Tank Wavemaker,
Proceedings of the 25th ATTC, Iowa City, Iowa,
1998.
Longo, J., Shao, J., Irvine, M., and Stern, F., PhaseAveraged PIV for Surface Combatant in Regular
Head Waves, 24th Symposium on Naval
Hydrodynamics, Fukuoka, Japan, 8-13 July 2002
Longo, J., and Stern, F., Uncertainty Assessment for
Towing Tank Tests With Example for Surface
Combatant DTMB Model 5415, Journal of Ship
Research, 2004
Stahl, R., Ship Model Size Selection, Facilities, and
Notes on Experimental Techniques, Carderock
Division, Naval Surface Warfare Center Report
CRDKNSWC/HD-1448-01, 1995.
Weymouth, G., "RANS CFD Predictions of Pitch and
Heave Ship Motions in Head Seas," MS Thesis,
University of Iowa, Iowa Institute of Hydraulic
Research, 2002.

ACKNOWLEDGEMENTS
This research was sponsored by Office of Naval
Research under Grant N00014-01-1-0073 under the
administration of Dr. Pat Purtell whose support is
greatly appreciated. Special thanks are extended to
University of Iowa mechanical and civil engineering
undergraduates Pieter Beyer, Brad Carlson, Brad
Flaminio, Tanner Kuhl, and Adrian Stamper for their
efforts in the data-acquisition phases of this study.

13

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Wilson, R. and Stern, F., Unsteady RANS Simulation


of a Surface Combatant, Forward SpeedDiffraction Problem, presented at ONR 2000
Workshop on Free Surface Turbulence and Bubbly
Flows, California Institute of Technology,
Pasadena, California, March 1-3, 2000,
unpublished.
Wilson, R. and Stern, F., "Unsteady RANS Simulation
of a Surface Combatant with Roll Motion," 24th
Symposium on Naval Hydrodynamics, Fukuoka,
Japan, 8-13 July 2002.
Wilson, R., Carrica, P., and Stern, F., Unsteady RANS
Method with Application to Roll Motion for a
Surface Combatant, Computers and Fluids,
Submitted April 21, 2004a.
Wilson, R., Carrica, P., Hyman, M., and Stern, F., A
Steady and Unsteady Single-Phase Level Set
Method for Large Amplitude Ship Motions and
Maneuvering, Proceedings of the 25th Symposium
on
Naval
Hydrodynamics,
St.
Johns,
Newfoundland and Labrador, Canada, 8-13 August
2004b.

14

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Table 1. Summary of DTMB model 5512 and fullscale particulars.


DTMB 5512
3.048 m
0.386 m
0.133 m
2
1.371 m
0.506

Length, L
Beam, B
Draft, T
Wetted Surface Area, S
Block Coefficient, CB

Table 7.

Full-Scale
142.04 m
17.99 m
6.20 m
2
2977 m
0.506

BX (%)

PX (%)

UX (%)

Heave, x3

88.3

11.7

0.70

Roll, x4

60.8

39.2

0.64

Pitch, x5

61.2

31.8

0.81

Table 8.
Table 2.

Test conditions for coupled pitch and heave


motions tests.

Froude Number, Fr
0.0, 0.190, 0.280, 0.340,
0.410

Table 3.

Test conditions for free roll decay motion


tests.

Table 4.

Table 9.

Initial Roll Angle, x4i ()


0.0, 2.5, 3.0, 4.0, 5.0, 7.5,
10.0, 12.5, 15.0, 20.0

0.069, 0.096, 0.138, 0.190,


0.280, 0.340, 0.410

Wave-field elevation measurement grid


spacing.

Region
Near-Field (1)
Wake (2)
Wake (3)
Far-field (4)
Far-field (5)
Far-field (6)
Total

30.48 mm
30.48 mm
50.80 mm
50.80 mm
101.60 mm
101.60 mm

30.48 mm
30.48 mm
50.80 mm
50.80 mm
101.60 mm
101.60 mm

UA summary table for forces and moments


measurements.
BX (%)
1.1
0.2
0.3
3.5
0.3

CT
CS
CH
CM
CY

Wave Steepness, Ak
0.025, 0.050, 0.075

Froude Number, Fr

UA summary table for roll motion


measurements.

PX (%)
98.9
99.8
99.7
96.5
99.7

UX (%)
1.4
12.4
15.2
2.1
12.8

UA summary table for wave elevation


measurements.
Bnf (%)

Pnf (%)

Unf (%)

Position 1

0.30

99.70

5.0

Position 2

0.01

99.99

7.8

Table 10. Roll time history correction bias summary


table.

#
Points
1632
418
182
220
52
108
2812

Fr

Bias ( )

0.069

0.020

0.096

0.015

0.138

0.027

0.190

0.040

0.280

0.074

0.340

0.122

0.410

0.206

Table 5. Ballasted properties of DTMB 5512.


Parameter
VCG
Pitch Gyradius, k5
Roll Gyradius, k4

DTMB 5512
0.163 m
0.762 m or 0.25*L
0.149 m or 0.385*B

Table 11. Contribution of bare hull and bilge keel


damping to total roll damping.
% b44 (bare hull)

Table 6.

UA summary table for carriage speed


measurements.
Uc

BUc
93.3 %

PUc
6.7 %

% b44 (bilge
keels)

EFD

78.1

21.9

SMP

16.0

84.0

UUc
0.5 %

15

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Fig. 1.

Fig. 2.

DMTB model 5512, 1/46.6 scale geosym of


DDG-51 Arleigh Burke Class Destroyer.

Pitch transfer function vs. encounter frequency


for Fr = 0.28 and Ak = 0.025, 0.050, 0.075.

Fig. 6.

Roll time history for Fr = 0.138, initial roll


angle, x4i = -10, without and with bilge keels
showing effect of bilge keels on roll damping
and period.

Dantec 2-D towed PIV measurement system


shown in XY configuration, measuring region
near bilge keel of DTMB 5512.

Fig. 3.

Fig. 4.

Fig. 5.

Wave-field elevation measurement grid.

Flow-field measurement area near bilge keel,


showing PIV cuts at x = 0.675.

16

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Non-dimensional Roll Damping Coefficient, b44 (nd)

Circles = without bilge keels


Squares - with bilge keels
0.1

0.08

Damping Ratio,

x4i = 2.5
x4i = 3.0 o
o
x4i = 4.0
x4i = 5.0 o
o
x4i = 7.5
o
x4i = 10.0
o
x4i = 12.5
o
x4i = 15.0
x4i = 20.0 o
o
x4i = 2.5
o
x4i = 3.0
o
x4i = 4.0
o
x4i = 5.0
x4i = 7.5 o
o
x4i = 10.0
x4i = 12.5 o
o
x4i = 15.0
o
x4i = 20.0

0.06

0.04

0.02

0.2

0.4

Froude Number, Fr

Average damping ratio, versus Froude


number, Fr for seven Fr without and with
bilge keels, for all initial roll angles, x4i.

Fig. 7.

Circles = without bilge keels


Squares = with bilge keels

Fr =
Fr =
Fr =
Fr =
Fr =
Fr =
Fr =
Fr =
Fr =
Fr =
Fr =
Fr =
Fr =
Fr =

0.1

Damping Ratio,

0.08

0.06

0.04

Fr = 0.138

0.03

0.02

0.01

Fig. 9.

0.069
0.096
0.138
0.190
0.280
0.340
0.410
0.069
0.096
0.138
0.190
0.280
0.340
0.410

Total - EFD, b44


Bilge Keels - EFD, b44-BK
Bare Hull - EFD, b44-BH
Total - SMP, b44
Bilge Keels - SMP, b44-BK
Bare Hull - SMP, b44-BH

0.05

5
o

Mean Roll Angle, x4m ( )

Roll damping coefficient for the unappended


model (bare hull) b44-BH, the total roll damping
coefficient (bare hull + bilge keels) b44, and
the bilge keel roll damping coefficient b44-BK
versus mean roll angle, x4m for Fr = 0.138,
EFD and SMP results.

0.04

0.02

10

15

20

25

Initial Roll Angle, x4i ( )

Fig. 8.

Average damping ratio, versus initial roll


angle, x4i. for all seven Fr, without and with
bilge keels.
Fig. 10. Roll decay coefficient, n vs. mean roll angle
for 0.069 Fr 0.41 with free running model
data for DDG-51 from Bishop (1983).

17

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Fig. 11. Roll decay coefficient, n versus Fr, without


and with bilge keels.

Fig. 13. Log-normal of the absolute value of the crest


and trough roll amplitudes versus time for all
Fr and initial roll angle x4i = -10 (a) without
bilge keels. (b) with bilge keels.

Fig. 14. Resistance coefficient CT versus Fr for seven


Fr without and with bilge keels and all initial
roll angles.

Fig. 12. (a) Damped Frequency fd versus mean roll


angle x4M for all Fr, without and with bilge
keels. (b) Damped Frequency fd versus Fr,
without and with bilge keels.

18

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

10.0

(a)

x4( )

5.0

Roll signal (complete)

0.0

-5.0

-10.0
0.0

10.0

20.0

30.0

40.0

50.0

60.0

t (sec)
10.0

(b)
5.0

x4( )

Fig. 15. Heave force coefficient CH versus Fr for seven


Fr without and with bilge keels and all initial
roll angles.

0.0

-5.0
Roll signal (6 cycles)
Quarter phases for five roll cycles

-10.0
0.0

1.7

3.3

5.0

6.7

8.3

10.0

t (sec)

Fig. 17. Roll decay signal for one carriage run (a) and
truncated roll decay signal (b) where flowfield
and wavefield data are analyzed through five
roll cycles.

Fig. 16. Pitch moment coefficient CM versus Fr for


seven Fr without and with bilge keels and all
initial roll angles.

19

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Fig. 18. Contours of steady flow variables (a) U; (b) V;


and (c) W for model 5512 at Fr=0.138.

20

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Fig. 19. Contours of phase-averaged axial velocity (U) for five roll cycles at quarter phases for model 5512 at
Fr=0.138 and x4i=-10.

21

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Fig. 20. Contours of phase-averaged transverse velocity (V) for five roll cycles at quarter phases for model 5512 at
Fr=0.138 and x4i=-10.

22

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Fig. 21. Contours of phase-averaged vertical velocity (W) for five roll cycles at quarter phases for model 5512 at
Fr=0.138 and x4i=-10.

23

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Fig. 22. Contours of axial vorticity (x) for five roll cycles at quarter phases for model 5512 at Fr=0.138 and x4i=10.

24

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Fig. 23. Steady wave pattern (a) for all zones, for model 5512 at Fr=0.138 and typical unsteady residuals (b)
resulting from phase averaging free surface elevations at t=0 sec.

25

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Fig. 24. Phase-averaged wave elevations of the first five roll cycles at quarter phase angles for all zones, Fr=0.138 and x4i=-10.

26

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Fig. 25. Phase-averaged wave elevation perturbations of the first five roll cycles at quarter phase angles for all zones, Fr=0.138 and x4i=-10.

27

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics


St. Johns, Newfoundland and Labrador, CANADA, 8-13 August 2004

Viscous flow simulation past a ship in waves using the


SWENSE approach
R. Luquet, L. Gentaz, P. Ferrant, B. Alessandrini
(Ecole Centrale de Nantes, France)

1 ABSTRACT
This paper documents the development of the
SWENSE (Spectral Wave Explicit Navier-Stokes
Equations) approach, a new method for studying
wave-body interactions. This work has been motivated
by the accuracy and efficiency requirements for
simulating hulls manoeuvring in waves. The
SWENSE method is applied to the case of a ship kept
fixed at the dynamic trimmed and heave conditions in
heading regular waves with forward speed. After a
description of the theory, the test case is presented :
the DTMB Model 5415 has been chosen because of
the availability of experimental data. Unsteady
resistance, heave force, pitch moment, nominal wake
and free-surface elevations are compared to
experiments (Gui et al., 2001). Results show an
overall good agreement with experiments. The
accuracy and effectiveness confirm the viability of the
method. Finally, future developments of the proposed
approach are discussed.

2 INTRODUCTION
Free surface flow problems in the naval or
ocean engineering contexts are more and more often
addressed using viscous flow solvers based on the
solution of Reynolds Averaged Navier-Stokes
Equations (RANSE). This is especially useful in
situations where viscosity or flow separation effects
play an important role. A number of useful codes have
been developed and used by ship designers, and
computational fluid dynamics (CFD) techniques have
been recently incorporated into non linear hull shape
optimisation procedures. In this way, CFD simulation
now plays an important role in ship design.
The classical method used to simulate a ship

advancing in head waves is to impose an incident


wave field on the inlet boundary. It is modelled as
velocity and pressure perturbations which are added to
the uniform stream. Those perturbations are derived
from the linear or nonlinear potential flow solution for
free-surface travelling waves.
However such simulations require very large
computing resources because grids must be very
refined between the location of the structure and the
location of the wave generation systems (for
structured or non-structured grids used with finite
differences or finite volumes). This is indeed
necessary to propagate the wave from the paddle to
the structure with no noticeable damping. Moreover
successive wave reflections on the body or the paddle
affect the incoming wave train and reduce the useable
duration of the numerical simulation.
To overcome these difficulties an original
formulation is used here by modifying the initial
problem in order to solve the diffracted flow only.
This approach has previously been used in the frame
of potential theory, by Di Mascio et al. (1994) or
Ferrant (1996) in three dimensions. It consists in
splitting all unknowns of the problem (potential and
free-surface elevation) into the sum of an incident
term and a diffracted term. The incident terms are
described explicitly using a nonlinear potential flow
model. Thus only the part of the grid in the vicinity of
the structure needs to be refined. Far from the body a
stretched grid allows an efficient damping of the
diffracted flow.
Here this splitting of unknowns will be
applied to a 3D viscous flow solver (Alessandrini &
Delhommeau, 1995). The incident flow is supposed to
verify a non-linear potential problem and the
diffracted flow is solved by assuming that the total
flow verifies RANS Equations : modified RANS

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Equations verified by the diffracted flow are named in


the following SWENSE (Spectral Wave Explicit
Navier-Stokes Equations).
This technique has been already successfully
applied in 2D for the case of a non-linear regular wave
train on submerged horizontal square and circular
cylinders (with axis parallel to the wave crest),
(Ferrant et al., 2003) and (Luquet et al., 2003)
respectively. In the first case, hydrodynamic forces
were compared with numerical and experimental data.
In the second, free-surface elevations and its harmonic
components on lee the side of the cylinder were
validated. A first application in 3D, has been done for
a vertical cylinder in waves and has given very
accurate results on forces and wave runups comparing
to experiments (Gentaz et al., 2004). The numerical
procedure has been further developed to account for
forward speed. This paper is devoted to the simulation
of a ship advancing in various wave conditions.
The paper is organized as follows. After a
description of the principles of the method, governing
equations, boundary and initial conditions and
numerical method are presented. Then model,
computational parameters and grids are described
along with available data. A grid dependency test is
realized. Numerical results (unsteady resistance, heave
force, pitch moment, nominal wake and free-surface
elevations) are compared to experiments (Gui et al.,
2001) for different Froude numbers and wave
conditions. Finally, some conclusions are made and
future developments are discussed.

3 THEORY
In this section, the numerical method is
described. After defining the coordinate system,
governing equations (previous RANSE and modified
SWENSE) are presented and followed by boundary
conditions (previous and modified). The model used
for the incident flow is presented and the numerical
method for solving the whole set of equations is
presented.

3.1 Coordinate system


x3

x1

x2

A space-fixed coordinate system (o- x1 x 2 x3 ),


see Figure 1, is defined with the (o- x1 x2 ) plane on the
still free surface and the zaxis directed vertically
upwards. The origin of the coordinate system is placed
at the intersection of the undisturbed free surface and
forward perpendicular (FP) of the model. The axes of
the coordinate system ( x1 x 2 x3 ) are normalized with
model length and the x1 -axis is directed downstream.
Another coordinate system ( 1 , 2 , 3 ), fitted to the hull
and free surface field is used to define the curvilinear
computational domain.

3.2 Governing Equations


The set of equations governing the diffracted
flow are obtained from the classical RANS Equations
of the 3D solver (Alessandrini et al., 1995). The set of
equations solved by the initial solver is first given.
Then SWENS Equations are derived.
3.2.1 Previous RANS Equations
The SWENSE approach is applied here on a
viscous flow solver which uses 3D RANS Equations
under fully non-linear free-surface boundary
conditions. In Cartesian form, the unsteady
incompressible RANS and continuity equations are
written as
U i
=0
x i

(1)

U i
U i
2U i
i j
1 P
+U j
=
+

uu
t
x j
xi
x j xi x j

(2)

Where
components, xi ,

Ui

, i {1,2,3} are the mean-velocity


i {1,2,3} are the Cartesian coordinates,
i
P = p + g x 3 is the dynamic pressure, u , i {1,2,3} are

the fluctuant velocities, u i u j the Reynolds stresses


and the kinematic viscosity. Reynolds stresses are
related to the mean rate of strain through an isotropic
eddy viscosity t :
U i U j
+
ui u j = t
x j
U i

2
+ ij k
3

(3)

Where ij is the Kronecker delta and k the

Figure 1: Definition of the coordinate system

turbulent kinetic energy. The equations are normalized


by reference velocity U , Length between
perpendiculars of the model L and density .
Substituting (3) for the Reynolds-stress term

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

in (2), the momentum equations become :


1 2U i
U i
U i
P
+U j
=
+
+Ti
t
x j
xi Re eff x j xi

vg =

U i U j
T i = t
+
x j
U i

Eddy viscosity and turbulent kinetic energy


are calculated by a k- turbulence model.
The equations are then transformed from the
physical domain in Cartesian coordinates (xi ) into the
computational domain in non-orthogonal curvilinear
coordinates ( i ) by using a partial transformation.
Details of the coordinate transformation can be found
in Alessandrini et al. (1995). The vectors of the both
dual basis, covariant and contra-variant, used to
transform the equations from the cartesian to the
curvilinear coordinates are respectively expressed as
following :

xi
are the grid velocities.
t

The previous set of RANS Equations is


modified in order to formulate a problem for the
nonlinear diffracted flow. Primitive unknowns
(Cartesian components of velocity U i , i {1,2,3} ,
pressure P and free-surface elevation h ) are
decomposed as follows :

1
1
=
+ t
Re eff Re

(a ) ,

1
Jg ji are the control grid functions,
J i

3.2.2 SWENS Equations

2
P = P + k is the dynamic turbulent pressure
3

ij

(4)

Where :

(a ) ,

i {1,2,3}

j {1,2,3}

i {1,2,3}

j {1,2,3}

is the covariant matrix

is the contra-variant matrix

(5)

(6)

Thus, the continuity (1) and momentum (2)


equations in the transformed space are given by :
U i
a
=0
k

(7)

k
i

U
U j i
1

+ ai U vg i
f j aki ti akj

j
Reeff
t

k
2
1
k P
ij U
i t
j U
g
a
= a
+
+ ak
k Reeff
j i
i
j

U i = U Ii + U Di

P = PI + PD
h = h + h
I
D

Variables with the subscripts I and D


represent respectively incident and diffracted
variables.
This decomposition is then introduced in the
set of initial RANS equations in Cartesian coordinates
assuming that the incident wave flow fulfils Euler
equations:
U Di
=0
xi

g ij = aki akj is the contra-variant metric tensor,

(10)

U Di j
U D
2U D 1 PD
( + t )
+
+ U I + U Dj tj
2
t
x x j
x
x j

= U Dj

U I
2U I t
+ ( + t )
+ j
U Ij + U Dj
2
x j
x x
x j

aik

U Di
=0
k

(11)

(12)

U D j j
t j U D
1
i
f j a ki
ak
+ ai U I + U Dj v g

Reeff
t
i j

2U D
P
1
g kk

+ ak D
2
Reeff
k
k

(8)

The equations are then normalized and


transformed into the computational domain :

Where :

(9)

2U D
1
g jk
Reeff
j k

+ a ki
k l

t j
a
U Ik + U Dk
i
j

U I k j
1 jk 2U I
a j U D a ij ti
g
+
+fj

Reeff
j k
j

Copyright National Academy of Sciences. All rights reserved.

U I

(13)

Twenty-Fifth Symposium on Naval Hydrodynamics

We obtain a new set of equations called


SWENS Equations where incident variables (dynamic
pressure, velocities, free-surface elevations and their
gradients) are explicit. Their values are directly
computed at each time step knowing kinematics and
interface position of the incident flow. Then, only the
diffracted variables are unknowns of the problem and
have to be solved by the modified viscous solver.
To conclude, the solution obtained is the
solution of a modified problem where incident terms
are known and their values are replaced in the
SWENS Equations. There are no specific assumptions
and the diffracted solution summed with the incident
field give the solution of the initial problem.

3.3 Boundary Conditions


The boundary conditions are composed of
no-slip and free surface conditions (kinematics and
dynamic). In the first part, the previous conditions are
given and then the modified boundary conditions
corresponding to the SWENS Equations are presented.

n j t1i + ni t1 j

n t +nt
i 2j
j 2i

) Ux

(
(

=0

(16)

U
=0
x j

Where (t1i ) , (t 2i ) and (n j ) j {1,2,3} are


respectively the components of the two vectors
tangential and the one normal to the free surface at the
point where the free surface condition is applied.
All of these free surface boundary conditions
are applied on the actual free surface, which is
updated at each time step.
A classical no-slip condition is imposed on
the hull and also in the farfield. Even if the no-slip
condition on the farfield seems to be unphysical, a
stretching of the mesh is realized in order to damp the
diffracted waves far from the body, so there are no
reflexions. This condition is written as follows :
(17)

Ui = 0

3.3.1 Previous Boundary Conditions

3.3.2 Modified Boundary Conditions

The previous free surface boundary


conditions were composed of four equations. The
classical kinematic condition expresses the fact that a
fluid particle of the free surface remains on it.
Especially, the vertical variation of the position of a
particle is equal to the vertical velocity of the fluid :

The previous set of boundary conditions is


modified in order to define the diffracted problem.
The previous decomposition (7) is inserted in the
equations and the modified free surface boundary
conditions are then respectively written as :

h
h
h
+U1
+U 2
=U3
t
x1
x 2

(14)

P = g h 2 ( + t )

U
ni n j
x j

1 hD hI
U D +

x + x
2

h h
h
= U I D + I
t x1
x1
3
I

The dynamic conditions verifies the


continuity of stresses applied to the fluid through the
free surface. Then, this continuity condition is written
on the three axes. Neglecting the superficial tension
term (which has a weak influence on the type of
problems treated here) and using the dynamic
turbulent pressure definition, the continuity of the
normal component gives the normal dynamic
condition :
i

hD hD hI
+
+
t
x1
x1

(15)

And the two tangential components give the


two tangential dynamic conditions :

1 hD hI
U I

x + x
2

PD g hD = g hI PI 2 ( + t )

n j t1i + ni t1 j

n t +nt
i 2j
j 2i

i
D
j

) Ux

) Ux

i
D
j

2
U D U D3

i
I
j

) Ux

) Ux

= n j t 2 i + ni t 2 j

U Ii + U Di
ni n j (19)
x j

= n j t1i + ni t1 j

(18)

2
U I

i
I
j

(20)

The no-slip boundary condition generates the


diffracted field from the hull :
U Di = U Ii

Copyright National Academy of Sciences. All rights reserved.

(21)

Twenty-Fifth Symposium on Naval Hydrodynamics

The modified condition applied on the


farfield verifies the decay of the diffracted field far
from the body in the same way as previously :
(22)

U Di = 0

All these boundary conditions are then


expressed in curvilinear coordinates. So the kinematic
condition is finally written as :
h
hD
h
+ a1j D + I
j j
t

U D + a 2j hD + hI

j j

h
h
h
1
= U I3 I a1j D + I U I1 v g
j j
t

h
h
2
a 2j D + I U I2 v g
j j

2
U D U D3

(23)

and contra-variant a i = (a ij ) vectors as following :

nj =

a ij
ai
a 3j
a3

i {1,2}

j {1,2,3}

, j {1,2,3}

(25)

Then, the normal and tangential dynamic


conditions respectively become :

= g hI PI 2 ( + t )

a i3 a 3j a kj U Ii + U Di
2
j
a3

(26)

k 3
U Di
U Ii
k 3
k 3
k 3
a
a
a
a
a
a
=

+
a j a j a1i + a j a i a1 j
j
j
i
j
i
j
1
1

k
x j

i
i
n t + n t U D = n t + n t U I
i 2j
j 2i
i 2j
j 2i
x j
x j

(
(

( )

( (

))

(28)

And the free surface elevation by :


(24)

PD g hD

i (x1 , x 2 , t )
(x1 , x 2 , t ) =

i =0

cosh ikx 2

i x1 , x 2 , t = Bi cosh (ikD ) sin ik x1 ct

The components of the two vectors tangential


and the one normal to the free surface can be
respectively expressed using the covariant ai = (aij )

t ij =

pressure, free surface elevations).


In this study, we consider a non-linear
regular incident wave train. To compute it, an
algorithm based on the stream function theory
(Rienecker & Fenton, 1981) has been implemented.
This algorithm has been chosen as it can
generate the incident field for a wide range of depth,
amplitudes and wavelengths (in the limit of breaking
waves).
It is based on potential theory and gives the
numerical solution of steadily progressing periodic
waves on irrotational flow over a horizontal bed. No
analytical approximations are made. A finite Fourier
series is used to give a set of nonlinear equations
which can be solved using Newtons method. The
velocity potential is then given by :

(27)

The no-slip and farfield conditions still be the


same as previously.

3.4 Incident Flow Model


In order to apply the SWENSE method it is
necessary to be able to know at each time step all of
the characteristics of the incident field (velocities,

(x1 , x 2 , t ) =

A0
+
2

A cos(ik (x ct ))
i

(29)

i =1

This method allows the incident waves to be


described very accurately. What is striking is that this
incident field is independent of the Navier-Stokes grid
and can be computed everywhere. This attribute is
essential because values of the incident field will be
possibly needed above the undisturbed incident free
surface. The total free surface elevation can indeed be
higher than the incident elevation. That is why this
point is decisive for the choice of the incident field
model. What is more, the field created by this method
is infinitely derivable on the whole space and this
smoothness is useful for the implementation of the
equations and the behavior of the computations
The problem being defined in a frame of
reference linked to the ship, the forward speed is
accounted for by adding a uniform opposing current to
the incident wave field.
An irregular wave train could be prescribed
as well by using a spectral formulation. This kind of
procedure has been already developed by Ferrant &
Le Touz (2002) combining a spectral formulation
and a Boundary Element Method (BEM) to simulate
an irregular wave train interacting with a 3D body
under potential flow theory.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

4 RESOLUTION
This section gives some information about
the numerical method applied to obtain the solution of
the previous set of equations. Then an overview of the
organization of the SWENSE resolution is presented.

4.1 Numerical Method

Finally the combination of incident potential


flow and RANS solvers for the diffracted problem can
be summarized as follows : At each time step of the
computation the geometry of the fluid domain is
updated. The kinematics of the incident wave flow is
calculated on this updated grid and then the diffracted
problem defined by the set of SWENS Equations (10)

Incident flow at t
(potential model)

Calculation of the diffracted flow


by the modified RANSE solver

Diffracted flow at t

Iteration on time

4.2 Global organization

Geometry of fluid domain for


the total flow at t-1

Iteration on the
nonlinearities

The physical problem is now modelised by


the previous equations. In order to solve this problem,
the set of equations is computed and solved
numerically. The equations have been modified by
using a set of curvilinear space variables fitted to the
geometry of the fluid domain so it simplifies the
implementation of the boundary conditions. Equations
are also discretised by second-order finite difference
schemes in space and time. These equations are
discretised on a structured monoblock grid. An O-O
grid system is used in our study but an H-H grid could
be used as well. The advantages of the O-O grid will
be discussed further. A free-surface tracking method is
applied to update the interface : at each time step the
mesh is regridded following the new shape of the free
surface. A pressure equation is obtained by combining
the equation of mass conservation with transport
equations following the Rhie & Chow (1983)
procedure : thus checkerboard oscillations classically
associated with centred second-order schemes are
suppressed.
As mentioned previously, the turbulent
viscosity is computed at each time step by a modified
k- model proposed by Wilcox (1988) which
introduces a specific dissipation rate.
An original fully-coupled method, developed
by Alessandrini and Delhommeau (1995) is used to
solve the discretised set of equations : at each iteration
all equations (RANS Equations, pressure equation,
free-surface boundary conditions, no-slip conditions)
are assembled in a single large and sparse linear
system which is solved using a bi-CGSTAB algorithm
with an incomplete LU preconditioning. This method
offers a better accuracy and efficiency than weak
coupled algorithms (SIMPLER, PISO), especially for
the convergence of the velocity-pression coupling.

to (20) is solved using the viscous flow software


described previously. The diffracted solution is then
summed with the incident terms in order to obtain the
total field. Considering the total free surface elevation
it is now possible to update the fluid domain again.

Summation of the flows


total flow at t

Figure 2: Organization of the calculation

5 RESULTS AND DISCUSSIONS


To verify the efficiency of the SWENSE
method on the forward speed diffraction problem,
computed results (forces and moment, nominal wake,
free surface elevations) are compared with
experimental data (Gui et al., 2001). The experiments
were realized at the IIHR Institute, Iowa. All
experimental results used here are available online on
the IIHR website (www.iihr.uiowa.edu).

5.1 Validation Test Case


In this section, geometries, simulation
conditions and EFD validation data, and numerical
grids will be described
5.1.1 Geometry
The geometry selected for the validation of
the SWENSE method is the DTMB (David Taylor
Model Basin) Model 5415, representative of a modern
hull. It was conceived as a preliminary design for a
Navy surface combatant in 1980. The hull geometry
includes both a sonar dome and transom stern. This
model has been chosen for two main reasons.
First, there is a large EFD database for Model
5415 due to a current international collaborative study
on EFD/CFD and uncertainty assessment between
Iowa Institute of Hydraulic Research (IIHR), Istituto
Nazionale per Studi ed Esperienze di Architettura
Navale (INSEAN) and the Naval Surface Warfare

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Center, Carderock Division (NSWC, formerly


DTMB). The validation data includes boundary layer
and wake, longitudinal wave cuts, bow and transom
wave fields, and wave breaking. No full-scale data
and/or ship exist.
Moreover, this model has been also chosen as
a test case for the next CFD workshop in Tokyo, 2005.
The characteristics of the test are towing conditions in
incident waves (diffracted problem) with trim and
heave kept fixed (designated).
The DTMB Models 5415 and 5512 global
characteristics are given in Table 1.
PARAMETERS

UNITS

DTMB 5415

DTMB 5512

Linear scale ratio

28.8316

46.6000

Length between perpendiculars

5.72

3.048

Beam B

0.7242

0.3859

Draft T

0.4020

0.2142

Wetted surface area S

4.8273

1.3707

Waterplane area Awp

3.2228

0.9151

0.8426

0.1275

0.5060

0.5060

Displacement S

Block coefficient

Table 1 : Geometric parameters for DTMB models 5415 and 5512


for the static condition.

The two geometries are identical and only


differs in the scale.

Fn =

Ak =

(30)

gL
2A

(31)

Where U and L are respectively the forward speed


and the length of the ship, g is the local gravity
acceleration and A is the amplitude of the incident
waves.
The whole conditions used for the validation
are given in Table 3. The frequency of the regular
head wave f w and the frequency of the encounter
wave f e , also shown in Table 2, are determined as :
fw =

g
2

fe = fw +

(32)

(33)

For all Froude numbers, computations were


also run without waves Ak = 0 , (steady case).

( )

Fn

Re 10 6

Ak

fw

fe

0.19

3.153

0, 0.025

1.5

0.584

0.811

0.28

4.647

0, 0.025, 0.05

0.34

5.642

0, 0.025

0.5
1
1.5
1.5

1.012
0.715
0.584
0.584

2.016
1.218
0.919
0.991

5.1.2 Conditions
Test conditions applied here are the same as
the IIHR experiments. The tests were conducted on
the DTMB Model 5512 without appendages. The
transition to turbulent flow is initiated at x=0.05 (by
studs in experiments), and the ship is kept fixed at the
dynamic sinkage and trim condition, which are
determined in calm water condition (without waves)
for each Froude number, see table 2 (experimental
conditions).
Fn

FP (m)

AP (m)

0.19

-0.004

-0.0004

0.28

-0.0094

-0.0022

0.34

-0.0122

-0.0058

Table 2 : Dynamic sinkage and trim

Then, each computation is determined by


three main parameters : the Froude number (Fn ) , the
wavelength ( ) and the wave steepness (Ak ) . Fn and
Ak are defined as :

Table 3 : Test conditions

For the calculations, 15 coefficients are


retained in order to recompose the incident field
(nonlinear regular waves). The initial value of the
kinetic energy of the turbulence model is 10 8 , and the
transition to turbulent flow is realized at x=0.05 . The
time step is t = 2.10 2 seconds and simulations are run
for fifteen wave periods. A computation takes one
hour per wave period (for L = 1 ) on one processor of
1.5 GHz for grid 1 (94500 nodes) two hours for grid 2
(212784 nodes) and almost twelve hours for grid 3
(884802 nodes). To converge to the fully nonlinear
solution, nonlinear iterations are performed at each
time
iteration.
Undisturbed
incident
wave
characteristics are used to define initial conditions of
the calculation (at t=0, the free surface is fully
deformed). The ship is accelerated during 80 time
steps to its nominal velocity.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

5.1.3 Grid

C M (t ) =

T (t ) =

M Y (t )
0.5 U 2 SL

(36)

z ( x, y , t )
L

(37)

where FX , FY , M Y and z are the measured timevarying resistance, heave force, pitch moment and
free-surface elevation, respectively. S is the wetted
surface area for the static condition and is the water
density.
The previous variables are compared to
experiments in terms of Fourier components (FS).
This decomposition is realized in a moving window of
one encountered wave period long. The FS for time
history X ( X = CT , C H , C M , T ) is determined as
follows :
Figure 3 : view of the O-O grid system

X (t ) =

As the problem is symmetrical, the


computation is only realized on a half domain, Figure
3 shows a typical grid system. The O-O organization
is used in this study for a better control of the density
of the mesh in the boundary layer. As seen in Figure
3, the mesh is stretched in the farfield. This stretching
combined to a canceling of the pressure is sufficient to
realize an accurate damping of all of the diffracted
terms (velocities, pressure and free surface elevations)
in the outer part of the domain.
Body-fitted, structured, monoblock grids are
generated using the commercial grid generation
software GRIDGEN from Pointwise, Inc. Near-wall
spacing is determined by turbulence modeling
considerations where, in general, the first point from
the no-slip surface is placed at a distance of
d / L < 104 .

X0
+
2
2

cos(2nf e t + n )

(38)

n =1

X n = a n + bn

bn

an

n = tan 1

(39)
(40)

where X is the reconstructed time history; X n is the


n th order harmonic amplitude; n is the corresponding

phase. a n and bn are the classical Fourier coefficients.

5.1.4 Variables of interest


We will focus on the available experimental
variables. The first variables of interest are resistance
coefficient (CT ) , heave force coefficient (C H ) , pitch
moment coefficient (C M ) and non-dimensional free
surface elevation ( T ) , which are defined as :
CT (t ) =

FX (t )
0.5 U 2 S

(34)

C H (t ) =

FZ (t )
0.5 U 2 S

(35)

Figure 4 : Convergence of the Fourier components of the resistant


coefficient.

Figure 4 shows the history of the Fourier


components of the resistance coefficient for a classical

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

computation. As seen, the convergence of the


coefficient is effective after two or three time units
which corresponds to almost three encountered wave
periods.
Another variable interesting is the unsteady
perturbation response of unsteady free surface
elevation defined as:
(41)

z D = zT zT z I

Where zT is the time mean of the unsteady free


surface elevation and z I the incident free surface
elevation.

5.2 Grid Dependency Test


GRIDS
1
2
3

POINTS
90 x42 x 25 = 94 500
124 x 39 x 44 = 212 784
201 x 62 x 71 = 884 802
Table 4 : Size of the computational grids

To check out the grid dependency of the


SWENSE method, a numerical test was conducted
using a set of three grids. The number of grid points
are given in Table 4. The three grids used in the test
are shown in Figure 4. The simulations were run in
Fn = 0.28
and
regular
heading
waves
at
6
Re = 4.647 x10 . The wave-length to ship-length ratio
and the wave steepness were respectively chosen as
L = 1.5 and Ak = 0.025 .
Then, grid convergence is addressed by
performing steady and unsteady (for the case
described previously) simulations using the three
computational grids. Ship force coefficients are given
in Table 5 along with the absolute change, , between
the medium/coarse grid, fine/medium grid and
fine/experiments solutions (expressed in percentages).
The results show that the resistant coefficient,
heave force coefficient and pitch moment coefficient
are grid convergent. In the following comparisons,
results are given for computations realized on Grid 2
which presents a good compromise between accuracy
and computing time.

Figure5 : Comparison of the inner grids used for the examination of


grid-dependency, (successively grid 1, grid 2 and grid 3).

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

UNSTEADY
Grid

Grid 1

Grid 2

Grid 3

Exp.

CT ,0

0.00965

0.0092

0.00938

0.00925

4.3

2.1

1.4

0.0069

0.0071

0.0072

0.00608

2.8

1.4

18.2

-0.067

-0.064

-0.065

-0.068

4.7

1.5

4.4

0.041

0.038

0.04

0.036

7.9

11.1

-0.0019

-0.0014

-0.0016

-0.0012

( %)
CT ,1

( %)

C H ,0
( %)

CH ,1
( %)
C M ,0

( %)
CM ,1

0.012

( %)

35.7

12.5

33.3

0.01

0.0105

0.0108

20

4.8

2.7

Figure 6 : Numerical (up) and experimental (down) results on time


histories of resistant coefficient.

STEADY
Grid

Grid 1

Grid 2

Grid 3

Exp.

CT

0.0089

0.0083

0.0084

0.00884

(%)
CH

6.7

1.2

4.9

-0.067

-0.064

-0.065

-0.0633

( %)
CM

4.5

1.5

2.7

-0.001

-0.0017

-0.0015

-0.00187

41

11

19.7

( %)

Table 5 : Grid convergence of ship forces coefficients.

5.3 Forces and Moment


Time histories and Fourier components of
forces and moment coefficients are compared with
experiments for many test cases in the following
section.

Figure 7 : Numerical (up) and experimental (down) results on time


histories of heave force coefficient.

5.3.1 Time History


The first results for the forces and moment
tests are time histories of resistance (CT ) , heave force
(C H ) and pitch moment (C M ) for a test case at
Fn = 0.28 , L = 1 and Ak = 0.05 .
Figures 6,7 and 8 show a good agreement on
forces and moment amplitudes : only the resistant
coefficient (CT ) seems to have a slightly higher
amplitude than experiments. Each coefficient is
accurately in phase with experiments. The two curves
for each coefficient on the experimental results
correspond to Raw time history and FS reconstruction.
A FS reconstruction has been realized because (CT )
and (C H ) contain limited high-frequency signals at the
peaks and thoughts, which are associated with carriage
vibration. This noise is absent for (C M ) due to the
large inertia of the model for pitching motion.

Figure 8 : Numerical (up) and experimental (down) results on time


histories of pitch moment coefficient.

5.3.2 Fourier components


Then, harmonic amplitudes of the forces and
moment coefficients are compared in the following
figures. The characteristics of the incident wave is
L = 1.5 and Ak = 0.025 . The results are given for
three Froude numbers : Fn = 0.19, 0.28, 0.34 . Steady
components are given in Figure 9, and first harmonic

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

component in Figure 10, for resistance, heave and


pitch components. The agreement with experiments is
remarkable for heave and pitch components for both
steady and first harmonic components. Nevertheless,
there is a noticeable overestimation of the first
harmonic component of the resistance coefficient, for
the three Froude numbers.

Figure 10 : First harmonic amplitude of the resistance C T , 0 ,

heave force C H , 0 and pitch moment C M , 0 coefficients for test


cases at L = 1 Ak = 0.05 and Fn = 0.19, 0.28, 0.34 .

Figure 9 : Fundamental amplitude of the resistance C T , 0 , heave

force C H , 0 and pitch moment C M , 0 coefficients for test cases at

L = 1 Ak = 0.05 and Fn = 0.19, 0.28, 0.34 .

5.4 Nominal wake


In this section, Fourier components of
velocities, axial (U ) , transverse (V ) and vertical (W ) ,
are compared to experimental 2D PIV results for a test
case at L = 1.5 , Ak = 0.025 and Fn = 0.28 . Contours
are given in the propeller plane, x / L = 0.935 . Some

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

differences between experimental and numerical


results can be attributed to the fact that the density of
PIV measurements is of 22500 points although the
numerical grid can give no more than 3000
measurements. That is why high gradient can not be
plotted exactly.

Figure 12 : First harmonic amplitude of the axial velocity (U 1 ) ,

transverse velocity (V1 ) and vertical velocity (W1 ) for L = 1.5 ,

Ak = 0.025 and Fn = 0.28 . Left : experiment, Right : numeric.

Figure 11 : Fundamental amplitude of the axial velocity (U 0 ) ,


transverse velocity (V0 ) and vertical velocity (W0 ) for

L = 1.5 , Ak = 0.025 and Fn = 0.28 at the nominal wake


plane. Left : experiment, Right : numeric.

The steady components of the velocity, see


Figure 11, are in good agreement with experiments
and especially for the vertical velocity. First harmonic
amplitudes shown in Figure 12 show an overall
correct agreement with experiments. Moreover some
discrepancies appear, which may safely be attributed
to the rather low resolution of the computation (200
000 nodes).

5.5 Free Surface Elevations


In this section, total and diffracted wave

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

patterns are compared to experiments, comparisons


are presented for four different time instants in a wave
period, and also for the first harmonic of the wave
elevation. All results given correspond to L = 1.5 ,
Ak = 0.025 and Fn = 0.28 .
5.5.1 Total wave pattern

t/T=0.

Figure 13 shows the total unsteady wave


obtained by numerical simulation at
t / T = 0, 0.25, 0.5, 0.75 and the same results obtained
by experiments. The overall behavior of the
experimental results is correctly recovered, but short
wavelength components, associated with the steady
wave pattern, are clearly damped in the outer part of
the computational domain. This is clearly due to an
insufficient mesh density. Nevertheless the SWENSE
method is able to capture the unsteady Kelvin wave
pattern

pattern

A comparison on the first harmonic


amplitude of the total unsteady free surface elevation
is shown on figure 14. The experiments appears on the
lower half while the numerical results are in the upper
half of the figure. There is a good agreement on
distribution and only the amplitude seems to be
slightly underestimated far from the hull.
t/T=0.25

t/T=0.5

t/T=0.75
Figure 14 :Oth and 1st harmonic amplitude for the total unsteady
free surface elevation. Up : numeric, Down : experiment.

Figure 13 : Numerical (up) and experimental (down) total unsteady


wave pattern at t / T = 0, 0.25, 0.5, 0.75

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

5.5.2 Diffracted wave pattern

t/T=0.

t/T=0.25

numerical simulation at t / T = 0, 0.25, 0.5, 0.75 and the


same results obtained by experiments. Again, the
overall agreement is good. Differences in details are
noticeable. Here again the origin of these
discrepancies is probably in the too coarse
discretisation of the free surface, which affects the
estimation of the diffracted pattern in two different
ways.
The direct resolution effect is augmented by
the fact that the steady wave resistance pattern is
substracted from the unsteady pattern together with
the incident wave field, in order to extract the
diffracted pattern in the same way as in the
experiments. This affects the estimation of the steady
part of the diffracted pattern, since the steady wave
resistance pattern is not sufficiently accurately
estimated in the calculation.
This analysis seems to be confirmed by Figure 16 in
which the unsteady first harmonic components of the
diffracted patterns are compared. The agreement is
much more satisfactory, since the influence of the
differences appearing in the estimation of steady
components is eliminated.

t/T=0.5

Figure 16 : First harmonic amplitude for the diffracted unsteady


free surface elevation. Up : numeric, Down : experiment.

t/T=0.75

5.6 Perspective view


Figure 17 show a 3D perspective of the
instantaneous wave field and hull for the DTMB
model 5415 in head waves with forward speed. The
free surface mesh is colored by free surface elevation.
The upper drawing corresponds to the diffracted field
( D = T I ) and the lower to the total field. Results
given correspond to L = 1 , Ak = 0.05 and Fn = 0.28 .
Figure 15 : Numerical (up) and experimental (down) diffracted
unsteady wave pattern at t / T = 0, 0.25, 0.5, 0.75

Figure 15 shows the diffracted unsteady


wave pattern (as defined by equation 34) obtained by

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

hours per encountered wave length. Finally, the


simulations presented in this paper demonstrate the
effectiveness of the SWENSE approach for ship /
wave interaction.

6.2 Future work

Figure 17 : Perspective view of the DTMB 5415 in head waves with


forward speed. Up : diffracted free surface elevation, Down : total
free surface elevation

6 CONCLUSION AND FUTURE


WORK
6.1 Conclusion
The SWENSE new simulation method has
been presented for the prediction of the viscous flow
past a ship in regular heading waves. The accuracy
and efficiency of this method was examined by
simulations with a code-development benchmark
geometry, DTMB model 5415. Unsteady flow
verification and validation was presented by
comparisons to experiments made by Gui et al.
(2001). Many quantities as forces and moment
coefficients, free surface elevations and nominal wake
velocities were compared both on time histories and
Fourier components. The results give a good overall
agreement which confirms the viability of the method.
Especially, force coefficients and unsteady
components are estimated with a very satisfactory
accuracy. Results on steady components, have to be
improved, primarily by using finer grids. Such results
will be presented at the symposium.
What is more, the duration of the simulation
is considerably reduced by the use of the SWENSE
method. Indeed, a classical simulation takes only two

A work on the implementation of the


modified free surface conditions will be realized to
improve the accuracy of the results, especially for free
surface
elevations.
In
addition,
additional
complications that could arise due to the shallow draft
transom which could possibly become unwetted when
simulating large amplitude incident waves will be
treated.
Although the simulations dealt with in this
study are limited to those in regular heading waves
and hull kept fixed, the SWENSE method can be
extended to simulations in other wave conditions,
including oblique waves and irregular waves with all
degrees of freedom.
Additional applications will include the
simulation of a self propelled ship in waves
incorporating fins and rudder action. Seakeeping and
manoeuvring problems will thus be addressed within a
unified approach.

7 ACKNOWLEDGEMENTS
The authors express their thanks to the
Dlgation Gnrale pour l'Armement (DGA) which
is supporting this work. The IIHR Institute, Iowa, is
gratefully acknowledged for making the experimental
results available online.

8 REFERENCES
Alessandrini, B., Delhommeau, G., A multigrid
velocity-pressure-free surface elevation fully coupled
solver for calculation of turbulent incompressible flow
around a hull, Proceedings of the 9th International
Conference on Numerical Methods in Laminar and
Turbulent Flows, Atlanta, 1995, pp. 1173-1184.
Di Mascio, A., Landrini, M., Lalli, F., Bulgarelli, U.,
Three Dimensional non linear diffraction around
fixed structures, Proceedings of the 20th Symposium
on Naval Hydrodynamics, 1994.
Ferrant, P., Simulation of strongly non-linear wave
generation and wave-body interaction using a 3D
MEL model, Proceedings of 21st Symposium on
Naval Hydrodynamics, Trondheim, 1996.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Ferrant, P., Le Touz, D., Fully-non-linear


spectral/BEM solution for irregular wave interactions
with a 3D body, Proceedings of the OMAE
Conference, Oslo, 2002.
Ferrant, P., Gentaz L., Alessandrini, B., Le Touz, D.,
A potential / RANSE approach for regular water
wave diffraction about 2D structures, Ship
Technology Research, Vol. 50, N 4, 2003.
Gentaz, L., Luquet, R., Alessandrini, B., Ferrant, P,
Numerical Simulation of a 3D viscous flow around a
vertical cylinder in nonlinear waves using an explicit
wave model, Proceedings of the OMAE Symposium,
Vancouver, 2004.
Gui, L., Longo, J., Metcalf, B., Shao, J., Stern, F.,
Forces, Moment and Wave Pattern for Naval
Combatant in Regular Head Waves, Proceedings of
the 23rd Symposium on Naval Hydrodynamics, 2001.
Longo, J., Stern, F., Resistance, Sinkage and Trim,
Wave Profile, and Nominal Wake and Uncertainty
Assessment for DTMB Model 5512, Proceedings of
the 25th ATTC, Iowa City, 1998.
Luquet, R., Gentaz, L., Ferrant, P., Alessandrini, B.,
RANSE analysis of 2D flow about a submerged
body using explicit incident wave models,
Proceedings of the 6th Numerical Towing Tank
Symposium, Rome, 2003.
Rhie, C. M., Chow, W. L., A numerical study of the
turbulent flow past an isolated airfoil with trailing
edge separation, 1983, AIAA Journal, vol. 21, pp.
179-195.
Rienecker, M.M.; Fenton, J.D., A Fourier
approximation method for steady water waves,
Journal of Fluid Mechanics, Vol. 104, 1981, pp. 119137.
Wilcox, D., C., Reassessment of the scaledetermining equation for advanced turbulence
models, 1988, AIAA journal, Vol. 26, pp. 12991310.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Hoyte C. Raven
MARIN, Netherlands
In the approach proposed by the authors, the
velocity field is decomposed in a known field
connected with the incoming wave system, and an
unknown field containing the diffracted waves, the
steady wave pattern and the boundary layer and
wake. This seems a quite sensible step and avoids the
need to have a fine discretisation just to represent the
simple incoming wave accurately.
However, the diffracted field and steady
wave pattern also ask for a fine discretisation, and the
results shown indicate that this is still hard to
achieve. The question that comes to mind then is,
whether alternative decompositions could alleviate
those requirements as well. E.g. if the known field UI
is defined as containing also the diffracted field
according to potential theory, or even the steady
wave pattern, could the accuracy be improved
further? Could you please comment on this option?
AUTHORS REPLY
We have already think of considering the
inviscid solution as the incident field because as you
said we are sure that this would permit us to improve
our results and especially in the far field even by
using coarser grids. But the implementation of this
new incident field would lead to many problems. If
we add the steady wave pattern and the diffracted
field in the incident field we would not be able to
conserve a spectral method to generate it. As a
consequence the method that we would use would
create a field which would not be continuously
defined through and above its free surface. Or this is
a crucial point to apply the SWENSE method.
Nevertheless, I think that this idea should be
performed in the next future because it should give
accurate results of the viscous wave pattern of a ship
in waves in the both near field and far field by using
coarse grids so low CPU time requirements.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

St Johns, Newfoundland and Labrador, CANADA, 8-13 August 2004

Computation of Three-Dimensional Free-Surface Flows with an


Automatic Adaptive Mesh Refinement and Coarsening Strategy
A. Hay, P. Queutey, M. Visonneau
Laboratoire de Mcanique des Fluides, CNRS UMR 6598
Ecole Centrale de Nantes, B.P. 92101
44321 Nantes Cedex 3, France
ABSTRACT
This paper describes a modern free-surface capturing
methodology implemented in an unstructured finitevolume flow solver. A systematic grid-refinement study
and a comparison of several discretisation schemes for the
transport equation of concentration is performed to evaluate the influence of the discretization error on the interface
smearing. Detailed comparisons with experiments carried
out at model scale on the Series 60 bare hull illustrate
the high degree of reliability of this approach, provided
that adequate compressive discretisation schemes are used
with a reasonable number of grid points. A promising
adaptive grid refinement and coarsening strategy suited to
free-surface flows is described in the second part and applied to the computation of the free-surface flow around
the Wigley hull.

INTRODUCTION
The use of Computational Fluid Dynamics (CFD) tools
for predicting the powering performance of ships remain
challenging because of the numerous physical difficulties
which characterize the flow around a real ship. Among the
hard points which have to be solved to get a reliable simulation of ship flow, one can list, without being exhaustive, the accurate simulation of ship stern flows which is
highly dependent on turbulence modelling, the modelisation of the hull/propeller coupling and the simulation of
the free-surface deformation. On one hand, one must notice that the accuracy of CFD tools for predicting crucial
global quantities like resistance and self-propulsion factor
is still limited. On the other hand, CFD is complementary
to towing tank tests because it provides a large amount of
detailed information on the flow which helps the designer
to improve the performance of new ship. Therefore, as
long as simplified configurations are considered, the large
increase of computational power and memory makes possible to have recourse to CFD codes which may be inte-

grated in fully automated optimization procedures during


the design phase. When the challenges related with turbulence modelisation and ship hull optimisation have been
treated in (1) and (2), this paper is devoted to the simulation of free-surface flows and its requirements in terms of
simulation strategy, discretisation schemes accuracy and
computational power.
It has been observed during the last Gothenburg 2000
workshop (3) that the free-surface capturing methodology was more and more popular among the CFD developers dealing with viscous naval hydrodynamics. Recently, our team has successfully included this approach
in our code ISIS to compute off-shore flows or waveresistance problems (4). This increasing interest is due
to the fact that this approach is more robust than those
based on a free-surface fitting methodology since no regridding is necessary and the numerical wave-breaking
which may occur during the initialization period, is perfectly tolerated. When specific compressive discretisation
schemes are used to discretize the concentration transport
equation, one can ensure that the density discontinuity between air and water is captured on three to five control
volumes. However, if the discontinuity occurs in a region
where there are not enough grid points, the free-surface elevation is dramatically smeared, making the free-surface
capturing strategy far less accurate (and far more expensive) than the classical algorithms based on free-surface
fitting. Therefore, the almost perfect numerical strategy
should integrate a coupling between a free-surface capturing approach and an automatic adaptive mesh refinement
and coarsening methodology in order to maintain dynamically a prescribed density of grid points around the steady
or unsteady interface between air and water.
During the last two years, such a capability was first implemented in our CFD code for error-estimation purposes
during the PhD thesis of A. Hay (5). For instance, the
study described in (6) presented a new a-posteriori error
estimate for a finite volume method based on a transport equation for the error. Recently, our group has

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

also extended this method to the computation of threedimensional free-surface flows by using an explicit error
indicator.
This paper is devoted to a presentation of the specific discretisation schemes which are required to get an accurate
description of a density discontinuity. Moreover, the influence of the number of grid points and discretization
schemes is analysed by comparing the respective numerical solutions to an extensive experimental database (7)
related to the Series 60. In the second part of the paper,
an automatic grid refinement and coarsening is applied to
the simulation of the free-surface flow around a Wigley
hull. The advantages of this approach are demonstrated in
terms of accuracy, optimal use of discretisation points and
user-friendliness.

Ui dV +
Ui ( U U d )
n dS
V
S
Z
Z

= (ij Ij pIi )
n dS +
gi dV

Z
ci dV +

ci ( U U d )
n dS = 0

Flow solver
The ISIS flow solver, developed by DMN (Division Modlisation Numrique i.e. CFD Department of the Fluid
Mechanics Laboratory), uses the incompressible unsteady
Reynolds-averaged Navier Stokes equations (RANSE).
The solver is based on the finite volume method to build
the spatial discretization of the transport equations. The
face-based method is generalized to two-dimensional,
rotationally-symmetric, or three-dimensional unstructured meshes for which non-overlapping control volumes
are bounded by an arbitrary number of constitutive faces.
The velocity field is obtained from the momentum conservation equations and the pressure field is extracted from
the mass conservation constraint, or continuity equation,
transformed into a pressure-equation. In the case of turbulent flows, additional transport equations for modeled
variables are solved in a form similar to the momentum
equations and they can be discretized and solved using
the same principles. Incompressible and non-miscible
flow phases are modelized through the use of conservation equations for each volume fraction of phase.

(1c)

where V is the domain of interest, or control volume,


bounded by the closed surface S with a unit normal vector

n directed outward.
The effective flow physical properties (viscosity and
density ) are obtained from each phase physical properties (i and i ) with the following constitutive relations:
X
X
X
=
ci i ; =
ci i ; 1 =
ci (2)
i

COMPUTATIONAL APPROACH

(1b)

When the grid is moving, the so-called space conservation law must also be satisfied:
Z
Z

dV
Ud
n dS = 0
(3)
t V
S
Numerical framework

Spatial discretization All the flow variables are stored


at geometric centers of the arbitrary shaped cells. Surface
and volume integrals are evaluated according to secondorder accurate approximations by using the values of integrand that prevail at the center of the face f , or cell C,
and neighbor cells. The various fluxes appearing in the
discretized equations are built using centered and/or upwind schemes. For example, the convective fluxes are obtained by two kinds of upwind schemes. A first scheme
available in the flow solver, (HD) for Hybrid differencing, is a combination of upwind (UD) and centered (CD)
schemes. Contrary to a practical approach (8; 9) where
CD/UD blending is fixed with a global blending factor for
all faces of the mesh, the HD scheme results from a local blending factor based on the signed Peclet number at
the face. An other upwind scheme which is implemented
in ISIS, is the Gamma Differencing Scheme (GDS) (10).
Through a normalized variable diagram (NVD) analysis (11), this scheme enforces local monotonicity and
convection boundedness criterium (CBC) (12). A presConservation equations
sure equation is obtained in the spirit of the Rhie and
The flow solver can deal with multi-phase flows and mov- Chow (13) procedure and momentum and pressure equaing grids. Considering incompressible flow of viscous tions are solved in an segregated way like in the wellfluid under isothermal conditions, mass, momentum and known SIMPLE coupling procedure.
volume fraction conservation equations can be written as
Discretization schemes for the transport equation of
(using the generalized form of Gauss theorem):
concentration Except the convection terms and voluZ
Z
metric mass fluxes, interfacial quantities qf are rebuilt lin

dV +
( U U d ) n dS = 0
(1a) early from the quantities themselves and their available
t V
S

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

cell-centered gradients. In order to assume face bounded


reconstructions and to avoid unrealistic oscillations, the
search for an acceptable compromise between accuracy
and boundedness of 0 6 ci 6 1 is a key point (14; 15).
Moreover, the method must also preserve the sharpness of
the interface through the transport equation (1c).
Considering normalized variables, Eq. (4), the CBC criterium (11) implies that, in the NVD diagram Fig. 1(a),
the only scheme which unconditionally satisfies the
boundedness criterium is the first-order upwind differencing scheme (UDS).
q =

q qU
qD qU

0.7 Co
0.7 0.3
Co > 0.7

qf = qC

if










































































CDS










































1/2 








































UDS



















































































1

(4)

The second-order centered differencing scheme (CDS)


is only useful in the range 0 6 qC 6 1. In order to provide a continuous switch from UDS to CDS and CDS to
UDS, the GDS scheme illustrated into the NVD diagram
by Fig. 1(b) yields a first answer to that requirement.
However, the GDS scheme is based on upwind or centered differencing and could not be appropriate for maintaining the sharpness of an interface since numerical diffusion is introduced. On the other hand, this scheme does
not suffer from any Courant number limitation. An alternate treatment of numerical diffusion problem for interface capturing method is to use the donor-acceptor cell
concept (16). This can be realized with the GDS scheme
if one introduces downwind differencing (DDS) to build
the Inter-Gamma scheme (17), Fig. 1(c), since compressive characteristics are required for an accurate interface
capturing. The main disadvantage of the IGDS scheme is
a Courant number limitation : Co < 0.3 in multidimensional cases. To overcome that difficulty, the following
correction (5) is used which proved satisfactory in most
applications (18). Then, the adopted scheme for interface
capturing MGDS, results from the Courant number limited scheme IGDS with corrections from (5).
qf = qC + (
qf qC )

~
qf

if

0.3 6 Co 6 0.7
(5)

Temporal discretization The temporal discretization is


based on a three-step upwind discretization ensuring a
second-order accuracy. The convection and diffusion are
treated implicitly.

~
qC

(a) UDS and CDS Schemes


~
qf
UDS





CDS
















1/2 









1

UDS

~
qC

(b) Gamma Scheme (GDS)


~
qf
UDS



DDS







CDS



1/2 










1

UDS

m =1/2

~
qC

(c) Inter-Gamma Scheme (IGDS)

Figure 1: Numerical Schemes in NVD diagram

are carried out in a full three-dimensional domain for the


validations with a steady drift. To speed-up the converWall function conditions are imposed on the walls. Sym- gence towards a steady state, the computations are permetry conditions are written in the vertical plane of sym- formed around a moving ship with a constant acceleration
metry for the computations without drift. Computations rate from T=0 to T=1 followed by a constant speed.
Boundary conditions

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Validation on Series 60 bare hull in drift


Description of the test case
In order to assess the reliability of this free-surface capturing strategy, the Series 60 database (7) has been chosen because of the availability of several well documented
drift configurations. Two drift angles (5 and 10 degrees)
have been retained. The computations are performed on
a parallel computer using MPI and the k SST model
is selected to model the turbulent correlations. An original and robust wall-function (19) boundary condition is
employed on the wall instead of a near-wall low Reynolds
number formulation in order to avoid, until now, any difficulty related with the behavior of the interface in the
vicinity of the wall when very fine grids suited to a low
Reynolds number formulation are employed.
In order to get a well converged result, especially to satisfy the wave propagation in the far field, it is mandatory
to compute up to an adimensional time t=40 for the most
severe test case (drift angle of ten degrees).
This is why it has been decided to start the computation with the GDS scheme which has no constraints on
the Courant number, and then switch to the compressive
MGDS scheme with a time step ten times smaller to restore the sharpness of the interface at the end of the computations. This procedure which is only valid for steady
flows allows free-surface computations for reasonable cpu
cost. All the presented numerical simulations have been
performed on a SGI ORIGIN 3800 R14000/500Mhz machine.
Grid independence study
To evaluate the influence of the number of discretization
points on the solution, a grid independence study has been
carried out on three unstructured grids whose topology is
described in Fig. 2 with only some significant domains.
To avoid too many lengthy computations, this study is
limited to the drift angle of 10 degrees. In order to control
the smearing of the interface, hexahedric finite volumes
are used near the interface and tetrahedric elements are
preferred in the far field to reduce the number of points.
Hexahedric elements are also retained near the wall to
guarantee an accurate discretisation of the viscous diffusion effects.
The characteristics of the three grids employed are described in the table 1 with cpu time spent. Figs. 3 show the
free-surface elevation for a steady drift of 10 degrees computed on the afore-mentioned three grids with the MGDS
discretisation scheme. From these figures, one can notice that the coarse grid is definitely too coarse since it is
not able to capture the correct Kelvin angle although the
main divergent waves are already captured. The medium

Figure 2: Typical unstructured grid topology for a freesurface flow


Grid
Coarse
Medium
Fine

Number of
grid points
6.52 105
2.82 106
3.89 106

Number of
processors
16
32
32

Total
CPU time
7 hours
2.5 days
5 days

Table 1: Characteristics of the grids used in the grid independence study


and fine grids reveal far more accurate representations of
the free-surface with a correct Kelvin angle, steeper divergent and transversal waves and, for the fine grid, many
secondary divergent waves in good agreement with the experiments as it will be illustrated in the next sections.
Influence of the discretisation scheme
The challenge posed by the discretisation of a transport
equation for the concentration is the accurate modeling of
a contact discontinuity, i.e. the free-surface. From the
previous section, it appears that the compressivity is theoretically crucial to keep the interface as sharp as possible,
even if this property is satisfied through a severe Courant
number limitation. It is therefore interesting to evaluate
on fine grids and real-life problems, the influence of the
discretisation error to check if it is realistic to use noncompressive discretisation schemes which result in faster
computations since no criteria has to be satisfied on the
Courant number. Figs. 4 show comparisons of the wave
profiles on the wall of a Serie 60 hull obtained for a drift
angle of 10 degrees on the fine grid.
From these figures, one may conclude that the discretisation schemes play a minor role since the results are quite
similar. However, if one considers Figs.5 which compare

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

0.5

1.5

0.035
0.5

0.030

0.4

0.4

0.025

0.3

0.3

0.020

0.2

0.2

0.015

0.1

0.1

0.0

0.0

-0.1

-0.1

0.000

-0.2

-0.2

-0.005

-0.3

-0.010

-0.3
0
=10: Coarse (MGDS:KWSST)

0.5

x/L

1.5

=z/L

y/L

0
0.5

GDS
MGDS

0.010
0.005

= 0.001

-0.015

0.00

0.10

0.20

0.30

0.40

= 10 : Port : Fine(KWSST)

0.50

x/L

0.60

0.70

0.80

0.90

1.00

0.80

0.90

1.00

(a) Coarse grid


(a) 10 degrees - Port side
0.5

1.5
0.5

0.035

0.4

0.4

0.030

0.3

0.3

0.025

0.2

0.2

0.020

0.1

0.1

0.015

0.0

0.0

-0.1

-0.1

-0.2

-0.2

0.000

-0.3

-0.3

-0.005

0
=10: Medium (MGDS:KWSST)

0.5

x/L

1.5

= 0.001

=z/L

y/L

0
0.5

GDS
MGDS

0.010
0.005

-0.010
-0.015

0.00

0.10

0.20

0.30

0.40

= 10 : Starboard : Fine(KWSST)

(b) Medium grid

0.50

x/L

0.60

0.70

(b) 10 degrees - Starboard side

y/L

0.5

1.5

0.5

0.5

0.4

0.4

0.3

0.3

0.2

0.2

0.1

0.1

0.0

0.0

-0.1

-0.1

-0.2

-0.2

-0.3
0
=10: Fine (MGDS:KWSST)

-0.3
0.5

x/L

(c) Fine grid

1.5

= 0.001

Figure 4: Free-surface elevation on the waterline. Comparison of GDS and MGDS discretisation schemes for a
drift angle of 10 degrees
tion.
Additional illustrations of the influence of the compressive property will be given in the section relative to wave
cuts which will confirm this analysis.

Figure 3: Series 60 with a drift angle of 10 degrees. Influ- Free-surface elevation


ence of the number of grid points
Figure 6 shows the computed free-surface elevations compared to the experiments for a drift angle of 5 degrees.
the free-surface elevation field all around the ship hull, it Lastly, Fig. 7 shows the same comparison for a drift angle
is clear that the discretisation schemes still play a central of 10 degrees. These results are obtained on the fine grid.
role even if these results are obtained on the finest grid. One can notice the very good agreement between comNot only the amplitudes of the main divergent waves are putations and experiments since all the divergent waves
modified but one can observe that the free-surface field are accurately captured by the simulation. The transverse
computed with MGDS contain many secondary waves in- waves at the stern of the ship are also in very good agreeside the Kelvin angle in good agreement with the experi- ment with the experiments.
ments, a behavior which is not simulated by the cheaper
The wave contours are significantly influenced by the
GDS scheme.
drift angle since the wave amplitudes are drastically inParticularly, it is noteworthy to mention that, at 5 and 10 creased/decreased on the port/starboard sides. It appears
degrees drift angles, the bow wave spray sheet detaching that the influence of the compressive MGDS scheme is
from the hull and merging with the downstream breaking far more important on the port side because the water is
bow wave, is remarkably captured by the MGDS simula- pushed away from the hull resulting in a downwind de-

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

0.5

1.5

0.5

1.5

0.5

0.5

0.4

0.4

0.4

0.4

0.3

0.3

0.3

0.3

0.2

0.2

0.2

0.2

0.1

0.1

0.1

0.1

0.0

0.0

0.0

0.0

-0.1

-0.1

-0.2

-0.2

-0.1

-0.1

-0.2

-0.2

-0.3

-0.3

0
=10: Fine (GDS:KWSST)

0.5

x/L

1.5

-0.3

= 0.001

-0.3

0
= 5: Fine (MGDS:KWSST)

0.5

(a) GDS

0.5

x/L

1.5

= 0.001

(a) Computations
1.5

0.5

0.5

0.4

0.4

0.3

0.5

1.5

0.5

0.5

0.3

0.4

0.4

0.2

0.2

0.3

0.3

0.1

0.1

0.2

0.2

0.0

0.0

0.1

0.1

-0.1

-0.1

0.0

0.0

-0.2

-0.2

-0.1

-0.1

-0.2

-0.2

-0.3
0
=10: Fine (MGDS:KWSST)

-0.3
0.5

x/L

1.5

= 0.001

y/L

y/L

0.5

y/L

y/L

0
0.5

-0.3
0
= 5: Experiment

-0.3
0.5

x/L

1.5

= 0.001

(b) MGDS
(b) Experiments

Figure 5: Free-surface elevations for a drift angle of 10


degrees. Comparison of GDS and MGDS discretisation Figure 6: Free-surface elevations for a drift angle of 5
schemes
degrees
pendency, only accounted for by the compressive scheme.

Wave profiles cuts


The transversal wave cuts shown in Figure 8 illustrate the
influence of the discretization error on the free-surface elevations. One can notice that the compressive property
included in MGDS dramatically improves the solution on
the port side when the solution on the starboard side does
not seem strongly dependent on the discretisation scheme.
In particular, one may notice the occurence of breaking
waves in front of the hull which are not captured by a more
diffusive scheme (GDS). This phenomena is observed by
the experimental study (see 9 from (7). Figure 10 shows
the same transversal wave cuts for a drift angle of 5 degrees.
Figs. 10 shows comparisons between results obtained
with MGDS schemes to the experiments performed by
IIHR (7) for a drift angles of 5 degrees. The computational results obtained on the fine grid with MGDS agree
almost perfectly with the measurements. The same conclusions remain valid when the computations are examined through the longitudinal wave cuts shown in figs 11
and 12.

Evolution of the vorticity


Although the previous sections were focussed on the reliability of the free-surface simulation, it is very important to
examine the behaviour of the longitudinal vortices which
are created along the hull. A global view of the distribution of longitudinal vorticity around the Series 60 for a
drift angle of 10 degrees is shown in Figure 13. The computed results are obtained on the fine grid with the MGDS
scheme and the experiments are taken from (7). Effect
of drift angle on vortices generation is nicely reproduced
by the simulations : the forebody bilge vortex (FBV), the
forebody keel vortex (FKV) and the wave bilge vortex
(WBV) initiated on the forebody with inboard, inboard,
and outboard rotation, respectively. The FBV separates
from the hull and follows a trajectory outboard and toward the free surface. The ABV and ABCV structures are
generated on the afterbody and rotate approximately 90
in the wake from x/L = 1.0 to 1.2 in the direction of
the ABV rotation. This behavior is also captured with the
numerical simulation. Figure 14 illustrates the location
of the ABV, ABCV and FBV vortices near the propeller
plane x/L = 1.0 through the longitudinal vorticity distribution. The reduced scale is used to render the FBV

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

(a) Computations with the MGDS scheme

(b) Experiments

Figure 13: Evolution of the longitudinal vorticity around the Series 60 for a drift angle of 10 degrees - Comparison
between computations and experiments

structure. Apart from a weaker predicted level in the upper part of the ABV vortex, the fine mesh density is able
to capture the location of the near-wake structure. Endly,
it should be pointed out that, due to the free surface capturing method, the wave-breaking vortex (WBV) is even
observed numerically although much more diffused (in
the transversal direction) when compared to the experiments. The SST-K turbulence model is used without
any modifications to account for free surface effects and,
since this vortex results from a strong interaction of the
free surface with the flow, a high (abnormal ?) level of
turbulent viscosity is produced at the vicinity of the free

surface that rapidly dissipates the strength of this vortex.


ADAPTIVE MESH REFINEMENT AND COARSENING STRATEGY

Adaptive techniques
The ultimate goal of an adaptive procedure is clearly to
reduce the discretization error for reaching a solution of
prescribed accuracy for a low computational cost. But,
one also wishes to equidistribute the error over the whole
computational domain for the solution to be of similar ac-

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

0.5

1.5

1.05
0.5

0.4

0.4

0.3

0.3

0.2

0.2

0.1

0.1

0.0

0.0

-0.1

-0.1

-0.2

-0.2

-0.3

1.00
0.95
0.90
0.85

x/L

y/L

0
0.5

0.75

-0.3

0
=10: Fine (MGDS:KWSST)

0.5

x/L

1.5

0.80

= 0.001

0.70
0.65

(a) Computations

GDS
MGDS

=0.02L

0.60

y/L

0.5

0.55

1.5

0.5

0.5

0.4

0.4

0.3

0.3

0.2

0.2

0.1

0.1

0.0

0.0

-0.1

-0.1

-0.2

-0.2

-0.3
0
=10: Experiment

-0.4

x/L

1.5

0.0

y/L

0.2

0.4

0.6

(a) +0.60 < X/L < +1.00

0.50
0.45

-0.3
0.5

-0.2

=10: Fine (KWSST)

= 0.001

=0.02L

0.40
0.35

(b) Experiments

GDS
MGDS

Figure 7: Free-surface elevations for a drift angle of 10


degrees

x/L

0.30
0.25
0.20
0.15
0.10

curacy everywhere. In order to do so, each control volume marked for grid refinement is subdivided into several
smaller ones of the same topology. As the initial mesh can
possibly be too fine in some region for the desired accuracy, it can be coarsened by an agglomeration algorithm.
The steady adaptive procedure, described in Figure 15, is
included in the ISIS code making it an automatic single
numerical tool. It starts with the computation of a first
numerical solution on an initial uniform grid with an arbitrary number of cells. The numerical error is then estimated by various methods described in (5) or (6) and this
information is used to decide which changes should be
made on the current mesh. The criteria of selection of a
cell i on the current Gridj are the following :
Refinement : ErrGridj (i) > TolR ||ErrGrid1 ||L1
Unrefinement : ErrGridj (i) < TolD ||ErrGrid1 ||L1
The parameters TolR and TolD control the intensity of the
procedure during one adaptive step for the refinement and
the unrefinement process respectively. The numerical solution is then mapped on the new adapted mesh and the
computation is resumed on this grid and this procedure is
repeated until the error estimate is below a desired value.

0.05
0.00
-0.4

-0.2

=10: Fine (KWSST)

0.0

y/L

0.2

0.4

0.6

(b) 0.00 < X/L < +0.50

Figure 8: Wave profiles at X=cst. for a drift angle of 10


degrees - Comparison between GDS, MGDS discretisation schemes and experiments
Unsteady adaptive procedure
The above-mentioned steady adaptive procedure based
on the error estimation has already been successfully applied by the authors to the computation of steady turbulent
flows (6). Although the same approach can be retained for
multi-fluid flows, once generalised to unsteady problems,
it is wiser, for a first step, to use a strategy based on an
explicit physical indicator.
As indicated before, a free-surface capturing methodology is based on the solution of a transport equation
of a contact discontinuity which indicates the location
of the interface. Although the compressive discretisation

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

1.05
1.00
0.95
0.90

x/L

0.85
0.80
0.75
0.70
0.65

=0.02L

0.60

Figure 9: Experimental view of the bow wave breaking


for a drift angle of 10 degrees (from (7))

0.55
-0.4

-0.2

= 5: Fine (MGDS:KWSST)

0.0

y/L

0.2

0.4

0.6

(a) +0.60 < X/L < +1.00


schemes described above guarantee an interface captured
over three to five cells, it is crucial to keep as small as
possible the characteristic length of the cells on either
0.50
side of the interface in order to avoid too much a numer0.45
ical smearing of the discontinuity. Ideally, such an ap0.40
proach should make possible the use of a grid schematically described in Figure 17(a) instead of that shown in
0.35
Figure 17(b). Moreover, an automatic grid adaption frees
0.30
the user from a tedious task : the generation of a mesh
0.25
suited everywhere to the unknown interface since one can
0.20
start from a uniform grid which will be automatically re0.15
fined near the interface, thanks to the explicit interface
indicator. Lastly, in case of unsteady waves, an adaptive
0.10
grid strategy based on successive refinement and unrefine0.05
ment steps is optimally suited to follow the temporal evo0.00
-0.4
-0.2
0.0
0.2
0.4
0.6
lution of the interface. Therefore, the adaptive unsteady
y/L
= 5: Fine (MGDS:KWSST)
procedure is summarized on figure 18.
It is based on an explicit error indicator which is pro(b) 0.00 < X/L < +0.50
portional to the gradient of concentration in order to concentrate the points in the vicinity of the interface. A safety
margin has to be added around the detected zones as indi- Figure 10: Wave profiles at X=cst. for a drift angle of 5
degrees - Comparison between computations and expericated in figure 19.
ments
x/L

=0.02L

Grid alteration
During the refinement process, each control volume to
be refined is split into several smaller ones of the same
topology. Thus, the topology of the element on any part
of the mesh will always be the same and only their size
will be adapted. The refinement process can occur with a
possible directional sensitivity for flows with simple features. A non-refined neighbor of a refined cell presents a
so called hanging node which is accounted for naturally
by our face-based finite-volume method : a face with a
hanging node is simply seen as several smaller faces.
The agglomeration algorithm permits to unrefine the
selected parts of the mesh by grouping neighboring cells

marked for coarsening. The different groups are then


merged into a bigger cell. The agglomeration can result
from a fusion of a face or of a node as illustrated on figure 16. As shown, the resulting agglomerated cell is of no
usual shape but once again the face-based methodology is
designed to deal with such peculiarities.
All these possible grid alterations can be canceled as
the initial mesh can be recovered by the use of connectivities between the different generations of meshes. Thus,
the regions of the mesh that must be adapted are allowed
to change during the computation as it is the case for unsteady computation.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

0.35
min=-15, max=65, =5

0.30
0.25

0.00

0.20

z/L

0.15
0.10

y/L

0.05

=0.02L

-0.05

0.00
-0.05
-0.05

-0.10

0.00

0.05
y/L

x/L=1.0 : =10: Fine (MGDS:KWSST)

0.10

0.15

-0.15
-0.20

(a) Computations with the MGDS scheme

-0.25
-0.30
-0.35

-0.4

-0.2

0.0

0.2

0.4

0.6

= 5: Fine (MGDS:KWSST)

0.8

x/L

1.0

1.2

1.4

1.6

1.8

min=-15, max=65, =5

2.0

z/L

0.00

Figure 11: Wave profiles at Y=cst. for a drift angle of 5


degrees - Comparison between computations and experiments

-0.05

0.35
0.30

-0.05

0.00

0.05
y/L

Exp, =10, x=1.0, Fr=0.316

0.25

0.10

0.15

0.20
0.15

(b) Experiments

0.10

y/L

0.05

=0.02L

Figure 14: Longitudinal vorticity distribution at X/L =


1.0 for a drift angle of 10 degrees - Comparison between
computations and experiments

0.00
-0.05
-0.10
-0.15
-0.20
-0.25
-0.30
-0.35

-0.4

-0.2

0.0

=10: Fine (MGDS:KWSST)

0.2

0.4

0.6

0.8

x/L

1.0

1.2

1.4

1.6

1.8

CFD SIMULATION

ERROR ESTIMATION

Reduction of residuals
of several orders

Computation of the
error a posteriori

Initial Mesh

ADAPTATION

INTERPOLATION

Refinement
Unrefinement

Criteria
reached ?
No

Yes

2.0

Figure 12: Wave profiles at Y=cst. for a drift angle of 10


degrees - Comparison between computations and experiments

End

Figure 15: Adaptive steady procedure

Mesh update
When a new node is added on the surface of the body, two
requirements have to be fulfilled:
1. the node must be placed on the exact geometry which
should be described either analytically or by a CAD
description readable by the flow solver,
Figure 16: Agglomerated cells
2. for convex boundaries and high aspect ratio cells, the
inclusion of a new node on the surface of the body
should not generate neighboring cells with negative
generalizing the mesh deformation tools already included
volumes as illustrated by Figure 20,
in the flow solver and previously used for the shape optiTo avoid the apparition of ill-conditionned near-wall cells, mization studies.
it has been decided to move the three-dimensional grid in
For two and three-dimensional cases, a lineal and toraccordance with the discrete boundaries deformation by sional spring analogy is employed to control the defor-

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Free surface detected

Margin added

(a) Adapted grid

Initial Mesh

Adapted Mesh

Figure 19: Safety margin around the interface


(b) Unadapted grid

Figure 17: Adaption of grids around an interface

Exact Boundary
(2)

(1)

(3)

Initialization t c = t o
Initialization of the
mesh

Figure 20: Illustration of situations for which a grid deformation tool is required

Residual reduction
reached ?
Yes
No
Time of adaptation ?
tc = Mod (to ,Nt)
Yes
No

square of the length lij of the edge.


temporal loop

new convergence of the current time step

New non linear iteration

non linear loop

New Time Step


tc + t
tc

Need of adaptation ?
Yes

No

Adaptation of the
current mesh
Interpolation

Figure 18: Adaptive unsteady procedure

mation of the grid (20). The lineal spring analogy is commonly used to deform the mesh in shape design strategies.
A lineal spring is attached along each edge connecting two
nodes i and j, with a stiffness inversely proportional to the

kij

1
2
lij

(6)

The displacement of the nodes q is the solution of a linear


system, which represents the quasi-static equilibrium of
the discrete mechanical problem :
Klin q = 0 q = q

on boundaries,

(7)

where Klin is the stiffness matrix and q the displacement


of the nodes imposed on the boundaries. This method prevents two nodes from colliding, since in that case the stiffness tends towards an infinite value according to (6), but it
does not prevent a node from colliding the edge that faces
it. This behaviour is particularly damageable for viscous
calculations, since the high stretching of the control volumes is unavoidable in the boundary layer and may give
rise the failure of the deformation process by generating
negative volumes. Therefore, special treatments are usually applied near the wall, involving geometric considerations (21). In this paper, an other approach which associates a torsional spring analogy with the lineal one (20) is
employed. The method consists in attaching to each node

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

i of each volume a torsional spring with a stiffness defined


by :
1
,
(8)
Ciijk
2 ijk
sin (i )
at the node i of the volume.
where iijk is the angle jik
The system (7) is then updated in :
on boundaries, (9) Figure 21: Global view of the region where the adaption
criteria is activated
using an energy equivalence to express the effects of the
Number of
Number of cells
Total monoproc.
torsional springs in terms of displacements of the nodes.
generations
(volume) CPU time (hours)
The torsional springs prevent the volumes from interpen2
432000
10
etrating each other. The association of lineal and torsional
3
142400
48
springs provides a powerful mesh deformation tool, which
maintains the grid quality near the wall even for high dereference grid
1718100
960
formations.
(Klin + Ktors )q = 0

q=q

FREE-SURFACE AROUND THE WIGLEY HULL

Table 2: Characteristics of the grids used for the computations of the free-surface around the Wigley hull

Characteristics of the computations


The Wigley hull (22) has a shape which is defined by a million points is used to validate the convergence of the
adapted grid solutions towards a grid-independent solumathematical formula given by:
tion. Table 2 provides also the duration of the respective
y2
B
2
)
(10) computations expressed in mono-processor hours (IBM
z(x, y) = (1 4x )(1 (
2
D
Power4 P690) for 4 seconds of computations in physical
where B is the maximum thickness of the hull and D is the adimensional time. The CPU time needed to get a condraft of the hull. These constants are set here to 0.1 and verged steady solution on the best locally adapted grid (3
0.0625 respectively. This hull has been preferred to the generations) is reduced by a factor 20 when compared to
more realistic Series 60 hull because its analytical defini- that needed on the fine unadapted grid, which illustrates
tion facilitates the inclusion of the new grid points on the the potentialities of such a strategy.
surface of the body. Computations using slip wall conditions instead of wall-function boundary condition, are
performed for a Froude number of 0.316 on grids systematically refined according to an explicit indicator based
on the vertical component of the gradient of the concentration Ci . However, to avoid the inclusion of points in
regions where the free-surface is not deformed, this criteria is activated only inside a triangular region around the
hull. Figure 21 shows the typical grid concentration near
the interface after three levels of refinement. One can notice here that the criteria used in this preliminary study is
based on the density discontinuity without considering the
shape of the waves. This is the reason why no wave pattern is visible on this figure. It is clear that such a criteria
will have to be improved in the future to distinguish between moderate and steeper waves. Such an improvment
will require the use of a directional refinement approach
instead of the actual isotropic strategy.
Three refinement steps have been performed and the
characteristics of the grids are described in Table 2.
An additional computation performed without local
grid refinement on a very fine grid containing about 2

Free-surface
The elevation on the waterline for the grids with different generations of refinement, the fine reference grid and
the experiments are given in Figure 22. The convergence
towards the results obtained on the fine reference grid is
well established, even if some local differences indicate
that the grid built with 3 generations is likely too coarse
in some regions of the flow. Figures 23 illustrate the structure of the mesh near the waterline for all the generations
of refinement. On one hand, one can observe that the interface is very well captured by the last locally adapted
grid. On the other hand, one may fear a certain lack of
grid points away from the interface which may deteriorate the global quality of the simulation since the position
of the interface depends on the flow field in its proximity.
For the purpose of comparison, Figure 24 shows the same
information with the fine reference grid.
A gobal view of the free-surface elevation is shown by
Figures 25. Computations relative to the fine grid of reference are compared to that obtained on grids containing

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

0.02

Experiments
ISIS : NGen=0
ISIS : NGen=1
ISIS : NGen=2
ISIS : NGen=3
ISIS : Fine Mesh

0.015

Z
X
Y

Z/L

0.01
0.005
0
-0.005
-0.01

(a) Initial mesh


-0.5

X/L

0.5
Z
X
Y

Figure 22: Free-surface elevation along the waterline for


the different grids

two and three levels of local refinement in a prescribed region indicated above. These preliminary results are very
encouraging : the free-surface behaviour is very well captured in the immediate vicinity of the hull on the adapted
grids. The deterioration of the accuracy of the simulation observed in the far-field is likely due to a lack of grid
points in the vertical direction on both sides of the density
discontinuity.
From all these first results on the use of automatic local
grid adaption for the modelization of free-surface flows,
one can draw the following conclusions :
a complex automatic unstructured grid refinement/coarsening strategy has been successfully implemented in a complex CFD code,

(b) 1 refinement step

Z
X
Y

(c) 2 refinement steps

for three-dimensional steady free-surface flows, the


explicit indicator based on the density discontinuity
has provided encouraging results for a very impressive reduction of the required CPU time,

X
Y

a systematic study of the influence of the width of


the fine grid region around the interface on the global
accuracy of the flow field needs to be done,
(d) 3 refinement steps

one may think that the emphasis put on the interface


is exaggerated because the sources of discretisation
errors may come from other regions of the flow,
the generalisation to 3D free-surface flows of an approach based on a-posteriori error estimation for all
the dependent variables may be a rigorous way to
clarify the previous uncertainty.

Figure 23: Locally refined meshes along the waterline


CONCLUSION AND PERSPECTIVES
This paper has described a modern free-surface capturing strategy implemented in an unstructured finite-volume
viscous flow solver. Special attention has been paid to

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Z
X
Y

Figure 24: Reference fine mesh along the waterline

z/L=0.001
0.5
0.4

Ngen=2

0.3

reached, which makes this approach robust and reliable.


A new adaptive grid refinement strategy has also been described and applied to the free-surface flow around the
Wigley hull. Very promising although preliminary results
have been obtained, showing that an approach linking the
use of unstructured adaptive grids with several levels of
local refinement to a free-surface capturing strategy may
be optimal in terms of robustness and accuracy. The future developments will aim at generalizing this approach
to arbitrary (i.e. non analytical) hull geometries and at
implementing an efficient parallelisation with a dynamic
load-balancing. In the long term, one cannot forget that all
these local grid adaption algorithms may be used to implement multi-grid techniques to solve the RANS equations
on adapted unstructured grids.

0.2

y/L

0.1

ACKNOWLEDGMENTS

The authors gratefully acknowledge the scientific committee of CINES (project dmn2050) and IDRIS (project
1308) for the attribution of CPU time.

-0.1
-0.2
-0.3
-0.4
-0.5

Fine
-0.5

0.0

x/L

0.5

1.0

REFERENCES
[1] R. Duvigneau and M. Visonneau, On the role
played by turbulence closures in hull shape optiJ Mar Sci
mization at model and full scale,
Technol, Vol. 8, 2003, pp. 1125.

(a) 2 generations of refinement vs reference grid

z/L=0.001
0.5
0.4

Ngen=3

[2] G.B. Deng and M. Visonneau, Comparison


of explicit algebraic stress models and secondorder turbulence closures for steady flows around
ships, Proc. 7th Int. Conf. on Numerical Ship
Hydrodynamics, 1999.

0.3
0.2

y/L

0.1
0
-0.1
-0.2
-0.3
-0.4
-0.5

Fine
-0.5

0.0

x/L

0.5

1.0

(b) 3 generations of refinement vs reference grid

Figure 25: Global view of the free-surface elevation with


different grids
the influence of the discretisation schemes of the transport
equation for concentration and it was found that, despite
the relatively fine grids used in this study, the role played
by the compressive property is still fundamental to get a
reliable simulation of the free-surface. A detailed comparison of computations versus experiments performed at
IIHR showed that a quite satisfactory agreement has been

[3] L. Larsson, F. Stern, and V. Bertram, Summary,


conclusions and recommandations of the Gothenburg 2000 Workshop, A Workshop on Numerical
Ship Hydrodynamics, L. Larsson, F. Stern, and
V. Bertram, eds., Gteborg, Chalmers University of
Technology, September 2000.
[4] P. Queutey, M. Visonneau, and Ferrant P., Numerical investigation of wave interaction with a
fixed vertical circular cylinder. The Thirteenth
International Offshore and Polar Engineering
Conference, Honolulu, Hawai, USA, May 2003.
[5] A. Hay, Etude de stratgies destimation derreur
numrique et dadaptation locale de maillages
non-structurs pour les quations de Navier-Stokes
en moyenne de Reynolds. PhD thesis, Universit de
Nantes, 2004.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

[6] A. Hay and M. Visonneau, Adaptive error control [18] S. Muzaferija and M. Peric,
Computation
of free surface flows using interfacetracking and
strategy : Application to a turbulent flow. AIAA
CFD Conference, Orlando, June 2003.
interface-capturing methods, vol. chap. 3, pp. 59100. Computational Mechanics Publications, WIT
[7] J. Longo and F. Stern, Effects of drift angle on
Press, Southampton, nonlinear water wave interacmodel ship flow, Experiments in Fluids, Vol. 32,
tion ed., 1998.
2002, pp. 558569.
[8] I. Demirdic and S. Muzaferija, Numerical method [19] G.B. Deng, R. Duvigneau, P. Queutey, and M. Visonneau, Assessment of turbulence model for ship
for coupled fluid flow, heat transfert and stress analflow at full scale, Comp. Mech., WCCM IV, Beiysis using unstructured moving meshes with cells of
jing, China, September 2004.
arbitrary topology, Comput. meth. Appl. Mech.
Eng., Vol. 125, 1995, pp. 235255.
[9]
[10]

[11]

[12]

[20] C. Farhat, C. Degand, B. Koobus, and M. Lesoinne,


Torsional springs for two dimensional dynamic unJ.H. Ferziger and M. Peric, Computational methods
structured fluid meshes, Computational Methods
for fluid dynamics. Berlin: Springer-Verlag, 1996.
in Applied Mechanics and Engineering, 1998,
H. Jasak, Error Analysis and Estimation for the
pp. 231245.
Finite Volume Method with Applications to Fluid
[21] W. K. Anderson and V. Venkatakrishnan, AeroFlows. PhD thesis, University of London, 1996.
dynamic design optimization on unstructured grids
B.P. Leonard, Simple high-accuracy resolution prowith a continuous adjoint formulation, Computers
gram for convective modelling of discontinuities,
and Fluids, Vol. 28, No. 4, 1999, pp. 443480.
International Journal for Numerical Methods in
Fluids, Vol. 8, 1988, pp. 12911318.
[22] Cooperative experiments on Wigley parabolic
models in Japan. Nineteenth R.C. International
P.H. Gaskell and A.K.C. Lau, Curvature comTowing Tank Conference (ITTC), 1983.
pensated convective transport:
SMART , a
new boundedness preserving transport algorithme,
International Journal for Numerical Methods in
Fluids, Vol. 8, 1988, pp. 617641.

[13] C.M. Rhie and W.L. Chow, A numerical study of


the turbulent flow past an isolated airfoil with trailing edge separation, AIAA Journal, Vol. 17, 1983,
pp. 15251532.
[14] H. Jasak, H.G. Weller, and A.D. Gosman, High
resolution nvd differencing scheme for arbitrarily unstructured meshes,
International Journal
for Numerical Methods in Fluids, Vol. 31, 1999,
pp. 431449.
[15] V. Prulj and B. Basara, Bounded convection
schemes for unstructured grids,
15th AIAA
Computational fluid dynamics conference, AIAA
paper 2001-2593, Anaheim, CA, 11-14 June 2001.
[16] C.W. Hirt and B.D. Nichols, Volume of fluid
(vof) method for the dynamics of free boundaries,
Journal of Computational Physics, Vol. 39, 1981,
pp. 449468.
[17] H. Jasak and H.G. Weller, Interface tracking capabilities of the inter-gamma differencing
scheme.. Internal Report, Mechanical Engineering
Department, Imperial College of Science, London,
February 1995.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics


St. Johns, Newfoundland and Labrador, CANADA, 8-13 August, 2004

Experimental Measurements for CFD Validation of the Flow


About a Submarine Model (ONR Body-1)
Paisan Atsavapranee, Thomas Forlini, Deborah Furey, John Hamilton, Scott
Percival, and Chao-Ho Sung
(The Naval Surface Warfare Center, Carderock Division)
ABSTRACT
Todays use of potential flow programs for the
prediction of the performance of a maneuvering
submarine is severely limited, as empirically-based
formulations must be used to account for viscous
effects around the hull. It has been shown that RANS
is sufficiently accurate for problems with simple
geometries and that the extension for six degree of
freedom (6DOF) simulations is promising. Largeeddy simulation (LES) has also shown promising
results, even though one may not expect realistic
applications to large-scale naval problems in the short
term. As a result, a clear need exists for experimental
hydrodynamic data in support of the development,
verification, and validation of physics-based CFD
prediction tools, such as 6DOF RANS and LES. In
the short term, application of a CFD prediction tool
would provide the naval architect with a means to
apply ship-specific force and moment coefficients in
support of potential flow submarine motion
predictions. Longer term, the use of 6DOF RANS or
LES would allow for a rigorous and direct
assessment of submarine trajectories in a wide range
of maneuvers and for many candidate designs,
allowing the field to be narrowed for cost-effective
detailed experimental evaluation.
Within the context of this need, a systematic
series of model tests was performed in order to
explore the viscous flow field around the hull of the
submarine model, ONR Body-1, and the associated
pressure distribution and the overall forces and
moments on the model. Using stereo particle-image
velocimetry (SPIV), flow patterns with all three
velocity components around the model have been
gathered and analyzed for three model
configurations: bare hull, bare hull with sail, and
fully-appended.
These flow-field measurements
were then used to correlate to the measurements of
pressure distribution and the overall forces and

moments on the model to explain the effects of


dominant flow features, such as the sail tip vortex on
the flow development and the cross flow separation
around the hull. This paper will also discuss
technical issues related to the application of SPIV in
support of viscous flow measurements in the towing
basins at NSWCCD. Among the issues to be
addressed are the overall system configuration,
imaging optics, and particle injection method.
INTRODUCTION
The Department of the Navy has a clear and urgent
need to accelerate the integration of Computational
Fluid Dynamics (CFD) techniques into the design
cycle of its submarine fleet. At present, CFD based
on solving the Reynolds-Averaged Navier-Stokes
(RANS) equations has matured to the point that most
complex turbulent flows can be computed. However,
its application to the design of maneuvering
submarines continues to be limited by a low level of
confidence in the solution accuracy, a lack of
robustness in the computational techniques, and a
high demand on computational memory and time.
Remedies are needed for all these difficulties, with
the establishment of confidence in the solution
accuracy being of particular importance. To achieve
this goal it is essential to have accurate and reliable
experimental data, which are especially tuned to the
needs of CFD validations and for the purpose of
improving CFD techniques. The six components of
forces and moments constitute the major information
needed for the design of maneuvering submarines.
However, these integrated quantities do not provide
details of the hydrodynamic state that more
fundamental data such as flow field and pressure
field can provide. In the instance where forces and
moments are predicted inaccurately, it is difficult to
pinpoint the reasons if no local hydrodynamic data is

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

available for validation. By providing correlation


with the integrated data, local hydrodynamic data
provide crucial understanding of the flow physics,
which could lead to the improvements of CFD.
The main objective of this study is to
perform fundamental submarine captive-model
experiments to obtain detailed hydrodynamic data
suitable for the validation of CFD techniques based
on RANS and LES. Both issues of stability of
motion and maneuverability are considered, with a
special emphasis on the major flow phenomenology
unique to a maneuvering submarine. Straight-line
experiments were performed in the NSWCCD
Shallow Water Basin, Towing Carriage #1 using the
submarine model, ONR Body-1 at a Reynolds
number of 1.3 x 107. Three types of measurements
were undertaken, flow field about the submarine
model utilizing stereo particle-image velocimetry
(SPIV) to obtain all three velocity components, static
pressure field at three axial cross sections on the
model using variable-reluctance differential pressure
transducers, and six components of forces and
moments with block gages on the two support struts.
The model consists of a Body-1 bare hull, a sail and
four stern appendages. Three model configurations
were tested: bare hull, bare hull with sail, and fullyappended.
BACKGROUND
There is an abundance of data measured in the David
Taylor Towing Carriages and the Rotating Arm Basin
on various submarine models. However, most of the
measured data include only the six components of
forces and moments, which are inadequate for CFD
validation by themselves as mentioned earlier.
Forces and moments are integrated quantities. By
coincidental cancellations of local solution errors, it
is possible that erroneous solutions may turn out to
give reasonable predictions of forces and moments.
It is also possible that a reasonable numerical
solution might give a poor prediction of forces and
moments. For example, a small error of pressure
distribution in the stern region may give a poor
prediction of moments because of the long distance
from the stern to the center of gravity of the
submarine. In addition, the measured data mentioned
above are not available to the entire CFD community.
The interaction between the sail vortical
flow and the hull separated flow is a very interesting
and important flow phenomenon, which in certain
cases could lead to undesirable out-of-plane forces
and moments, severely degrading the performance of
a maneuvering submarine. Accurate prediction of the
sail vortical flow alone is an enormous challenge for

CFD, because the combined diffusion due to


numerical schemes and turbulence models can
rapidly diffuse the sail tip vortex, resulting in
inaccurate predictions. However, the difficulty must
be overcome for CFD to be useful to the submarine
design community, and towards this goal accurate
measurement of the sail vortical flow by SPIV and
the investigation of hydrodynamic phenomena
leading to undesirable out-of-plane forces and
moments represent the main objectives in this work.
Recent work on the same ONR Body-1
model, performed in the Rotating Arm Basin at
NSWCCD, as described in Fu, et.al. [4], is a step in
the right direction to address some of these complex
issues. However, as discussed in Sung, et.al. [5],
which attempted to simulate the flow around the
model under the same condition, the available
experimental data are far from sufficient for adequate
CFD validation. The most serious shortcoming is
that only one experimental condition was
investigated. Six components of forces and moments
and the flow field around the submarine were
measured under a single set of non-dimensional
angular velocity r', roll, pitch and drift angles.
Consequently, the important information about the
trend in the changes of forces and moments as the
submarine attitude varies and the relationship to the
changes in the flow field cannot be compared with
CFD computations. This shortcoming is corrected in
the present work, in which data for several drift
angles were measured. The measured data with the
variations of are also important in assisting CFD to
predict accurately the change in the bow pitch-up
resulting from the interaction between the sail
vortical flow and the hull separated flow. As the drift
angle increases, the bow pitch-up moment increases
but levels off and may eventually become bow pitchdown. Accurate prediction of this trend is a very
challenging task for CFD.
ONR BODY-1
The submarine model used in this study is DTMB
model #5484 with a length of 5.27 m, a diameter of
0.47 m and configured as ONR Body-1. ONR
Body-1 is an unclassified generic submarine shape
composed of an axisymmetric body, sail and 4
identical stern appendages. The sail has a NACA
0014 section with a chord length of 0.57 m and an
aspect ratio of 0.27. It is faired to the hull and
located at x/L = 0.2. The stern appendages, all
identical in shape and size, are NACA 0018 foils
with a chord length of 0.18 m and an aspect ratio of
1.2, all set at zero degree angle of attack. A
boundary layer trip wire, with an OD of 1.5 mm, was

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

mounted at x/L = 0.05. The model has no propulsor


and instead is fitted with a faired tail cone in the aft
end.

streaming images at the full camera frame rate onto a


real-time disk array system with a capacity of 480
gigabytes. For system redundancy, images can also
be optionally captured to 1.8 gigabytes of RAM.

Figure 1: Elevation view of DTMB model #5484


PARTICLE IMAGE VELOCIMETRY (PIV)
PIV measurement has found wide usage in the
academic community, thus the general concept of
PIV/SPIV will not be described here and instead the
discussion will concentrate upon unique features that
allow successful application of PIV/SPIV in largescale hydrodynamic facilities at NSWCCD. The
SPIV system utilized in this study comprises of two
independent 2D PIV systems, which can be
configured for simultaneous two-component velocity
measurements or, as in this case, for stereo-PIV
(three-component) measurement in a single plane.
Specifics of the SPIV system are described in the
following paragraphs:
Figure 2: Submersible PIV cameras

Cameras
The cameras (Roper Scientific ES 4.0), with a spatial
resolution of 2048 pixels x 2048 pixels and a full
frame rate of 15 Hz, were fitted with 50 mm lens and
placed in submersible housings with an umbilical of
30 m in length. For the SPIV measurement in this
study, no Scheimflug mechanism was installed on the
cameras. Instead, the lens was stopped down to an fnumber of 4.0 to provide the necessary depth of field,
allowing cameras to focus on the entire measurement
plane. The camera housings utilize modular design
with interchangeable nose, center body, and main
bulkhead sections and incorporate remote focus and
moisture-sensing features. The electronics in the
camera housings are kept dry by purging the internal
volume with dry nitrogen periodically during the
experiment. Images from each camera are captured
and stored on a data-acquisition computer capable of

Laser sheets
Illumination was provided by a flash lamp-pumped
dye lasers, operating at 585 nm. The pulse laser
operates with a maximum optical energy output of 1
J/pulse 10 microsecond-long pulses at a maximum
rate of 15 Hz. The beam from the laser was coupled
into 600-micron optical fibers and formed into laser
sheets at the output end, using a series of beamforming optics placed in submersible housings. The
relatively long pulse duration (three orders of
magnitude longer than the typical 10 nsec pulse for a
Nd:YAG laser) makes it possible to launch the beam
into the fiber. The usage of fiber optics greatly
facilitates the placement of laser sheets, as mirrors
and open beams are eliminated from the optical
setup, with acceptable performance in terms of

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

optical loss through the fiber. Roughly 700 mJ/pulse


is recorded at the output end of a 30m long fiber
when the laser is operating at full power.
Seeding
Silver-coated hollow glass spheres with a mean size
of 40 microns and a specific gravity of approximately
1.0 was used as flow tracers in this study. The dry
powder was premixed into a tank of water and
injected using seeding manifolds into the basin along
the model tow path.
Typically, seeding was
performed after waiting a certain period (45 minutes)
since the end of a test run to allow the fluid motions
in the tow basin to dissipate.
SHALLOW WATER BASIN AND TOWING
CARRIAGE #1
The experiment was conducted in the deep-water
section of the Shallow Water Basin using the Towing
Carriage #1 at NSWCCD. The deep-water section of
the basin is approximately 271 m long, 15.5 m wide,
and 6.7 m deep. The basin is spanned crosswise by
Towing Carriage #1, which rests along the main rails
and can traverse at a maximum speed of 9.3 m/s. The
accuracy of towing speed is better than 0.005 m/s.

Figure 3: CAD rendering of the test setup in the


Shallow Water Basin at NSWCCD

TEST SETUP
The ONR Body-1 model was held, with the sail
mounted sideway, rigidly to the towing carriage by
two vertical struts, placed 1.83 m apart. The two
struts were fixed to a planar motion mechanism,
which allowed precise adjustments of the hull drift
angle. During a test run, the model was towed past
the stationary SPIV setup, set in place in the middle
of the basin. The submersible cameras and laser
sheet optics were fastened to an array of support
structures on a submersible hydraulic table. The
hydraulic table is a modified manlift, equipped with
pressurized sealed lines and an upper deck with a
rotating platform. The table allowed the vertical
position of the entire system to be set and easy
accessibility to the instruments during rigging. The
two cameras viewed a common area of interest but
from two different perspectives. SPIV calibration
was achieved by taking pictures of a calibration grid
at two axial positions 15.24 mm apart. Figure 3
illustrates a CAD rendering of the test setup (the
hydraulic table, except for the rotating platform, is
not shown), while Figure 4 is a photograph of the
overall setup.

Figure 4: Photograph of the test setup in the


Shallow Water Basin
TEST PROCEDURES
For each test condition, the model was adjusted to the
desired drift angle by using the planar motion
mechanism as a tilt table (pitch down on the tilt table
means positive drift angle on the model). Before the
start of each test run, a cloud of seeding particles
were laid within the measurement volume with the
aid of three seeding manifolds. The carriage was
then brought to the desired test speed, and the model
was towed past the stationary SPIV setup, setting off

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

2.00E-04

0.00E+00

-2.00E-04

X'

an opto-electronic trigger and starting the PIV data


acquisition. PIV images were collected at the rate of
5.88 image pairs/sec. A significant waiting period of
about 45 minutes between test runs ensured that any
fluid motion in the basin had adequately dissipated.
For each condition, at least 10 repeat runs were
performed.

Fully Appended
Hull w Sail
Bare Hull

-4.00E-04

-6.00E-04

FORCE AND MOMENT MEASUREMENTS


-8.00E-04

M =

-10

-5

10

15

20

(degrees)

Figure 5: Longitudinal force coefficient

1.40E-02
1.20E-02
1.00E-02
8.00E-03
6.00E-03
Fully Appended
Hull w Sail
Bare Hull

4.00E-03
2.00E-03
0.00E+00
-2.00E-03
-4.00E-03
-6.00E-03
-10

-5

10

15

20

(degrees)

Figure 6: Lateral force coefficient

4.50E-03
4.00E-03
3.50E-03

(Z 1 + Z 2 ) (Z 1 + Z 2 )0

(1)

1
U 2 L2
2

3.00E-03
2.50E-03

Z'

Z =

-1.00E-03

Y'

A forward and an aft set of block gages were used to


measure the longitudinal, lateral, and normal force
components with respect to the body axes, denoted X,
Y and Z, respectively. The forward and the aft set of
block gages were attached to a channel inside the
model through two gimbals.
The longitudinal
distance between the centerline of the two gimbals
was 1.83 m. The origin of the body coordinate
system or reference point was located 2.35 m aft of
the forward perpendicular along the centerline of the
hull. The pitching and yawing moments about the
reference point, M and N, were determined from the
difference in the measured reaction forces at each
gimbal, multiplied by one-half the spacing between
the two gimbals, Lh. The roll moment, K, was
measured by a strain gage unit located at the forward
strut.
Before a run was made, small nonzero
forces and moments present at each gage (electrical
zeros) were measured and recorded so that they could
be removed from the underway force and moment
data. The underway force and moment data were
then non-dimensionalized. Representative equations
for normal force, Z , and pitching moment, M ,
are shown in Equations (1) and (2), and similar
equations apply for the remaining quantities,

[(Z 2 Z1 ) (Z 2 Z1 )0 ] Lh

Fully Appended
Hull w Sail
Bare Hull

2.00E-03
1.50E-03

(2)

1
U 2 L3
2

1.00E-03
5.00E-04
0.00E+00
-5.00E-04
-10

where Z1 and Z2 are the normal force measurements


from the forward and aft block gages, respectively.
The subscript 0 indicates the values of the
electrical zeros, and the prime symbol denotes
dimensionless values. Figures 5-10 present the
block-gage data for all six components of forces and
moments.

-5

10

(degrees)

Figure 7: Normal force coefficient

Copyright National Academy of Sciences. All rights reserved.

15

20

Twenty-Fifth Symposium on Naval Hydrodynamics

2.00E-04

1.50E-04

K'

1.00E-04

Fully Appended
Hull w Sail
Bare Hull

5.00E-05

0.00E+00

-5.00E-05

-1.00E-04
-10

-5

10

15

20

(degrees)

Figure 8: Rolling moment coefficient

5.00E-04

planes along the hull outer surface at x/L = 0.318,


0.579, and 0.886, respectively for a total of 46
pressure taps or measurement locations distributed
azimuthally around the three rings. The axial
location x/L = 0.318 is only 0.014*L aft of the
trailing edge of the sail, so measurement in this plane
will elucidate the effects of the sail on the pressure
distribution immediately aft. The measurement at the
axial location x/L = 0.579 represents typical pressure
distribution that may be found along the parallel midbody aft of the sail. And the measurement at the last
station is 0.032*L forward of the leading edge of the
stern planes and will provide insights into the flow
going into the stern appendages. For the presentation
of pressure distribution measurement around the hull,
zero degree starts on the crown of the model and
increases in the clockwise direction looking forward
(see Figure 12). Figures 13-18 present the pressure
distribution data for various cases.

4.00E-04

3.00E-04

2.00E-04

M'

Fully Appended
Hull w Sail
Bare Hull

1.00E-04

0.00E+00

-1.00E-04

-2.00E-04
-10

-5

10

15

20

(degrees)

Figure 9: Pitching moment coefficient

3.00E-03
2.50E-03

Figure 11: Pressure measurement system

2.00E-03
1.50E-03

N'

1.00E-03
5.00E-04

Fully Appended
Hull w Sail
Bare Hull

0.00E+00
-5.00E-04
-1.00E-03
-1.50E-03
-10

-5

10

15

20

(degrees)

Figure 10: Yawing moment coefficient


PRESSURE MEASUREMENT
A rack of 12 Validyne DP15 variable-reluctance
differential pressure transducers (Figure 11) were
used to measure the static pressure at three axial cross

Figure 12: Schematic showing cross flow component


on hull

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

0.1
0.1

0.05

0.05

-0.05

Cp

-0.05

Cp

-0.1
-0.15

Beta=18
Beta=12
Beta=6
Beta=0

-0.1
-0.15
-0.2

Fully Appended
BH w Sail
Bare Hull

-0.2
-0.25

-0.25
-0.3

-0.3

-0.35
0

-0.35
0

30

60

90

120

150

180

210

240

270

300

330

30

360

60

90 120 150 180 210 240 270 300 330 360

Angle, degrees (CW looking Fwd)

Angle, degrees (CW looking Fwd)

Figure 13: Pressure distribution at x/L = 0.318,


12 degrees

Figure 16: Pressure distribution at x/L = 0.318, fullyappended configuration

0.1
0.1

0.05

0.05

-0.05

Cp

Cp

-0.05
-0.1
-0.15

Beta=18
Beta=12
Beta = 6
Beta = 0

-0.1
-0.15
-0.2

-0.2

Fully Appended
BH w Sail
Bare Hull

-0.25

-0.25
-0.3

-0.3

-0.35
-0.35
0

30

60

90

120

150

180 210

240

270

300

330

360

30

Angle, degrees (CW looking Fwd)

Figure 14: Pressure distribution at x/L = 0.579,


12 degrees

90

120 150 180 210 240 270 300 330 360

Figure 17: Pressure distribution at x/L = 0.579, fullyappended configuration

0.1

0.1

0.05

0.05

-0.05

-0.1

Fully Appended
BH w Sail
Bare Hull

-0.15
-0.2

Cp

-0.05

Cp

60

Angle, degrees (CW looking Fwd)

Beta=18
Beta=12
Beta = 6
Beta = 0

-0.1
-0.15
-0.2
-0.25

-0.25
-0.3

-0.3

-0.35

-0.35
0

30

60

90

120 150 180 210 240 270 300 330 360

Angle (CW looking Fwd)

Figure 15: Pressure distribution at x/L = 0.886,


12 degrees

30

60

90

120 150 180 210 240 270 300 330 360

Angle, degrees (CW looking Fwd)

Figure 18: Pressure distribution at x/L = 0.886, fullyappended configuration

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

SPIV MEASUREMENT
A small sample of SPIV measurement around the
ONR Body-1 at a hull drift angle of 12 degrees is
shown in Figure 19 for the bare hull case and in
Figure 20 for the fully-appended case. These results
include only 1/64 of the numbers of velocity vectors
obtained with the measurement, and only every other
SPIV velocity vector field is shown.
RESULTS AND DISCUSSION
In-plane forces and moments
Figures 5-7 present the force coefficients data as a
function of drift angle, for all three model
configurations, while Figures 8-10 show the plots of
the moment coefficients. The set of curves in Figure
6 show that the lateral force on the bare hull increases
monotonically with drift angle.
This trend is
expected since the cross flow velocity component
increases with drift angle and contributes to an
increasingly larger lateral force on the hull. The
presence of the sail and stern appendages contributes
even more to the magnitude of the lateral force due to
the fact that they act as lifting bodies at an angle of
attack.
As for in-plane moments, Figure 10 shows
that the highest yawing moment is exhibited in the
bare hull with sail configuration. This is obviously
due to the presence of the lift force on the sail acting
forward of the center of gravity, contributing to a
positive yawing moment. Even though a fully
appended model experiences the highest lateral force
for all drift angles, the yawing moment on a fully
appended model is actually the lowest of all three
configurations, due to the fact that the forces on the
sail and rudders create opposite signs of yawing
moments and act to counterbalance each other around
the center of gravity of the model.
Out-of-plane forces and moments
Figure 7 along with Figure 9 show that the bare hull
model experiences little normal force and pitching
moment at all drift angles. This is entirely expected
since the bare hull model is azimuthally symmetric
around the centerline of the model, and cross flow
velocity on the bare hull model being at a drift angle
can only exert in-plane forces. The flow field around
the bare hull model at angles of drift is expected to be
symmetric deck and keel sides. An example of SPIV

plots for the bare hull model at = 12 degrees is


presented in Figure 19.
On the other hand, for the bare hull with sail
and fully-appended configurations, the model
experiences a significant normal force (downward)
and pitching moment (pitch up) at high drift angles.
However, when the drift angle increases past a
certain point, the rate of increase in the downward
force is observed to taper off (towards = 18
degrees), and the pitching moment actually starts to
decrease (around = 13 degrees).
Pressure distribution data at the two forward
axial cross planes (x/L = 0.318 and x/L = 0.579)
strongly reinforce this observation. It is observed in
Figures 13 and 14 that, at = 12 degrees, the
presence of the sail (and the sail tip vortex) tends to
decrease the pressure along the keel side of the model
(from 90-270 degrees measuring from the deck
clockwise, looking forward) and to increase the
pressure along the deck side of the model, resulting
in a positive normal force and pitching moment. At
increasing drift angles, up to = 18 degrees, this
trend of increasing positive normal force and pitching
moment continues in the regions immediately aft of
the sail and along the parallel mid-body. However,
the trend in the pressure distribution at the aft most
axial cross plane (x/L = 0.886, Figure 18) for the
fully-appended model shows a different trend. With
increasing drift angles from = 0 degrees to = 12
degrees, the pressure on the deck and keel side is
both observed to decrease almost uniformly. At the
highest drift angle of = 18 degrees, the pressure
along the keel side of the model at the aft end is still
decreasing with drift angle; however, the pressure on
the deck side exhibits an even more significant
decrease, which would result in a counter-balancing
negative normal force and pitching moment to cancel
the positive normal force and pitching moment
immediately aft of the sail and along the parallel midbody. In other words, the pressure measurement on
the bare hull with sail and fully-appended model
seems to be consistent with the observed trends of
increasing positive normal forces and pitching
moments at moderate angles of drift and the tapering
off of the normal force and the decrease of pitching
moment at the higher drift angles.
A hypothesis of the physical phenomenon
leading to the out-of-plane forces and moments and
the behavior of the normal force and pitching
moment at moderate to high drift angles is presented.
At angles of drift, the presence of the sail to the
oncoming cross flow creates a strong sail tip vortex
that trails along the hull aft of the sail trailing edge,
as shown in the SPIV plots in Figure 20. For
moderate drift angles where cross flow separation is

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

not severe, the presence of the sail tip vortex induces


a circulation around the hull, which is analogous to
the presence of a mirror vortex of the opposite sense
inside of the hull. For both positive and negative
drift angles, the oncoming cross flow, coupled with
the induced circulation, creates a lift force (magnus
effect) on the hull aft of the sail trailing edge in the
positive Z direction (downward). Since most of this
lift force acts upon the hull aft of the sail trailing edge
and C.G., the resultant pitching moment is positive
(pitch-up) for both positive and negative drift angles.
As the drift angle increases, the effect of
cross flow separation becomes more severe,
especially towards the aft end of the model. The
induced circulation due to the presence of the sail tip
vortex would also act to rotate the pattern of cross
flow separation around the hull. This is evident in
the SPIV plots presented in Figures 19 and 20. In the
bare hull case, where no sail tip vortex is present,
separated flow region is observed on the lee side of
the model. However, with the presence of the sail tip
vortex, the separated flow region appeared to have
rotated counter-clockwise (looking forward). The
result is that the stagnation region (area of high
pressure) would now be rotated towards the keel side,
while the separated flow region (area of low
pressure) would be rotated towards the deck side.
This rotation of the separation pattern, coupled with
induced velocity from the sail tip vortex, appears to
be responsible for the large dip in surface pressure
along the deck side/windward part of the model (see
Figure 18). This effect results in a negative (upward)
normal force and pitching moment (pitch-down) and
acts to counterbalance the lift on the hull due to
magnus effect.

phenomena affecting the out-of-plane forces and


moments come into play. For moderate drift angles
where cross flow separation is not severe, the
presence of the sail tip vortex induces a circulation
around the hull, creating a lift force (magnus effect)
on the hull aft of the sail trailing edge in the positive
Z direction (downward), resulting in a net positive
pitching moment. As the drift angle increases, and
the effect of cross flow separation becomes more
severe, the induced circulation due to the presence of
the sail tip vortex would also act to rotate the
pattern of cross flow separation around the hull,
which results in a counterbalancing negative
(upward) normal force and pitching moment (pitchdown), explaining the tapering off in the rate of
increase of normal force and the decrease of pitching
moment on the model at high angles of drift.
REFERENCES
1.

Meier, H.U. and H.P. Kreplin, Experimental


Investigation of the Boundary Layer Transition
and Separation on a Body of Revolution,
Zeitschrift
fuer
Flugwissenschaft
und
Weltraumforschung, Vol.4, No.2, 1980, pp.6571.

2.

Wetzel, T.G., R.L. Simpson and C.J. Chesnakas,


Measurement of Three-Dimensional Crossflow
Separation, AIAA Journal, Vol.36, No.4, 1998,
pp.557-564.

3.

Huang, T. T, H.L. Liu and N.C. Groves,


Experiments of the DARPA Suboff Program,
DTRC/SHD-1298-02 December 1989.

4.

Fu, Thomas C., Paisan Atsavapranee and David


E. Hess, PIV Measurements of a Turning
Submarine Model (ONR BODY-1) Part 1:
Experimental
Setup,
NSWCCD-50-TR2001/036, 2001, Hydromechanics Directorate,
Research and Development Report.

5.

Sung, Chao-Ho, Ming-Yee Jiang, Bong Rhee,


Scott Percival, Paisan Atsavapranee and InYoung Koh, Validation of the Flow Around a
Turning Submarine, The Twenty-Fourth
Symposium on Naval Hydrodynamics, July 813, 2002, Fukuoka, Japan

CONCLUSIONS
A systematic series of model tests to explore the
viscous flow field around an unclassified submarine
model ONR Body-1 at angles of drift has been
performed. Using stereo particle-image velocimetry
(SPIV) measurements, flow patterns with all three
velocity components around the model have been
gathered and analyzed for three model
configurations: bare hull, bare hull with sail, and
fully-appended.
These flow-field measurements
were then used to correlate to the measurements of
the pressure distribution and the overall forces and
moments on the model to explain the effects of
dominant flow features such as the sail tip vortex on
the flow development and the cross flow separation
around the hull. It was found that the presence of the
sail tip vortex exerts a significant influence on the
downstream flow development and pressure
distribution. It is hypothesized that two competing

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Longitudinal
Velocity
u/U

12 Drift
Bare Hull

1.8
1.6
1.4
1.2
1.0
0.8
0.6

x/L =
0.634

0.4
0.2

Figure 19 (a): Bare hull, = 12 degrees

(d)

x/L =
0.284

x/L =
0.808
(e)

(b)

x/L =
0.459

x/L =
0.983
(c)

(f)

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Vorticity, x
(1/sec)

12 Drift
Bare Hull

40
24
8
-8

x/L =
0.634

-24

(g)

(j)

x/L =
0.284
x/L =
0.808
(h)

(k)

x/L =
0.459

x/L =
0.983
(i)

(l)

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

x/L =
0.634

Longitudinal
Velocity
u/U
1.8
1.6
1.4

12 Drift
Fully Appended

1.2
1.0
0.8
0.6
0.4
0.2

Figure 20 (a): Fully-appended, = 12 degrees

(d)

x/L =
0.808

x/L =
0.284

(e)

(b)

x/L =
0.983
x/L =
0.459

(c)

(f)

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Vorticity, x
(1/sec)
140

12 Drift
Fully Appended

100
60

x/L =
0.634

20
-20
-60

(g)

(j)

x/L =
0.284

(h)

x/L =
0.808

(k)

x/L =
0.459

x/L =
0.983

(i)

(l)

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Jerome P. Feldman
Naval Surface Warfare Center, Carderock Division,
USA
This paper provides valuable experimental
data for the continued development and validation of
physics-based six degree-of-freedom ReynoldsAveraged Navier-Stokes (RANS) and large eddy
simulation (LES) computer codes. These codes are
used for predicting the hydrodynamic forces,
moments, and pressure distributions that are
developed over the hull and appendages of
submarines.
Reference (1) and other papers indicate that
the hydrodynamic force distribution at small angles
of attack or angles of drift can be predicted
satisfactorily by potential flow theory over about 80
percent of the length of the hull from the nose. The
remaining portion of the hull increasingly deviates
from potential flow due to viscous effects. Potential
flow theory indicates that the nondimensional yawing
moment derivative Nv for the unappended hull is
equal to X u& 'Yv& ' which is the difference between the
longitudinal and lateral nondimensional added mass
derivatives. This destabilizing moment is reduced by
the addition of rudders and other stern appendages
and increased by the addition of forward appendages.
It would be interesting to compare the experimental
value of the yawing moment coefficient shown in
Figure 11 with the value calculated using potential
flow theory.
In a turning maneuver the local angle of drift
x varies along the length of the hull according to
x
tan x = tan
R cos
where is the angle of drift at the longitudinal
location of the center of buoyancy, x is the distance
from the center of buoyancy (negative aft), and R is
the steady radius. Hence, lift is developed on the
bridge fairwater due to its local angle of drift. There
is a bound vortex at the quarter chord of the bridge
fairwater and a tip vortex that trails aft. An image
vortex is located inside the hull. This system of
vorticity sets up circulation around the hull aft of the
sail. The combination of the circulation and the local
cross flow velocity causes the hydrodynamic pressure
over the deck to be greater than the hydrodynamic
pressure over the keel. This difference in pressure
results in a downward normal force over the hull aft
of the sail and an up pitching moment as seen in
Figures 8 and 10.
The lift on the sail YF is proportional to
Ub and the normal force on the hull ZH is

proportional to va. The symbol is the density of


water, is the strength of the tip vortex, U is the
forward speed of the submarine, v is the cross flow
velocity, b is the span of the sail and its image, and a
is the distance from the sail to the aft end of the
submarine. The ratio of these forces is
ZH
v a
a
=
=
YF U b
b
Since the lateral force on the sail varies linearly with
the angle of drift for moderate angles as can be seen
in Figure 7, the normal force on the hull varies with
the angle of drift squared as can be seen in Figure 8.
Flow will separate from the leeward side of
the afterbody when the local angle of drift becomes
relatively large. The separation of the flow from the
afterbody will be strongly asymmetrical due to the
combination of the circulation around the hull and the
cross flow velocity. The strength of the vorticity shed
from the lower leeward side of the afterbody (near
the keel) will be greater than the strength from the
upper leeward side (near the deck) due to higher local
cross flow velocity.
It is assumed that the vorticity shed from
both the upper and lower leeward sides of the
afterbody induces additional circulation around the
afterbody. The circulation induced by the vorticity
shed from the lower leeward side is in the opposite
direction to the circulation induced by the vorticity
shed from the sail. Hence, at relatively large angles of
drift the normal force and pitching moment will
begin to decrease as indicated in Figures 8 and 10. If
the vorticity shed from the lower leeward side of the
afterbody is strong enough, it is assumed that this
may cause a down pitching moment on the
submarine.
It is recommended that similar experiments
be performed on the rotating arm to investigate the
effect of the variation of local angle of drift along the
hull.
REFERENCE
1. Van Randwijck, E. and J. Feldman, "Results of
Experiments with a Segmented Model to Investigate
the Distribution of the Hydrodynamic Forces and
Moments on a Streamlined Body of Revolution at an
Angle of Attack or with a Pitching Angular
Velocity," NSWCCD-50-TR--2000/08, February
2000.
AUTHORS REPLY
The discussers comments added valuable
insights to the discussion already presented in the
paper. The authors strongly agree that similar

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

experiments should be performed in the Rotating


Arm Basin to investigate the effect of the variation of
local angle of drift along the hull. Even though the
straight-line experiment contains relevant physics of
cross flow separation, a submarine in a steady turn
would experience varying local angle of drift along
the hull. In a rotating arm experiment, for the same
drift angle at the sail as in the steady case, the local
drift angle at the aft end would be more severe.
Performing experiments at different and r would
reveal more complex physics and improve our
understanding of the relevant hydrodynamics.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Joseph Gorski
Naval Surface Warfare Center, Carderock Division,
USA
The authors are to be commended on getting
such a comprehensive data set including forces,
pressure, and velocity under the same operating
conditions as such complete measurements are
invaluable for understanding physics as well as good
CFD validation. The authors show PIV data for a 12
degree drift angle. Is PIV data also available at other
angles of drift, which might help verify the authors
hypothesis regarding the out of plane forces? Also,
since the model is being towed past the stationary
SPIV setup how stationary is the vortex on
subsequent runs and how different is the ensemble
averaged flow field, which is what is needed for
RANS
comparison,
to
the
instantaneous
measurements?
AUTHORS REPLY
SPIV measurements were conducted at 0, 6,
12, and 18 degree drift angle for the fully appended
configuration. At 12 degree angle of drift, SPIV
measurements were also performed with the bare hull
and bare hull with sail configurations. For each of
the six conditions, 10 experimental realizations with
the model towed through the stationary SPIV were
performed. In order to gauge the consistency of the
data, the velocity residual for each of the velocity
components was calculated. The velocity residual for
u for N averages is defined as the following:

Ru , N =

(u
u

avg , N

uavg , N 1 )

AllFrames AllVectors

avg , N =10
AllFrames AllVectors

The residual for v and w are defined in a


similar manner. The plot of the velocity residual as a
function of N for the fully-appended case at 12
degree drift angle is shown below. As can be seen,
all the residual components decrease to very low
values as N approaches 10, indicating that the total
average represents a good comparison for CFD
steady flow field calculations. At N=10, the u, v, and
w components of the residual are 0.0096, 0.0045, and
0.0019, respectively.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics


St. Johns, Newfoundland and Labrador, CANADA, 8-13 August 2004

INVESTIGATION OF THE
TURBULENT BOUNDARY LAYER FLOW
ON A MICROFILAMENT TOWED ARRAY

D. Furey,1 K. Cipolla, 2 P. Atsavapranee, 1


(1Naval Surface Warfare Center, 2Naval Undersea Warfare Center, United States)
ABSTRACT
By understanding the nature of turbulent
boundary layers on long, thin cylinders, the
performance of acoustic sensors used for
surveillance and tactical operations can be
improved. Towed array detection performance is
a function of several factors including ambient
acoustics, array gain, flow noise, and signal
processing. To advance existing systems the
primary goal is to increase the system gain by
introducing longer apertures. Longer arrays of a
conventional design lead to additional stowage
requirements that are difficult to implement. The
development of very small diameter towed array
technology is required. Therefore, the results of
this research will directly contribute to the work
being done on small diameter arrays and multiline array systems.
This is an experimental study to evaluate
the development of the boundary layer thickness,
, and momentum thickness, , along long thin
cylinders or arrays (L/a=1.5*105 and 3.0*105,
a=radius). The experiments use proven test
methods in conjunction with Stereo Particle
Image Velocimetry (SPIV) measurement
techniques to evaluate the flow in the boundary
region of a small diameter array towed in the high
speed towing basin at Naval Surface Warfare
Center Carderock Division (NSWCCD). The
array cylinders were approximately neutrally
buoyant and were towed at speeds ranging from 3
to 15 m/s. The objectives for this experiment are:
1. Determine the streamwise development of w
and in axisymmetric turbulent boundary layers
where >> a, using drag measurements;
2.Determine the streamwise development of the
axisymmetric turbulent boundary layer flow for
cases where >> a. Results for a=0.45 mm and
1.25 mm and lengths between 15 m and 152 m
will be presented. Tow speeds of 3.9, 7.7, 12.9
and 15.5 m/sec were tested with images acquired
over the entire length of the cylinders. Drag
measurements revealed that the shear on the wall
is large and that the momentum thickness grows

slowly when compared to flat plate boundary


layers. The velocity field data shows that the
boundary flow remains turbulent over the entire
length of the array and that the turbulent profile is
significantly different than that of the flat plate
boundary layers.
INTRODUCTION
Boundary layer flows have been studied to
understand and characterize the effects of fluid
dynamics adjacent to solid boundaries. The viscous
effects and dynamic properties of this flow region
can have a significant impact on the performance of
some systems. Flat plate boundary layers have
received significant attention due to the simple
geometry of the flow. More recently attention has
been given to geometries with transverse curvature of
the boundary. This boundary geometry is relevant
for flows around submerged bodies, missiles, ship
hulls and, in particular, towed arrays.
A towed array is a long cylindrically shaped,
flexible hose-like acoustic sensing device. They are
typically hundreds of feet in length and are deployed
trailing behind a towing vessel. Acoustic array
performance is influenced by several factors
including ambient acoustics, array gain, flow noise,
and signal processing.
Improvements can be
achieved through signal processing advancements as
well as implementing systems with higher array gain.
Higher array gain can be achieved through longer
arrays (longer aperture) or through volumetric
configurations (volumetric apertures).
Both
implementations require more linear feet of acoustic
array. Using longer conventional arrays requires
additional stowage capabilities, which is difficult and
sometimes impossible to accomodate. Therefore, the
development of very small diameter towed arrays is
required. Very small diameter towed arrays are being
developed with diameters between 0.5 and 3 mm
with lengths on the order of 200 m and longer to
achieve improved system performance.
Given that acoustic arrays are listening for
pressure variations in the ocean, the inherent
fluctuations in a turbulent boundary layer flow can

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

have a significant impact on system performance.


The operation of these acoustic sensors can therefore
be improved by understanding the nature of turbulent
boundary layers.
However,
boundary layer
development on surfaces with severe transverse
curvature, cases where the boundary layer thickness
is on the order of or larger than the radius of
transverse curvature, are not fully characterized. This
research will investigate turbulent boundary layer
development along high aspect ratio cylinders as it is
related to towed array performance.
Relevant
boundary layer parameters will be determined from
these results and used to scale characteristic wall
pressure spectra (See Figure 1). These spectra can be
utilized for developing flow noise models as applied
to acoustic signal analysis for towed arrays. Further,
this data can be used to develop turbulence models
for axisymmetric turbulent boundary layers. By
combining the drag measurement results and velocity
field results, a relationship between boundary layer
and momentum thickness for axisymmetric turbulent
boundary layers will be developed.

Figure 1: Nondimensional wall pressure spectra


scaled on mixed variables. (Keith, et. al. 1992)
BACKGROUND
Initial boundary layer research largely concentrated
on investigating the development of two dimensional
planar turbulent boundary layers. However, while
the axisymmetric boundary layer is also two
dimensional, it is more complicated due to the
additional effect of the transverse curvature of the
surface. In addition, the boundary layer equations
were developed for cases where the boundary layer
thickness is much smaller than the characteristic
length of the body. For cases of long aspect ratio
small diameter cylinders, the boundary layer
equations become incompatible as the boundary layer
thickness becomes much larger than the model

radius, the characteristic length of the body.


Experiments have been designed to determine the
scaling parameters that are most appropriate for
describing this flow. Lueptow (1988) provides a
complete review of work on axisymmetric boundary
layers (3). To date, there is no model that has been
developed to include general transverse curvatures.
Experiments designed to investigate the effects of
transverse curvature on the development of turbulent
boundary layers interrogate finite regions of the
parameter space, Rea, /a, based on the experimental
limitations of the apparatus and test facility.
Specifically, values of 0.1 /a100 and 100
Rea106 have been used in the previous work. (See
references 1-3, 5, 8-14).
In relation to thin line towed arrays, a
considerable amount of work has been done to
investigate the development of the boundary layer on
small diameter cylinders (References 1-3). In brief,
tests on small diameter cylinders (0.9 to 2.5 mm)
have been performed to characterize various relevant
flow quantities.
The initial microfilament
experiments conducted by Naval Undersea Warfare
Center, NUWC, were designed to measure the mean
wall shear stress and were used to evaluate the
resulting drag coefficients for comparison to
established flow quantities, i.e. theoretical flat plate
flow. The results supported the presence of a
turbulent boundary layer on the arrays and were used
by Cipolla et. al. (2003a) to develop analytical
methods relating the tangential drag coefficient to the
momentum thickness at the end of the line for a fully
developed axisymmetric turbulent boundary layer.
The results indicated that for a fixed diameter array
and tow speed, the momentum thickness at the end of
the array increases with increasing length. Therefore,
through systematic variation of array diameter,
length, and speed, an accurate measure of the spatial
growth of the momentum thickness can be
determined. These measurements were done in a
variety of tests including basin tests at the David
Taylor Model Basin (DTMB), trial evaluations (river
tests), flow visualization tests at Langley, NUWC
Quiet Water tunnel, and Rutgers Free Surface Water
Channel. The data verify that the momentum
thickness in an axisymmetric boundary layer grows
more slowly than that of a flat plate, while the mean
wall shear stress is 2-4 times greater.
PROJECT DESCRIPTION
In addition to the momentum thickness data, direct
measurements of the boundary flow can be used to
evaluate the effects of transverse curvature on
boundary layer development. Through the joint
efforts of NSWC and NUWC, velocity measurements

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

have been taken of the boundary flow along high


aspect ratio cylinders with applications to towed
array technology. Stereo Particle Image Velocimetry
(SPIV) measurements were taken in discrete
transverse planes along a small diameter array towed
in the High Speed Basin at DTMB. This data will
provide the necessary information to further evaluate
the boundary layer parameters as a function of axial
location.
There are two primary objectives for this
work. The first objective is to measure the mean wall
shear stress in the turbulent boundary layer of
cylinders with diameters of 0.9 and 2.5 mm and up to
150 m in length for comparison to previous results
and to add to the existing dataset. From the
measurements of total drag, the average tangential
drag coefficient, Cd, is calculated. This data is then
used to determine the momentum thickness at the end
of each cylinder, (L). The second objective is to
determine the growth of the boundary layer (x)
along the array and to evaluate and characterize the
turbulent quantities in the flow. Three-dimensional
velocity field measurements were obtained over the
length of the cylinders with a stationary SPIV system.
The images are processed using cross correlation
techniques and gradient methods. The boundary
layer profiles and turbulent quantities are evaluated
using circumferential averaging of the velocity data
around the array. Combining the drag results and the
velocity results will lead to a more complete
characterization of the boundary layer flow and to a
relationship between the momentum thickness and
the boundary layer thickness for axisymmetric
turbulent boundary layers. (Note, throughout this
discussion the terms array, model and cylinder are
used interchangeably.)
DRAG MEASUREMENTS AND MOMENTUM
THICKNESS DETERMINATION
The experiments by Cipolla et. al. [2003a, 2003b]
utilize an analytic approach to extract boundary layer
information using spatially and temporally averaged
measurements. The total drag on different lengths of
axial cylinders was measured for several speeds and
different length cylinders. The drag coefficient was
then determined based on the following relationship:

Cd =

D
1
U 2 As
2

(1)

where D is total drag, is density, U is tow speed,


As is the surface area of the cylinder, and Cd is the
drag coefficient.

The relationship developed between


momentum thickness and drag coefficient is derived
from a control volume analysis of the momentum
balance for the geometry (See Figure 2) [Cipolla
2003a].

Cylinder
Figure 2: Control volume for cylindrical boundary
layer analysis.
The presence of the boundary layer causes a
reduction in the mass flow rate as well as momentum
flux through the boundary of the control volume.
The relationship between momentum thickness and
the velocity profile for a cylindrical boundary layer
is:

2 + 2 a = 2

a +

u (r ) u (r )
r dr (2)
1
U
U

The cylinder imposes a shear force on the inner


surface of the fluid control volume. This can be
related to the average shear on the wall, w, and to the
momentum loss in the flow:

1 a + u (r ) u (r ) r
1
w
dr = Cd
1
=
2

2
U L a U U a

(3)

Relating equations (2) and (3) results in the following


relationship between the drag coefficient and
momentum thickness at the end of the cylinder:
[Cipolla et. al 2003a]

2 + 2a aLCd = 0

(4)

This relationship is used to extract momentum


thickness values from the drag measurements.

STEREO PARTICLE IMAGE


VELOCIMETRY (SPIV)
Stereo Particle Image Velocimetry (SPIV) is a flowfield measurement technique that provides an
instantaneous distribution of three velocity
components over an entire two-dimensional
measurement plane. Two high resolution cameras
are positioned close to each other and oriented to

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

record images of the same plane of interest which is


defined by a thick laser sheet (see Figure 3). The
working fluid is seeded with microscopic tracer
particles. As the model moves through the field of
view, two exposures of the particle field are recorded
on each camera. This provides a stereo view of the
particle displacements through the light sheet. Crosscorrelation techniques are performed on the image
pairs from each camera to resolve the two 2-D fields.
These results are combined using a 3-D gradient
calibration to resolve the 3-D velocity field. Reviews
of SPIV techniques may be found in the literature.
(Ref. 6 & 7)

dx1

dx2
Laser Sheet

d0
Lens

Lens

d1

reconfigured manlift equipped with pressurized


sealed hydraulic lines and a modified upper deck
with rotating table. The laser sheet optics were
attached to a cantilevered arm approximately 2
meters upstream of the cameras to produce a 15mm
thick light sheet oriented perpendicular to the tow
direction. The local fluid volume was seeded with
micron sized particles (40 m silver coated glass
spheres) for imaging the fluid motion. Three buoyant
seeding manifolds were anchored to the basin floor
upstream of the laser sheet. The manifolds were
raised to the test depth where seeding material was
pumped into the flow. An optical trigger was
installed adjacent to the test set up. When the
cylinder was towed through the fixed laser sheet, the
SPIV system was triggered to pulse the laser while
the underwater stereo camera system collected the
resulting images at known distances along the array
based on image acquisition rate and carriage speed.
The computer system was programmed to collect
images over the entire length of the array. The
camera system collects images at a rate of 12 frames
per second, or 6 image pairs per second. Therefore,
the spacing between frames varies with speed. The
field of view is 500 mm x 500 mm.

Image Plane

Figure 3: Stereo Particle Image Velocimetry camera


and image plane relationship.
TEST SET UP
The experiment was conducted in the High Speed
Basin at NSWC. The basin is approximately 904 m
in length, 6.4 m in width, and 4.9 m in depth. The
towing carriage is capable of speeds up to 25.7 m/s
(50 knots) and is equipped with a 3 m long towing
strut that allows array depth adjustments. The
towpoint is located at the bottom of the strut and is
enclosed by a fiberglass fairing. A load cell
mounted inside the fairing provides model drag load
measurements.
Cylinder models (.89mm and 2.5mm) up to
150 meters in length were tested. Each cylinder was
attached to a 15-m long, 2.5-mm diameter neutrally
buoyant leader line that in turn was connected to the
load cell. The leader line, or leader cable, separates
the cylinder from the wake turbulence generated by
the strut. The load cell has been demonstrated to
resolve drag forces to within +/- 0.1%. The tow point
was approximately at the centerline of the tow tank;
therefore, the cylindrical boundary layer could
develop unobstructed.
The submersible cameras and laser sheet
optics were configured on a submersible hydraulic
table (See Figure 4a and b). The table is a

Figure 4a. Photo of test set up shows hydraulic lift


with cameras and light sheet optics.
Test
Cylinder

Tow Strut

High Resolution
Cameras

Tow Point
Fairing
Seeding Particle
Manifolds

Laser
Probe

Hydraulic
Lift

Figure 4b. Schematic of test set up showing all test


hardware including towstrut with model and seeding
manifolds.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DRAG RESULTS
Total drag measurements were used to calculate the
tangential drag coefficient and momentum thickness
from equations (1) and (4). For the 0.89 mm
diameter cylinder, data was obtained by towing the
cylinder at four speeds and decreasing the length
from 145 to 25 m in 7.5 m decrements. This data
was combined with previously reported results
(Cipolla, 2003b) to determine the variation of as a
function of L, where L=nx, 1 n 19 are discrete
values . The results show that for a fixed diameter
array and tow speed, the momentum thickness at the
end of the array increases with increasing length and
that the momentum thickness in an axisymmetric
boundary layer develops more slowly than that of a
flat plate (see Figure 5). In particular, the values of
are 20-50% of the values predicted for a flat plate
turbulent boundary layer (Schlichting, 1979). A
thinner momentum thickness is consistent with a
higher wall shear stress, based on flat plate results.
Figure 6 shows values of the wall shear
stress w(xn) at discrete streamwise locations
xn=x/2+nx, n=0,1,2. for the 0.89 mm diameter
cylinders at four tow speeds. These values were
calculated from the difference in the measured drag
for the cylinders of decreasing length and combined
with previously reported results (Cipolla, 2003b).
The spatially averaged values of w have been
associated with the center point of each length x,
which assumes a linear change in w over each x.

Momentum Thickness calculated from Drag Coefficient


120

3.1 m/s
5.2 m/s
9.3 m/s
14.4 m/s

Momentum Thickness (mm)

100

flat plate at 3.1 m/s


flat plate at 5.2 m/s
flat plate at 9.3 m/s
flat plate at 14.4 m/s

80

60

40

Flat plate estimates from


Schlichting eq. 21.9

20

0
0

20

40

60

80

100

120

140

160

180

Total Length (m)

Figure 5: Momentum thickness for the axial cylinder


grows considerably more slowly than that for a flat
plate. Momentum thickness is determined from the
control volume analysis relating to average drag
coefficient.
Incremental Values of
Wall Shear Stress
0.89mm diameter
600

500
2

The hydraulic table was lowered into position such


that the field of view was in the center of the basin.
The strut was positioned with the array in the center
of the field of view. As the carriage was backing up,
the seeding manifolds were raised into position to
seed the test volume. The region was manually
stirred to completely mix the particles with the fluid.
Drag measurement tare runs were conducted with the
leader line attached to the strut. Tare runs for the
velocity measurements were conducted with no array
or leader line attached to the towpoint to evaluate the
wake flow of the strut.
Drag and velocity
information was simultaneously collected during
each data run. For each time series of load cell data,
the ensemble-averaged force represents the spatial
average of the mean wall shear stress over the surface
of the cylinder. In a separate series of runs, the array
was incrementally reduced in length to evaluate the
variation of drag and wall shear stress as a function
of length.

The resultant data shows that the wall shear is on the


order of 3 times larger for the cylindrical boundary
layer than the equivalent flat plate conditions. In
addition, the data reveals fluctuations in the average
wall shear along the length of the array. The drag
measurement data was consistent with previous
findings that indicate an increase in the coefficient of
friction with an increase in transverse curvature.

Wall Shear Stress w (N/m )

TEST PROCEDURES

6 kts
10 kts
18 kts
28 kts

400

300

200

100

0
0

20

40

60

80

100

120

140

160

Average Length (m)

Figure 6: Wall Shear stress as a function of length is


calculated from the average drag over the surface of
the axial cylinder.
VELOCITY AND TURBULENT FLOW FIELD
ANALYSIS
The analysis for the SPIV data required a multi-step
process. To extract the boundary layer information
as well as the turbulent character of the boundary

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

flow region, the tare data was analyzed first to


characterize the flow due to the tow strut when no
array was present. The array flow fields were then
averaged together and evaluated by subtracting out
the average tare information. Finally, the turbulence
levels were evaluated for the flow with the array and
compared to the turbulence levels of the flow for the
case when no array was present. Based on results
shown in the previous section (Cipolla et. al. 2001,
2002), the boundary flow is developing gradually.
Considering this result, it was decided to average
together all planes of data occurring within 10 meter
increments. This provides a sufficient number of
data samples (approximately 40/plane) to evaluate
average flow fields as well as unsteady quantities.
Averaging the tare data required summing
and averaging the vector images collected over each
10 meter increment downstream of the strut. This
data shows the decay of the strut wake as a function
of axial distance and provides a benchmark for the
background turbulence in the wake. On average,
mean velocities and turbulence quantities in the wake
are less than 0.03 and 0.004 of tow speed respectively
at a distance of 15 m downstream (the end of the
leader line). This location corresponds with the
forward end of the array.
Evaluating the average flow field with the
array present required image manipulation due to
array motions within the field of view. Images with
the model present first required locating the array in
the flowfield. To do this, a Regional Norm technique
was developed to identify the high speed fluid which
necessarily occurs next to the array boundary. A
search window of 8x8 vectors was analyzed by
evaluating a localized Euclidean norm to identify the
fastest moving fluid. This window was further
interrogated with a 3x3 window to more precisely
locate the center of the high speed fluid. This was
then identified as the new image center and was
overlaid, or collocated, with the image centers of the
other relevant planes of data. This method allowed
velocity vectors to be averaged based on their relative
position from the array, not relative to their positions
in the field of view.
Simultaneously, a tare field was generated
as a composite of the wake velocity maps. For each
array image, a corresponding wake field was
repositioned according to the result of the Regional
Norm, creating an effective composite wake. This
net composite wake was subtracted from the
averaged array fields resulting in the net flow field
from the array boundary layer only. Figure 7 shows
the averaged velocity field with array present for 25
knots and array size of 2.5 mm. The inplane vectors
are the cross velocity components. The color
contours correspond to the axial velocity magnitude,

red being high speed and black being zero speed.


Figures 8 a,b, and c are 3-D renditions of the
averaged flow field with array present, rotated 90
degrees to illustrate the mass flux out of the plane of
measurement.
Turbulent quantities were evaluated using
the average flow fields and each instantaneous field
respectively. For the tare data, the average field is
subtracted from each instantaneous image to evaluate
the perturbation in the flow. The turbulent field is
evaluated using the expression below:
N

I'=

(I I )
i =1

avg

(5)

For the array data, image manipulation was again


required to evaluate the turbulent velocity profiles.
For each instantaneous image the array was located
within the field of view. This location was collocated
with the center of the corresponding averaged array
flow field and the difference was evaluated. The
fluctuating quantities were evaluated over the entire
averaged field.

Figure 7: Average velocity field with array present


for a tow speed of 13 m/s show the wake flow from
the strut as well as the high speed fluid in boundary
flow of the array at approximately 140 meters from
leading edge of model.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

BOUNDARY LAYER RESULTS


The velocity fields were analyzed to extract boundary
layer thickness values, boundary layer profiles, and
turbulent boundary layer profiles. Boundary layer
thickness was defined as the average radius over
which velocities were larger than 1 percent of the tow
speed. This was accomplished by calculating the
total area over which there is high speed fluid. Each
vector corresponds to a cross sectional area, A, in the
plane of measurement of 2.1 mm by 2.1 mm.
Therefore, the area over which the velocity was
larger than 0.01 Utow was evaluated by finding the
total number of vectors and multiplying by the
corresponding unit area.
The boundary layer
thickness was calculated using:

8a.

( x) =

n* A
n =1

(6)

Figures 9a and 9b show the boundary layer thickness


as a function of axial position for a tow speed of 25
kts and array diameters of 1 mm and 2.5 mm. The
data shows that the boundary layer thickness exhibits
a relaxation in growth periodically along the length.
This was observed at all speeds.

8b.

8c.
Figure 8: Average velocity profiles for a tow speed
of 13 m/s illustrate the flow pattern resulting from
the array boundary layer flow in the downstream
array portion of the test apparatus. a. 50 meters aft of
the leaderline along the array b. 100 meters c. 140
meters.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

9a.

10a.

9b.
Figure 9: Boundary layer profiles as a function of
distance for 25 knots a. 0.9 mm array. b. 2.5 mm
array
Boundary layer profiles were determined from the
averaged flow fields by azimuthally averaging
around the array. Figures 10a and 10b show the
boundary layer profiles for the 25 knot 2mm in
planes occurring at 10 meter increments along the
array.
All velocities are scaled with pixel
displacement where 1 pixel corresponds to 4% of the
towing speed. From the profiles, it is again evident
that there is a relaxation occurring. The profile in
plane 1, figure 10a, shows there is high shear, high
momentum fluid near the boundary. There is a
reduction in planes 2 to 4. In figure 10b, planes 5
and 6 have lower momentum and lower shear in the
near boundary region, which then increases in planes
7 and 8. This corresponds well with the previous
shear and boundary layer growth data.

10b.
Figure 10: Boundary layer profiles for the 25 kt 2.5
mm case. a. 10m through 40 m. b. 50 m thorugh 80
m.
Turbulence profiles were calculated by
averaging the turbulence field azimuthally around the
array. Figure 11 shows the turbulent profiles for the
25 knot 2.5 mm case and Figures 12, 13 and 14 show
the turbulent levels in the boundary layer flow and
wake for the 7.5 knot condition. The profiles
illustrate that the fluctuations approach wake values
far from the boundary of the array. Near the array, at
distances of 50 mm and below, the fluctuations grow
significantly. Comparing the wake turbulence to the
boundary layer turbulence, the boundary layer
turbulence levels are more than 10 times the
background fluctuation values. In addition, for the
25 knot 2mm case, looking at plane 6, 7, 8, 10, and
12, the turbulence levels peak at a distance of 8 to 16
mm away from the boundary. For the 7.5 kt case,
there are peaks in planes 2,3,7,10,11,12,14 and 15 at
distances 8mm to 24 mm from the boundary. These
distances are on the order of 1000 viscous lengths
defined as:

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

l+ =

(7)

This peak occurs in planar boundary layers at


approximately 15 viscous lengths from the boundary
and corresponds to the point of turbulence
production. This suggests that the axisymmetric
turbulent boundary layer is characteristically different
from the planar boundary layer in that the location of
turbulence production occurs much further from the
boundary.

13a.
13b.
Figure 13a. Turbulence profiles for 7.5 knot, 2mm at
planes 5,6,7 and 8 (50, 60, 70 and 80 m). b.
Turbulence profiles for wake at planes 5,6,7 and 8.

14a.
14b.
Figure 14a. Turbulence profiles for 7.5 knot, 2mm at
planes 9, 10, 11 and 12 (90, 100, 110, and 120 m). b.
Turbulence profiles for wake at planes 9, 10, 11 and
12.
CONCLUSIONS

Figure 11: Turbulence profiles for 25 knott, 2mm


array at planes 5 through 12.

12a.
12b.
Figure 12.a. Turbulence profiles for 7.5 kt, 2mm at
planes 1,2,3, and 4 (10m, 20m. 30m and 40m). b.
Turbulence profiles for wake at planes 1,2,3 and 4.

The results show that the boundary layer flow is


strongly affected by transverse curvature of the
boundary. In cases where the boundary layer
thickness is much larger than the radius of curvature,
the wall shear is much greater than the case of a flat
plate boundary layer. The velocity data show that
the boundary layer remains turbulent over the entire
length of the array and that the boundary flow does
not relaminarize. Additionally, the results show a
significant increase in turbulence levels near the
surface of the array and suggest the peak in
turbulence production occurs much further from the
surface of the array than for a flat plate. Studies are
currently underway to more closely examine the near
field flow to investigate the location of turbulence
production.
The boundary layer thickness as a function
of axial distance suggest that the boundary layer
growth is bounded and undergoes periodic
relaxations. This is supported through the wall shear
data since it is not a monotonically varying term and
may be related to array motions. Turbulence in the
boundary layer is speculated to be affecting the array
motions. These motions force the array into new
fluid resulting in higher shear levels and stimulating a
relaxation of the boundary layer.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Current efforts are concentrating on the


investigation of the relationship between the
turbulence and the array motion and further
examining the turbulence production in the boundary
region. Future work will also investigate multiple
array configurations. For cases of multiple arrays, the
influence of the boundary flow on adjacent arrays
may effect array geometry due to array motions as
well as turbulence production in the array volume.

7.

8.

9.
References
1. K.M. Cipolla and W.L. Keith, Momentum
Thickness Measurements for Thick
Axisymmetric Turbulent Boundary Layers,
Journal of Fluids Engineering, vol. 125,
2003, pp. 569-575.
2.

3.

4.

K.M. Cipolla and W.L. Keith, High


Reynolds Number Thick Axisymmetric
Turbulent Boundary Layer Measurements,
Experiments in Fluids, vol. 35, no. 5, 2003,
pp. 477-485.
R.M. Lueptow, Turbulent Boundary Layer
on a Cylinder in Axial Flow, NUSC
Technical Report 8389, Naval Underwater
Systems Center, New London, CT, 29
September 1988 (UNCLASSIFIED).
H. Schlichting, Boundary-Layer Theory 7th
edition, McGraw-Hill, New York, 1979, 63.

5.

W. Keith, D. Hurdis, and B. Abraham, A


Comparison of Turbulent Boundary Layer
Wall-Pressure Spectra, Transactions of the
ASME, Vol. 114, September 1992, p. 338347.

6.

Soloff, S.M, Adrian, R.J, and Liu, Z. C.,


Distortion Compensation for Generalized
Stereoscopic Particle Image Velocimetry,
http://ltch.tam.uiuc.edu/Research/Soloff/Cali
braton/index.html

10.

11.

12.

13.

14.

Stereoscopic Digital Particle Image


Velocimetry, von Karmon Institute for
Fluid
Dynamics,
http://www.vki.ac.be/research/themes/mt/ste
reo.html
Rao, G., The Law of the Wall in a Thick
Axisymmetric Turbulent Boundary Layer,
Journal of Applied Mechanics, March 1967,
pp.237-238
Rao, G. and Keshavan, N., Axisymmetric
Turbulent Boundary Layers in Zero Pressure
Gradient Flows, Journal of Applied
Mechanics, March 1972, pp. 25-32.
Youssef, F., Kassab, S., and Al-Fahed, S.,
Low Reynolds Number Axisymmetric
Turbulent Boundary Layer on a Cylinder,
Mechanics Research Communications, Vol.
25, No. 1, 1998, pp. 33-48.
Ackroyd, J.A.D., On the Analysis of
Turbulent Boundary Layers on Slender
Cylinders, J. Fluids Eng., Vol. 104, 1982,
pp. 185-190.
Denli, N. and Landweber, L., Thick
Axisymmetric Turbulent Boundary Layer on
a Circular Cylinder, Journal Hydronautics,
Vol 13, No. 3, Article No. 79-4108, pp. 92104.
Lueptow, R., Leehey, P., and Stellinger, T.,
The Thick , Turbulent Boundary layer on a
Cylinder:
Mean and Fluctuating
Velocities, Physics of Fluids 28 (12),
December 1985, pp. 3495-3505.
Neves, J.C., Moin, P., and Moser, R.D.,
Effects of Convex Transverse Curvature on
Wall-Bounded Turbulence. Part 1. The
velocity and Vorticity, J. Fluid Mech., Vol.
272, 1994, pp. 349-381.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Neil Bose
Memorial University, NL, Canada
You found that the filament moved around
in the lateral plane which might be considered a form
of galloping or partial galloping. How did this
movement affect the boundary layer growth and was
this the cause of the periodic reduction/gain in
thickness of the boundary layer?
AUTHORS REPLY
The movement of the array caused external
fluid to be conveyed close to the boundary of the
array.
This effectively changes the upstream
conditions for the local boundary layer development.
This incoming flow introduces low momentum fluid
closer to the array wall which effectively thinned the
local boundary layer. With this understanding, the
movement of the filament was a direct contributor to
the periodic reduction and growth in boundary layer
thickness results.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics


St. Johns, Newfoundland and Labrador, CANADA, 8-13 August 2004

Internal Wave Generation by a Horizontally Moving Sphere


at Low Froude Number
James W. Rottman1 , Dave Broutman2 ,
Geoff Spedding3 and Patrice Meunier3
(1 Science Applications International Corporation, USA,
2
Computational Physics, Inc., USA,
3
University of Southern California, USA)
ABSTRACT
A linear theory is derived to describe the internal wavefield generated by a horizontally moving source in a stratified fluid. The theory, a modification of the methods described in Broutman, Rottman & Eckermann (2002) and
Broutman, Rottman & Eckermann (2003), involves a ray
solution in wavenumber coordinates, which is mapped
into a spatial solution by inverse Fourier transform. This
is a more practical approach than calculating the ray solution directly in the spatial domain, and it is general
enough to treat background flows with depth dependent
shear and stratification. The theory allows the source to
oscillate vertically while translating horizontally, a combination that has been used to model unsteady wave generation by a turbulent wake. Here comparisons are made
of the theory with tank experiments for a towed sphere in
a uniformly stratified background.

INTRODUCTION
The prediction of the internal wavefield produced by a
submarine as it traverses horizontally through a vertically stratified ocean has been an objective of naval hydrodynamics since the 1960s. These waves are produced by the vertical displacement of the fluid as it flows
over the submarine body and by the disturbance of the
fluid from motions in the submarines wake. The wake
motions consist of turbulent eddies and the bulk motion due to the collapse of the partially mixed wake region towards its equilibrium density level. In littoral regions of the ocean, where stratification is strong and submarines travel slowly, waves generated by the submarine
body predominate. In the open ocean, where stratification is generally weaker, the waves generated by waketurbulence predominate.
We use ray theory and laboratory experiments to
gain some insight into the generation and propagation
of internal waves produced by submarines travelling in
the littoral regions of the ocean. We focus on the ide-

alized case of a sphere moving horizontally at constant


speed through a vertically stratified fluid at moderate
Froude number. The Froude number is defined as F r =
U/(N a), where U is the speed of the sphere, a is the radius of the sphere, and N is the buoyancy frequency of
the fluid.
The present theory is a modification of the methods
described in Broutman et al. (2002) and Broutman et al.
(2003). It involves a ray solution in a wavenumber domain, which is mapped into a spatial solution by inverse
Fourier transform. As explained in the above references,
this is a more practical approach than calculating the ray
solution directly in the spatial domain, and it is general
enough to treat background flows with depth dependent
shear and stratification.
In the theory, we allow the source to oscillate vertically as well as moving horizontally. We plan eventually
to use a distribution of such sources so that the horizontal movement of one source models the body-generated
waves while the vertical oscillation of the other sources
models the unsteady wave generation by the turbulent
wake. The theory is also derived for a depth-dependent
background, but here we only show results for a uniform
background. This is to compare with tank experiments
that were conducted by towing a sphere through a background at rest with uniform stratification. Wave reflections from the upper and lower boundaries of the tank
make an important contribution to the wavefield, so these
reflected waves are included in the theory.

THEORY
We consider a stratified fluid in the Boussinesq approximation. Internal waves are generated by a source moving
horizontally through the fluid while oscillating vertically
at frequency . The coordinate system is r = (x, y, z),
with z positive upwards, and is fixed to the mean position of the oscillating source. In this reference frame,
the background flow is U = (U (z), V (z), 0). The background buoyancy frequency is N (z).

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

The internal waves have wavenumber k = (k, l, m)


and intrinsic frequency

Eq. (4) can then be rewritten as


r (k, l, z) = C |G0 /G|1/2 ei

= kU lV.

C = 2i(
mQ/Bm )0 ,

B=
2 m2 (N 2
2 )(k 2 + l2 ),

Q(k, t) =

(4)

The
R wave phase has the usual WKB form of an integral
mdz (e.g. Bender & Orszag (1978)). For a depthdependent background the wave phase generally needs
to determined by numerical integration. The wave amplitude |b| can be determined from the conservation of
wave-action. Here this reduces to the constancy of the
vertical flux of wave-action cg3 A, where A(k, l, z) is the
wave-action density and cg3 =
/m is the vertical
group velocity. The wave-action density A is related to
the wave-energy density E by (e.g. Lighthill (1978))
A = E/
,

E=

1
0 N 2 |b|2 ,
2

Summarizing, we have the ray solution (8), with


C given by (9), B given by (10) and Q given by (11).
Substituting these values into (8) and taking the inverse
Fourier transform (3) gives the spatial solution. Modifications for buoyancy frequency turning points are given
in Broutman et al. (2003) and Broutman & Rottman
(2004).
To compute the spatial solution for other variables,
such as u, the x component of the perturbation velocity,
we need only to relate the vertical displacement r to
the corresponding ur by standard WKB relations, and
then take the inverse Fourier transform (3) of ur . These
relations can be found in Gill (1982).

(5)

where 0 is a reference mass density.


In terms of the wave amplitude |b|, the constancy of
the wave-action flux cg3 A is
G|b|2 = constant,

R EFLECTED WAVES
(6)
So far we have only considered waves that move upward from the source. To compare with our experimental
results, we must allow wave reflections from the upper
and lower boundaries of the tank. From here on, we

where
G N 2 cg3 /
.

 j1 (Ka)
3 2 3 
i a U0 k + heit m
, (11)
4
Ka

where K = |k|, and j1 (z) = (sinz)/z 2 (cosz)/z is the


spherical Bessel function of order unity. The background
flow at the mean position of the source is U0 , which is
aligned in the x-direction. In the limit of large Froude
number F r = U0 /N a the source function (11) represents the effects of a vertically oscillating solid sphere
of radius a. The vertical displacement amplitude of the
oscillation of the sphere is h, and its vertical velocity is
heit .

We write the ray approximation to (k, l, z) as


m(k,l,z 0 )dz 0

(10)

The source is represented by Q. We will use the


following form, from Dupont & Voisin (1996),

The disadvantage of this method of solution is that


the vertical eigenfunctions are not generally known analytically and are not easily obtained by numerical methods, especially when the wavefield has critical layers and
turning points. The method is simplified by approximating the vertical eigenfunctions with ray theory.

(9)

with B = 0 defining the internal wave dispersion relation


and Bm indicating partial differentiation with respect to
m.

The depth is treated parametrically and is not transformed. The factor eit accounts for all of the timedependence in the present model. This can be considered as the long-time limit of an initial value problem in
which the motion is started from rest.

Rz

(8)

where the zero subscript indicates evaluation at z = 0,


the mean depth of the source. The function B is

(k, l, z) ei(kx+ly) dk dl . (3)

r (k, l, z) = b(k, l, z) ei

The factor C = C(k, l) in (8) depends on the spectrum of the source and on the form of the dispersion relation. Broutman & Rottman (2004) derived the following
form for C

(2)

We introduce the vertical eigenfunction (k, l, z) for


the vertical displacement of the fluid. The spatial solution for the wavefield is obtained by inverse Fourier
transform:
= eit

m(k,l,z 0 )dz 0

The factor G0 (k, l) = G(k, l, z = 0) is included for later


convenience.

m = (k 2 + l2 )1/2 (N 2 /
2 1)1/2 .

(r, t)

(1)

The internal-wave dispersion relation is

Z
Z

Rz

(7)
2

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

will restrict attention to a uniform background, which is


the condition of the experiments.
To account for wave reflections, we follow the
method of Broutman et al. (2003), which we briefly summarize here. Consider a fixed height z above a source in
a tank of height H. Each time a ray returns to that z after
reflecting once from the top of the tank and once from the
bottom of the tank, the wave phase mz has changed by
an amount m2H. To include the effects of this and subsequent reflections, we must multiply r (k, l, z) in (8) by
the sum
S

ein2mH

n=0
imD

= ie

/2 sin mH.

(12)

Figure 1: A sphere of diameter D is towed horizontally at


speed U through a linear density gradient in water of depth H.
Wave motions propagate away from the source and through a
carpet of neutrally-buoyant polystyrene beads with density 0 .
In the experiments reported here, the beads are 2 sphere diameters above the midplane of the sphere. Model vibrations are
reduced by operating at high tension T in the oblique support
wires.

(13)

The infinite sum diverges, since the individual terms do


not vanish as n , but the sum can be evaluated in the
sense of generalized functions (see Eq. (1.2.2) of Hardy
(1949)). Still, the sum diverges when mH = , the condition for perfect constructive interference between all of
the reflected waves. This divergence can be eliminated
by adding a small damping factor in the form of an imaginary wavenumber, e.g. mi for the vertical wavenumber,
or by limiting S to a finite number of terms. The number of terms can be chosen to represent the number of
reflections at a given time of interest, for given k, l, as
determined by a group velocity calculation. We have experimented with both methods but have used only the
first method for the results presented below, with a value
of mi D = 0.02.

To improve the spatial and temporal resolution, the


image window was centered on one side of the wake
only (assuming bilateral symmetry), and exposures were
made over fast and slow timescales using a nested pair
of short and long exposure times at each timestep. The
components of the horizontal velocity field u, v were calculated directly from the particle image displacements
and interpolated onto a regular grid by a two-dimensional
smoothing spline.

EXPERIMENTS

RESULTS

Experiments were conducted in a 2.4 2.4 m tank filled


to a height H = 26cm with a linearly stratified salt solution. The density gradient was created by the standard
two-tank method, and density values were checked at an
array of taps in the side-wall. In the experiments reported here, the buoyancy frequency N = 1.88 2%.
The sphere was towed at a height of 11.3cm above the
bottom tank. The sphere diameter is D = 2.52cm and
F r = N/U a (where a = D/2) was varied by varying the tow speed U . For F r = 1, the tow speed
was U = 2.37cm/s. The Reynolds number Re varies
by the same amount (i.e. for F r = 1, 2, 4, we have
Re = 600, 1200, 2400), but we assume the variations in
Re to be of minor importance for the wavefield. A horizontal (isopycnal) bead sheet is left at a height z = 2D
above the mid-plane of the sphere, and particle motions
are tracked using a custom DPIV technique, detailed in
Fincham & Spedding (1997) and Spedding, Browand &
Fincham (1996). The experimental setup is sketched in
Figure 1.

We consider Froude numbers F r = 1, 2, 4. The sphere


is towed in the negative x direction, and each plot shown
below is a horizontal cross section at a height of two
sphere diameters (z = 2D) above the centerplane of the
sphere. The plots show the x component u of the wavefield velocity, except for Figures 6 and 8, which show the
horizontal divergence ux + vy .
For the theoretical calculation, the inverse Fourier
transform (3) was approximated discretely on a
wavenumber grid of 512 by 512 in k, l, for F r = 1, 2
and 1024 by 512 for F r = 4. The grid spacing in the
spatial domain is in the range of about about 0.6cm. A
smaller number of wavenumbers suffices in some cases.
In addition to resolving the flow features, the extent of
the wavenumber grid must be chosen to limit periodic
wrap-around errors, which result from the discrete approximation of the inverse Fourier transform. For a single depth, the theoretical solution takes only a few seconds to calculate on a standard PC.
3

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Figure 2: Experimental results for u in the case of F r = 2.


Each image is a horizontal cross section of one side of the wake.
The 12 images are for equally spaced times over the duration
of time N t = 29.8. Time increases from left to right and from
top to bottom. The total distance traversed by the sphere in this
time is 29.8D. The minimum and maximum values for u are
-0.2cm/s to 0.14cm/s.

A sequence of images from the tank experiment is


shown in Figure 2 for the case of F r = 2. Here and
in the following Figures, the image grid size is 74 x 54,
with spacing 1.072 cm x 1.056 cm. This gives a 78.3 x
56 cm window (inside a 2.44 x 2.44 m tank). In units of
D the window is 31D x 22.2D. The window is fixed in
space and the sphere moves through it in time.
As noted above, the images from the experiments
were taken on one side of the wake only, assuming
a symmetric wake. In the following figures, we have
flipped the experimental images symmetrically to give
the more familiar picture of the full wavefield. The data
points directly behind the obstacle, along the line y = 0,
are at the edge of the field of measurement, and this
shows up as an anomaly in the following figures.
Figure 3 show results for the case of F r = 1. The
tank measurement is shown in the top panel, and our linearized theory is shown in the lower panel. At this relatively low Froude number the turbulent eddies in the
wake are not strong and are thus not important for the
generation of internal waves, an effect omitted from our
present theoretical calculations. However, the source
function we have used, (11), is most accurate for the
representation of a solid sphere only in the limit of high
Froude number. Thus some of the discrepancy between
the theory and measurements in this example may be due
to the source representation. In the future, we plan to try
other source representations that depend on the Froude
number.

Figure 3: A comparison for F r = 1 of tank experiments (upper panel) and theory (lower panel). The plotted variable is u,
the x-component of the velocity of the wavefield, in units of
cm/s.

Our calculations indicate that wave reflections from


the top and bottom of the tank are an important contribution to the wavefield. We show this by removing the
4
Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Figure 4: As in the lower panel of Figure 3 but calculated without the contribution of the waves reflected from the upper and
lower boundaries of the tank.

reflected waves from the theory. The result is shown in


Figure 4. Comparing this result with that of Figure 3, we
can identify the reflected waves as those having a more
transverse orientation to the towing direction.
Figure 5 shows the case of F r = 2, again with the
experimental results in the upper panel and the theory in
the lower panel. The peak theoretical amplitude of the
wavefield is down by a factor of about two compared
with the peak amplitude of the experiments. This difference in wave amplitudes is maintained in the region
closest to the obstacle, so there must be factors other than
turbulent eddy generation in the wake. One possibility is
evanescent modes, which we neglect. These could be
included in our theory by the usual WKB methods, and
they might account for some of the difference in the amplitude predictions.
Another possibility is that the experimental results
include vortical modes, which are also neglected by the
theory. To check for this, we have computed the horizontal divergence ux + vy in order to reveal the internal
wave motions without contamination from the vortical
modes. For the experimental results, the differentiation
of the velocity field was done using a spline fit to the
data. For the theory, the transformation to horizontal
divergence was done in the Fourier transform domain,
using the WKB relations between the vertical displacement r and the horizontal divergence. The horizontal
divergence for F r = 2 is shown in Figure 6. Here the
peak amplitudes are in much closer agreement than for
the corresponding results for u in Figure 5.

Figure 5: A comparison for F r = 2 of the experimental result


(upper panel) and the theory (lower panel). The plotted variable
is u.

Figure 7 shows the final case, for F r = 4. Here


our linear theory is a poor representation of the solution,
though it does give a reasonable prediction for the lat5
Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Figure 7: A comparison for F r = 4 of the experimental result


(upper panel) and the theory (lower panel). The plotted variable
is u.

Figure 6: As in Figure 5 but for the horizontal divergence, in


units of s1 . The experimental result is in the upper panel, and
the theory is in the lower panel.

6
Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Figure 9: The theoretical result for F r = 4, as in Figure 7 but


including an oscillating source term to model waves generated
by the eddies in the wake of the sphere.

eral extent of the wavefield. This is the Froude number


regime in which eddy-generated internal waves are important, as appears to be the case from the experimental
results in the upper panel. There is an order of magnitude difference in peak amplitudes between experiment
and theory. The difference is reduced to about a factor of
two for horizontal divergence, as shown in Figure 8. But
clearly the pattern of waves in the theoretical solution is
still very different from the experiments.
We mentioned previously that an oscillating source
has been used to model internal wave generation by turbulent eddies in the wake of an obstacle (Dupont &
Voisin (1996)) based on some experimental work by
Gilreath & Brandt (1985). We added an oscillating
source to our calculation for F r = 4 and obtained a
closer match between theory and experiment, as shown
in Figure 9. Following the guidelines of Dupont & Voisin
(1996), we chose a value of 0.2 for the Strouhal number
St = a/piU and a source frequency of = N StF r,
or about 2.51N . The amplitude of the vertical oscillation of the source, represented by the factor h in (11), is
chosen to be h = 0.8a.

Figure 8: As in Figure 7 but for the horizontal divergence, in


units of s1 . The experimental result is in the upper panel, and
the theory is in the lower panel.

CONCLUSIONS
This work is a step in seeing how well ray theory can
simulate the internal wavefield generated by stratified
flow past an obstacle. In Broutman & Rottman (2004),
we used the present method to study the refraction of
such a wavefield by a depth-dependent current without
boundaries. Here we considered a uniform background
with reflecting boundaries, for comparison with tank experiments. Our view based on these studies is that the
present method can effectively describe the diffraction
7

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

of the wavefield, but it needs a better representation of


the source.

translating and oscillating sphere, Dyn. Atmos. Oceans,


Vol. 23, 1996, pp. 289298.
Fincham, A. & Spedding, G., Low-cost high-resolution dpiv
for turbulent flows, Exps. Fluids, Vol. 23, 1997, pp.
449462.

We have used the analytic source function (11),


which, for flow past a sphere, is justified only in the limit
of large Froude number. However, at large Froude number, eddies appear in the wake of the obstacle, and it is the
eddies, rather than obstacle directly, that generate most
of the internal waves, as described in Gilreath & Brandt
(1985). This suggests that our analytic source function is
most appropriate for moderate Froude numbers, and this
is indeed where we found the best agreement with the
tank experiments, i.e. for F r = 1, 2 but not for F r = 4.
We achieved better agreement for F r = 4 by adding an
oscillating source term to mimic the wave generation by
eddies, as suggested by Dupont & Voisin (1996).

Gill, A., Atmosphere-Ocean Dynamics, Academic Press, 1982.


Gilreath, H. & Brandt, A., Experiments on the generation of
internal waves in a stratified fluid, AIAA J., Vol. 23,
1985, pp. 693700.
Hardy, G., Divergent Series, Clarendon, 1949.
Lighthill, J., Waves in Fluids, Cambridge University Press,
1978.
Spedding, G., Browand, F., & Fincham, A., Turbulence,
similarity scaling and vortex geometry in the wake of a
towed sphere in a stably stratified fluid, J. Fluid Mech.,
Vol. 314, 1996, pp. 53103.

We are planning to follow this up with better representations of the source, dependent on the Froude number and including a distribution of oscillating sources to
model the wave generation by eddies in the wake. A
useful approach might be to match the ray solution to
a numerical simulation of the near-field around the obstacle. The numerical model would in effect replace the
analytic source function used here. The matching would
take place in a Fourier-transform domain, and thus could
probably be done at short distances from the obstacle.
The results of this approach may suggest better ways to
parameterize the source function.

Acknowledgements
This research is supported by ONR under contract numbers N00014-01-C-0191 and N00014-96-1-0001. Dr.
Ronald Joslin is the program manager.

References
Bender, C. & Orszag, S., Advanced Mathematical Methods for
Scientists and Engineers, McGraw-Hill, 1978.
Broutman, D., Rottman, J., & Eckermann, S., Maslovs
method for stationary hydrostatic mountain waves, Q. J.
Roy. Met. Soc., Vol. 128, 2002, pp. 11591171.
Broutman, D., Rottman, J., & Eckermann, S., A simplified
fourier method nonhydrostatic mountain waves, J.
Atmos. Sci., Vol. 60, 2003, pp. 26862696.
Broutman, D. & Rottman, J., A simplified fourier method for
computing the internal wavefield generated by an
oscillating source in a horizontally moving,
depth-dependent background, Phys. Fluids, submitted.
Dupont, P. & Voisin, B., Internal waves generated by a

8
Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

WBB%g*!Vp!BfLLB'
Ba> wts>LT!!ZvV'g

   !
"#%$&' $)(*,+ -  ./0 %213 $
4658795;:<>=@?BADCEGF@H65;IJ<>KB?L?BAM=NEGF;O5;PQ@K%<>RTS@EVU WXFZY[5;\T]T<>=@=^RN<>K%_`WaFZb[5@c dfeg=TW
h E Oji@kadglm<`K%?Y.=en]M<>KB?LenopSqAarsHt<>uvi=AMdgAa_XSwF
W \wCj<>xef?Li yz<{r|<`=?B<}49<>?B<`kXKBuvic_a<>=@uGSTFZPJ~p
,>Z,J

gjL^fpf|0' LpZfTtnVL,gg>ff
BMpjpvjfp,|Lp
fTp @6E|WpE|;wf%{wp^p
fgffL|TfL'LLp
fZ|fwapfvf^p w%%
LLLBpDa%gf!! 6LL
pLV!D`f%@p-!`
BLL>fag{`wTD`MfX`gg
'f|w^L'pf,BfNLVfp
LJ|Xf;pp;f>NgTL
f%fw|'@VfNLfp
gMM@pgf^fff%fTvpf
fVTpff|Lv^
f>f!VM@p|gB
afVfLfLL>!gtfVf
fJ|G|pwJftZ,D`,p
ffVT!`|Lf%p^pXVp
|ffwZfLtf|NLL
f'p^B8VfB|%gnB
LLL^'fJwpLfp|gpn
ffL{qsvp[ LLsgV
Vf-wp@LLNsfq- V
f'!`f L!'V[BL
pf%w^sfNLf%VBfZ|'
Lp %p%tf^fvff
npf@f@fMg
f>fgafTLG|`v|nT;fanM
fL;'LLapTfp>f
%;ZZBL{!VM`MJfV;f
LtLL^|pwfV%^VTf
%%pf,L%%|fn'LL

ZJs
wzp-BfpLfq
vJfffssVLX
n}pf%Vpnp;#
|fpnpBpn^p!^f0fp
Vpgfp%>f@M%@|%Dw
Zs|pff8BLfN L
BgLDwBBfLLB'TfpLf
fnffB%pp|nMw
pfptVLpp|BLLp
%p>NL{f!L`LgNgfL`|
LNf;f%f,!vgfL
V ;gvp%%
p|fB
B!w%;vV
|Gn%p 'fE%^
ps'fEB,LLv Jfp
vpp>fL>| .pf@T
ffBff%Tftf^>f
gf|f>-,ff|Bp'wg%f|
^f-pfZp^f-ffLf
MfXf'f*fJ|JL
gfLVN^gjBpffwV
vpf|LLw^pBfvf
p%fpNgf%g9tf
fLnsf^|%f@n^fff
v;pfgDf,V
fTZvfp%!|Jfp
wLpTpg'npf{%^f!f@LL
fnfVp'pfBLvf|f
vVppnv|nTL|f%!
v|nTfftfL^TpfL^sL
awfVgf|f`fTwp>w

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

VfL @gf>pp%|
)Bgf !-gp
gg M` M|f,f
V6nf' Bvp9!`Gp
vLsfpN'LLMwZ nw`
gf L0p-fffpjpL6f
tf%w^f;|',V|pfB
.fJgXgpsLL8wppf
gfLfwvpN`V9fwL|pD-
fB!|BfLsf!gfLs^f' 
LD`!|Lpf'^fnpGf
p'Ba^fjpvpw,fVf
pfNJvffN!LVp
fLB8fpgpnLL'B
wL g|
wfpNtV
-gpfB;L|! ^ 0Bvpt'
ft s|Lv%p`f|ffG|XLfff
 Z%D'g8fNLLV
TN%sfBp>fL
LLfV;Zvjgjpj|vft|'
LpfMf|gJf;fJfg-
MfVTN!`TLVf>aVMZ!`
LVVpZpf^f8v;>
,|VvMNf^f-g'||g
pLBG
t M
@
-!|fLa ^ BvpD`
>s|Lpfa!pa%fp;|
fnjpfB8%8ptn
v^f}69BtnL
>%TL|f-,D`|L
VBpBn%Vfvpf Z VMJg@p,
fffJf G fBL
'%wwsfg,f;p
f%VMLJwf9fp'p
NntppTLV@fLp
vff;ftVw8{XjBjff
f-|Vg,|V.f
|ff9f[ >0gfL0
fV^L|pMNfNZpg'fZpB
LV D`G,|'f'L
pfMp ,vp'TVf

f@@ftpp|ff|L8 @L
pg,V6gff
Tpff%ppZV
pp L||,p!L'
V6@ B|wf%
f%9fptLV^ff-Lfpf|
>V^p'^'p-.
f-psfv gs-VsV
LLV|8fp wfpNML
pfTD.Bps8T f;V;L
gff; Vfj B>p
!BfBf !Bw`pa%
|{a%{Bv'%pf-LLXp
|Jp}gfLgV9V|LLVw
J|zfpgfT^tf
gfLsv@pZgp%pp@tf
VN'ftLVp|TqBf|
%f{
TpLVf%9pV%L}B
>f L.B;pf @fBV%
Tfj% gj% pL
p| L-BVf. Lppf
%@^p%'LVfLLv|nT
@qp !@mV|LznV
gff{@f^|f|gpwDfVp
LL,-fpn,VV}|
pNpLV-gfLfffLBp'
pf|.ff|Bgf
| VB6|p
 pjgpg}!gwf
|Lpf|L n>gfpL{pLpp
ff G%w^fTffL
ft; 6pV0,fpfvf,p
VBjpJnLVjV
ffMgff>pZvMLVfp@fg{
p6;fffpJVVX^zL
%f|f^f6f-p
%ZBgVffL
'Tf|!'LLwfg'f
ftv'BLf;%V
LgTV`fLBpg
%|gp6fpstf
wp,papg-fLp
fDf^fVBfgL ZZfw g8f
LL8!gG @vv
fs.%.pVX fBT



 







Copyright National Academy of Sciences. All rights reserved.



Twenty-Fifth Symposium on Naval Hydrodynamics

w@nMNMp%Tgff>``M@ggf
nff!wB8Mf|ffN
LVX{fVJff{p,L
!L>p^LfZ|%XL%
JvpZf'gfLvff
|f|fgfVftvpMfp
f'f%!f@ f^f
pg-pfNf,|LLff|N,Lpf
>gftNvp
. N  N
wTBp%wNpvBf ;%p
9pVffjpf
sft!Bgn%@%VDf sLL
8p|DwMff'vVftptLff
B,BL,^f'f '
^U
wU
n
; U  U U
^ ff JL
N f^B w-ff
Bgn%faVfTptfsN^ff

8  U 
U
^ 9'Bn f9
ff W fBfNgfp
f
MfpfNgfp
fLp tTfNBV@Bffn
wzMpfN LL
f^Bgn%fN%pf'p
,fLgfff

N|'Lpffa vU 
GU
G^ vU
TfZ;Vtpf
wff vV
-B-ftBpaJB
fBVf!f
^.|G| pfg8fqp
 fvZ%f ^fBDv^LL{
%Lfpp}VpfpJf%fg
V@ !BVpp%LJpNpp


Dzw[pv  Bvp
!`LLX!ff^ps|

.

DC

JALKGM

ON

@ JM PC DQSR<M

@ "T

Y\

=K

]4Z0[

cb

fe
g
3

63 d
YhZU[ Y:i

Y \jY

_

"b

.

43

65 +7

pkq>nF
[ @
K

Yr\:YZs]ut

Kv^

_

8/ 1

Za

^K`_

&% (' &%

0/21

@ BA

E>GF

VXW YZ0[

+%

-!

DC

UN

"! 

#

BA

&, &%

=

<;

HI>



*) +%

4

?>

V
w-|fL|v6p%p
@VG@fXB^f;'f%
afV^JNwvfJv!Bv
TN>LXZw|vBp
 saBLLj!pz sapp
VLLsap9;NwBD@pLfB
|f`psJpL'fBf
pv LasDwTpBDLVZp`Mf'L;L
fwpB8%%MLL@V %
pN`fp!LTVT p^s.pZLf
f%fgf|fMpgpaJ{f@f
VT%Xp %sw^agffJ
|fgN!!TtfZV  
E  BW B^ s9 wJ^f,f
pJwJgffppLpfNfL
m^fp'ZV 
Lv%ppTf%
wVpVt!vfVBL,%J@B D
 LL{wf@ ^|>%p @
EW'LE,gf'|gfpBn
p;BV%|fn Gwwf>8 Df
|fVp%X>p@ff
g@fp%nsf'Lnpt
gfV;fpvL>W 0fwf|nfpL
n wwLLL|;^;f^f
%pJf^fz|Z'^9|B
%f;pXpwpapvLXB
$#

:

_

Copyright National Academy of Sciences. All rights reserved.

K Yj\
[
lkm>nF
0km>nF

o Y
[ o @

Twenty-Fifth Symposium on Naval Hydrodynamics

faLp;fLfNVDf
gg8pBfTfLnJvtf
LgpzT%;v m -tf
BgTfNV
J W E W E

W
U
^jB E
 '
fffV^ 'fVG BBf
fzgf6f-M

^ {wfff;B
fffJ wM| %
ff%st0LLt^
f'zf9-Lff%.BBL
fB^%
X^-
fZJfpfvf!%pf V
pTL|LTvf%g{p T @f
w^ @%fw p s|
^!gfJ!f'pnpJZnf
%pDg!f@!|'LfafXLL%%
%DBanL;XfV^fpXf!p
[g.%p-LLNsLgp!


Lpv TBv%p| a%gf


L f!fB'fVBL ; ^U `afB'g
tf%Bp|p9fp f 0{

;L ;gDv^fpfnN
fvNvfZL|,Bv  fp
pBfB;Bpw|'Lpf6f
ff  |XaVMf;p
Lvf%fVfLN|'Lffw
f%n%Dw|G|apvfg^Xp^p
^tLgfBLsptfgnMgB|L
LL`|Lg^fTNLf;fZ!fpf
!fpg'Zvp%fLff
p,|Lq'wp{'LL
w

z}| | _~||

| | R

(_

$>

JA

xN

D_z

y_{z

?_

?b

mw}LL,f[g8fmnff
Nf XZffTffBBfL
f!gp@w! B
fLL|p'L;Bffn fg8f
B|fGDpLL,LBpf;g
p 

U 
wW
J
wfpXLLpZ|fsf
p}pf pg'fV
f|Gtp!ff.'zff
pMJ@LLBp,f'pp
BGp|T`,|ff;LL8TZs|B
gJ 9p6^[p6Lf
fgpV

 !t|ffnw pLvLB M@ af@ f  %%pfL


B a^sLLfg%pf.
t^f{
%
n 8n XjW
8f p
pJN,B|%gn|'Lft
N U

^f T qMf@Lfpf, wp
p,LLw'f X 
e
g

tI

n+X+

Z0[

Z0[

3<P

g2

D3

e

vO N

.

$O

5 &,

|| d||

I8


<

Y\

Y\

Y(

65 2

rO

(K~

L**6O

g Z

3<P

D
e R R

jX

r.

3<P || d||

=.

X

V W

"_

f'L^tafL%B|%gns'LL
@f'% ^f wfpf
v|nTfgnM.p%fft0fV^
sVN`f J> n%Jw,L
Lfn Bw BsT n%

!fpLBpfj9f
Tff,BBvpff@pgVJv|ff
fpf'|f;^fBv'%X@
 {%BB
8
w`B%pf{[fM|>VgXf
fL,p@wB

U
%
ff-fLBN-vJ|Lf


3<x

D.

X?-+

<

63<x d

_5

gB

(_

R*

.

43SP

~_K

3 P

"_

P| | DQu| n| 

h3<P

&""X(~-&

&0Gjn"

Y g

y_

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

-gT@>wf- gfs0v
pLZ@wpf ^fTf
ZBVgfp W{
W
{  @Lp
gpfB[fzB|pffp
p'fBn `^

^fZTgpgf^w;%
vvf'%ff ^
wp|X%p|^jZff,fJ
@{w;ffp{L,ffqw
LLf'gsLLDBs'ppX
Tfp^fVBLfN|f|w'p
^f !wpDNpf%{Z-'L
gfa fMN|LDT%LLf|`^Zf
 9DV^Ln  8p|
TjL|6BptfL|n
*fNw8|wp
[EE WWU

! EE WW g
wMpfLVfaTBvp;;v
%fjB-LffwpB|%gn
ff'|pBffn fTXf
VwZgL8ft%|fn|
t
N `
U g%
^f;f|L Lpf^V;%n
vLVfpzVppfp,g
fJv%!T*fNwp
@ @ J
a,LTBp'a,vp
f|pf'LfsfZ'fpf%
VffTpfTVNNNfB!w'LB
ZL|TpfwfLZD%fpf
f`M%pBff^^Lf wfpfJ
ZLp>g 8 
^s
NgL8f,fgp' Xfs
w}|'Lpf0fg%f}p
LZ^f;;LfffVf
^pfL|V;vp
p-f|`ff'gn
^ffpf 'p.-f

w-fpZpftpZjg8f p
fVa%f|X^f!fg'fpXp'fBa
gfVs
E
E
nU
g
W f W
^f9Lp6f}p 9
LLwp^s|v%pfL9nff
fffp|^9ftfgf
p
BpjLLpfp|jp
 Mf!BVsBf BB
^8p|v^p-pfLpfBf;
Zf-|'Lf p9gfZ|Lpg
LLJ8Tpf8X%w, NL%ff
fNLLT,fNB

D_

N4

lNS

 s LB'JgLGfL
%pTfLVL`p N%8fg`^f
fffNgfXfB^VTapw
Bp@npXfX>@'LLfafg
fjfp'pB'pLgV
f> t;pf;gf
fB8 j'LL{fZ%
fv|nTf!fg'fpXf@gX8f
pNp'LLB^f
j
fLL{f;ffpg^v|nTf;
wffJ^ L Ba
fJL-LLMf-Z
fZg8f,gfTfg^wLL
E
g
W U
^f NLf X j-
|'%fp!Z''%Vp!B
wJgfp%g8wppgpnz
T|f{


Y\
K

Yl\

Y g

63 P d

4jD~U8:n8

.

e

8

r.

&U

&

v3

9l

:

N

'pGTvL 'V|f
|LvfJVZ ^JVffLNfg
ff!L%pp.vftL,

65


V / V P|

3<P

0| |

G3

tZ

$3 P

.s.

3<P

.s.

3S

.

r.

=.

p3 |

&~&

0.

-3<P

=

>

*?

(N4!+

N| |
g
x| |

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

 GM.f|J|Lfff
!pg|Lfp'gBff
^fsNwpB8M`LVpN
pf,^fs%f%|Lf^JgfL
V TBJfn
w |f LL fp
VfBnT|G|w,n;;fp;f
p!psfnfB |Lpf f
fg0Lp gv|
;|BL^wp^|tfB|fff
ff9LL'^T9.pvf
fffJLgfLf9f-,Bwtf|
f9f^;gf.|Lv%
fftfpgfff%f
V^a|Lf!|'Lpf
LL@gfqB,f'LVffpf
DfJf9^J;|Lv-Jf
vp>TpvLMwfg
{fa!`XLVfpf
LLLffN'fN|JV0pfw
^LV^JfLVfg!'LL

MLpfJv gf.
wLG%ppf`'!Z
M>MLV@^fJE E^p WE
'fXp.j!`MfV^
.L ^W ,GE^f-f
`fL'fgvf@|%pgDf
Nft!`,|LpfsT|nTf
`pX^fNfMvpapp
f|fpBXB^fZwwpLnp{g
Gp tp ,T@Np|%pg
fv
wpjgLpf ^ff
fpn{%`p{VjNMg
pf,m !Vs`BTp
.wfsp.pf B 
^
p  E|
aN g
XgE
f!%f L
p
@fM>f TVpBLp >
 qf %fp|p ,B,g B ; v! E|WtE
^   LE L W  
E nE|
 tw^fgg|%
fNsf;vp,!pfB`^
]

D^DwfwLMf|Dfffpvf
!`'LV
|B|ffsv6f %pG;
fLpw,nM|Lvg
|8TB^G !^^f
pp| 9%g,%sp p'%n
Tf
w%pfwffpftf'f|
pfJpf%-v%^f,%V
ffpf'fp|fpf'f%%
w tL|n%8ffJpjf
g !nBv!L|f`w@pf%Xf;
T{B;vXgffXgn%p;T{
B;vV ^vgf,NL|
%V^BpLnpn'p|>f'
LL|'vfffNT^f^
pp|Lf
NT@ @Z |
wj|LpfZgpLLNsv
%|B@paLp|  %
%g0z-9|nf0;fj^
f8^fsZ|!N%`p
gfpwBLpZfwLf|DnM@f
W LV^gfT|`
fpgwZLpGp!L
p.|L'pJf T'ftL
p|LfffsL|f

(

Q

5,)+,

 e

e

e

/ 1 

X5

R 

&

65 '

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

f@fL|n,L%fZTDgp
gf'.zf'wpt^fgf
|f  EZff|nfpqp
^BM  w'f  |sDw^f
pf%>`f%>gpT|
f B,fNfJfLft^Vf!
G
N>



DN ffN%f^L|nJ
w^Vf'fBfnXpGa| f@af8vLL@pMD`TLV
Df;npV,vpwfN%gN
fB8f^f'V^f8vLL
pNf;p`Tgff>ppZfTT`TDf
%gpMwT@f'DvpN
^tLaL^;!fff!L
%DwD`N|BfDV`g`p
fw|LLMvpZv
ff,wBV V@|fL
^ZgBp
w6fLfmgfm*npf D@MM%Tp %$ '& tfNB8
v|wp^f'pT|LL;{
w'fgLptfZf
LpXf@L|vL@Tfffp
LnsfVf@vpfv%
ff%gBv|f@f TV>pfDw|
LzBB|Vgpgf
p ^;fNpj V
p# JpT^Jf,LwT
fgDVafpfXBgp;%Bf'p
|f`V@B@fwfgD%ppg
%!fp
^!f;fpfLZvp,
fp9gfLNpBn|fpf D >%f'`pNf!gg|>pLgn
|Lv|f9fff`
ZpwfM%vfa%D9vp>V| pwfZgffpDfZ>N|Lpf
|aL||f'V`f^gDfwfV
;TpgNfB'fpp
fMBfV^|`!ZZDL G^gf'pX'g %( pL|
|MNvpZ|Vp{v n;pfNgg|wp
%agV^>
Lfj !jLLJ@LVp 9fp9VNX||Vjtp
fstf%fsp>NB
vp^^g%pgppv%
`VT^f;pvnp'f -p,f9wftf,fpfvff
|ffgffp{Gff fgZLnpJpBTpff;LV^>

" j

 *!4:"

(N



?| | d||xK| | [

Copyright National Academy of Sciences. All rights reserved.

||

Twenty-Fifth Symposium on Naval Hydrodynamics

1
0.8

0.2
0

0.2
0

0.5
v/v

0.4
0.2

0.5
v/v

0.6
0.4

r/r

0.4

0.6

r/r

0.4

0.6

r/r

0.6

midship

1
0.8
midship

1
0.8
midship

1
0.8
midship

1
0.8

r/r

r/r

midship

OEEVM
MM
OEEVM fine grid
MM fine grid
EXP DTMB AFF1

0.2
0

0.5
v/v

0.6
0.4
0.2
0

0.5
v/v

0.5

v/v

D L|nf-.|V9|Lf%0gffmV

|Uf  Ug W VUf p)U|p !`


fTpf'|'J|gp% fwJBfg9wf^fT%f
gLgfp%|.g'^ ;v@fpp|L6f
,fV^,B
LVpTB8a%fpftfs
J@T@p@'pnfp favspB@N`>;gf!Tvf@'^fZ^`fD|L|
^gLn%NVLVf ^LL8
gf> `M%fgfVff
fzDgMzfv-!pv ;ML|f.{pw^90fLnpJgnvf|nTffJfL|
vpf9 NLLN|;fpNv|nT
f,tV*pfs|Lpff pLLDfw^pff a p pv,Vf`!VgM@',ff
fgp|VwVXBff NBfNLL`gwfpp@Jpf
vB@f'pf|%pgpf ||V,fLpg 
{`wTL;L|nfaDgfvp %ffnXVTD'fV|L-fwpf
TLgf!fmfNfwMff
gfpL|n9^sffff |^fpp|gZpfn9f|sfTT{DftvfBf
-gBpfpgf fV`f f LLvf;tfvp%LG|;v|nT
|Lvtp%fffLgff0
fN|pfNTwNff wB|XLLpfs' fpff'Tf^f^J
;f; ||-j,fp8;Jj fV%Z9ft|fjVmB;^
fV 0szffL|vf tg;-fp-gZZf|
vpBBfVsfs fLf Vftf|pf!p'vf%!pf!f;fL
w'gL^DfN,v;,f,'f ffLTTvtvffJpg
LpsLL>fp!!pNftpff;%|
pf8wp^VNNLV| ;f ^wfpL@'GB|v '%'gfX
LL! V%^tp
fzg|LM g68p%|zw
sL^fVpVfZp!T{fVf
Lvf|pfvsLJ'fLL T;pfJnLtf'pgf;^gL|
f'9Lnpfzfjfjnfp nJpMfwp>NBX`''gp
vp|Lf'LfN|BL@% LVwfV p0-fpf tf
fMf^wVG|;tfff,Ln VV,VVf'MT|L'|G@p
fLffp-|G|Zgp@ftf'M ZffNL%wfpf%TZL

65 )+,

65 ' 2

s!

5 ' +,

65 '

65 '+,)

r.

L.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

OEEVM coarse grid


MM coarse grid
OEEVM fine grid
MM fine grid
channel flow, OEEVM
channel flow, MM
channel flow, EXP Wei (1989)
Flat plate, no pressure gradient
Flat plate, with pressure gradient

70

60

50

u+

40

30

20

10

D wB BnNfgD!` D aDM^VvpfX;fT^gVVM^'fM>N|;Lp


%VNpNTf|p%j!V pf
V
0
0
10

10

10
+
y

10

10

afLDf!fLfL|n;fpf
TL^wf|pLBfffp;f^T
LLMfZ|wL|nJZ|
|fgfT8 VB-gJw
fT8Tpjf'fpNf8Np
zpB@fVfg
|LjnBp|Lvv,N9fn
ffL%^fLpfffV,f|f|
fL| LgVBv'%pMVs
ff|fpDfNB^!pD|L D N ffN%f^L|nJ
,@fpBvp,w f%Lp%  f@af8vLL@pMD`LV
p,Bv'%Z%pGVN
.

"+*,j
.- +

 *!4:"

[g|fpfLjf
v LLT;|pNppMwfp
f;LDp;fwLLLf
pB`|V
wV0VTf;'BLLgL
fLVtBtffB|gpvJf
|{Tpffff8.B
f, %Zf-ppBf-V
gZftfpwtffB|ff|
Zfftf8>fLpMf
ff|B'B|X%f|f^ff!L
gfLf%>VwpvXL
!pnpLZf|ff|pw'fp
%fJ^ffgfL.%f%

DfVsvgv|fp'f9V n
ffVT>ZBva>nfpfv
fpfpff|fBff|-f%
fp^|%fwfZL|ntpfL
.|nf'pfB8!f-
fLVf>^fVgtf8vLL
LVtBfZgpfLvwsp
ftpJfftL%9wtvp
JppJzf^L,v|
^sfNpgL%|pf
fV}V@f'ffLfsN
|pffMvaf%

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

CP

0.5

/1032

/5462

-0.5

towed hull OEEVM


towed hull MM
self-propelled hull OEEVM
self-propelled hull MM
EXP DTMB AFF8

D !>|DpMV9p!fTL8vLL
fp'f9B;fV^sf9ggB
g|p v0%ff ^ B D qw gpfg|B|J87 9
vpJ^pJZf{wp
 'f;'fLpsNfZB{^fZfVT
>
p;ffB|g|{ZN|L
gfLTMZTp8DB|MfTggB
g|'|X,LnwpfpVT fNvp@V`gfNpvf
f;B`,tv;p!f;fpJLZ B@pv>%JNLgg
fwp%fpf n%|G|pfwB|fVV fL%,0fjgff*B VVff
pfnT,fffLnpL%pffsL|  ffJfg;pfLp,Tfp
Tt'%fV0'fVwpXfNp !,pT |Lpgp%TVp
LL|np|%Xp'| `^` Z>fwvpf%fpBpf
%t!JLL-ffB|ffZL TfgLZ`fLJ|%pZfwgL8fNLL
.fftVw,pffft| %V %ffff|pf
p9pgLfVwpf%8 !f-f j'gf nggf fL
L|Js%fJ^f-z LpTf,fL`fgs sMpv|
Bf}ffV }| Bfpfpvp,f
pf8wpV^;vLjgpL pvfj pTftB8f|fLgf
ffn@L|sV Lf Zf Bnfffppf9
pgfff.ffLjpf9B%fLL Zp@!,@gffLffNgf'|
fpgvf,pNf%fB8^ VNpffZpgNBfTgfLZpBTa
VG|af@Lnpf%wnffTLnfL ggB8ff|{
fzs%8vpjwVV
fp'aVfB;fp'f; gpfDgj|tB|gJVZ f7 9LpnfL9jp^fL
fV6JZL'%T`Nfg p[jjBNsfVT[!`f
fJp^VTf nfv|nT,fNp{pJf |XfgJT@@L,fNB
p
v'%pBpVff@D^NfBv
`VT f
n  f T f,!`tf,jg;
vpXfwp,%|a%f|f V-fNNfgv!`9f
v|jff vf ff[f L%p%pzfjvX fLp
q
pM';!wvXZVM`gg Mw%L`;vp;
pp;|f8pV^ff|`ff;V fLLL fJppXB
ffwpDfN%ff^^ggB vLLg,p{B8# 7 9 w {%^JT
fpvff|@f-f-pwfJ BB'%^f>V@
LNwfff|fpL j8Lpps B{% ffp

-1

0.2

0.4

0.6

0.8

x/L

=

n

.s.

.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

@LG|`7 9pZvX;nMNML
%N -aLVf VG|p
f;g
ff,f9p
tZ|JfM!JfpMfJ
fLJ|%f@pLNLnfL: 7 9
VXna@f!vXL^pG^B8
L9jgfjnVfJfff
|B%f!p'v@ 9
gf;f;fV^Jg
f|af%; 7 9jV < pf
Mjv;fg' 79p p= < 
gfNpvp^^pf>DMpf" 7 9wLgf
L pJ|Lf%pNT^
BL@pJvp'X ppf'L
v^ntMvp
p j
^f,BgpLvf |nMg8f
LL^L` V%L%LL
f@f@MX@f'fvp
ffpDnpp|
ff.'fsVM E  'gffzpL
vgV^vT!,VM> 7 9;Vpap
wpffpLwfVf0'%
^ff%wzfLvg
|? 7 9v|nTfVMJn
ppTp%fpf!X^f
gTVB MLVp>ffLf
pwp`fZ spf%
Dfg%;|'pfj^f
f!p';Lp&;gf p^fwL
nMs tp,|Lv
nJ|v%" $
%`p^fZfVTsfp z {
MpgT|L'!LfL
NV.6 0fLf, j!`
g @BBswv|nTZ s,f
|LvpBVDLpNZw^a
^f|Lvpp'Vp fLN
f $ & pf% fBLfX fp !ffL
p'Gv|nT,

fLsLL@pf{!w;f'p
pLfp}pvp}-G
LTvpffVpBL%DV>fwg
paBLLjfjp^,BL'fL
ftf'LfZgfL%vV!f
fJgwp>fNgft,f@Ltg
f|
.

1
towed hull OEEVM
towed hull MM
self-propelled hull OEEVM
self-propelled hull MM
EXP DTMB AFF8

r.

#

0.5

-0.5

&%

-1
0.2

0.22

0.24



towed hull OEEVM


towed hull MM
self-propelled hull OEEVM
self-propelled hull MM
EXP DTMB AFF8

.s.


!

0.5

-1

0.87

0.88

0.89

/14@2

0.9
x/L

0.91

0.92

0.93

Dqw gpfg|B|J87 9
j%f,fMpZsf,v'fL
Tp= < p p < pNppspwf
fvLfVM,!`f
v

2.5

5 '&,*)

towed hull OEEVM


towed hull MM
self-propelled hull OEEVM
self-propelled hull MM
towed bare hull OEEVM
towed bare hull MM
EXP AFF1 (bare hull)
EXP AFF8 (hull+sail+rudders)

1.5

r/r

/21

-0.5

0.3

*) +%

/ 1

0.28

midship

r.

/1032

CP

0.26
x/L

0.5

.

.s.

0
0

0.2

0.4

0.6

0.8

v/v

D w' ;Lp 


gf;^g@Lnp^V j {
0

-!

Copyright National Academy of Sciences. All rights reserved.

65 '+,)

Twenty-Fifth Symposium on Naval Hydrodynamics

BA DC(-C(&&&
n

-fVftf8fgf;f
fpfLL;gMpfN@pwf;pf
f%f^vpvf!f`wg
f|ftfJ^ft;wp%LfZ;
gg|@ffgf*LV
f9gwp Tfg%f9fjpg
fLL|f@^pVf;|g
pL|f0ZfvtfBz|LvL
fZfsvg-fp@L
6VGpwfpf%Jf.g
p.VfB8ff|
Lff'%f %pf-f
p'f!fp@|pf^fZVNvB
ffg^jgnfp^fpVf> zf
fgw%fNgwf@JLn
fLfNV{fMLfXLXLV|
LLD|LgDpB!%f|`fTv`|
f,p,fNgfp*VN
fMfGGg;@f'V
pf9gf8Jf%JV^
n%6wvtg9jL
fJBZpwLZfsfp;V;
fN^@^f!|%^pDp^fZL
fp[LVpZZ|L[[f
LfzwsVf.pppff
pwfD^ffTV9LnNp
Lf0fsp[p8pg.s
LNw'LLV}L|n|
pVf'v@fpfVf'pJ
|^afXjpf;tLVm|
Bnspp8pgGMw!npfsn
zf@V B8pfp
L|0wGv|nT
vX^fJLf!%;va{Lap'V
p, f^B;v-p
|fpfwf @n MVV8f
fD' ng %p| w^t%n
ff9sfgz%JLnfL
fVgpNXv;pV
pfjp@fLVNtwJ|L
|pTfLLaZsfJ|
ffpfmBn,|pVf
fLpjf'|
fsBV^v!vgZ,'fVNf'|L

/1032

Kt, 10*Kq

1.5

0.5

0.5

/14@2

1.5

D Mpfp!|f f
VfL'LL^pvV,vTpf
g
BVpfaZLp|B|J|%`
>LfTp8pg^DLG%`L8
w'fVfgLfZfgv!gp%s
zfLVLfwLVJpsf
ZffJLntVG^fZv
v^TpVM@fL%,ff
fst%|.|G|JM!'fL
Bpf8ff|Mpp^v,ffg%f
Dfg fg'; Vf
mTpffn0f L
vfV @BDB LL-Tg
 w^%MgV^pffjn%Tw
 L{N|fp86v%

Kt
10*Kq
Kt EXP Taylor et al
10*Kq EXP Taylor et al
Kt EXP Chesnakas & Jessup
10*EXP Chesnakas & Jessup

.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

zLg[LV|-B;vsp  %B*w
p af|Lv%T@f%gfpff
vsL|8f[ffVNp[L|p
pVgJw|pf-'  LDp
,%fV^'{vVpsf'p
%f fsf fB8}Df
vDfgXN|pf'pvfLp
Egfgsf%-|B|J@
1 F=G %ZpzI HK
J 1 F=G BXfv
^f E >fafg) Jza'% GsfTL
pwL|n@ FsNLgn,p *fNLpf^
f'f @BL'%LpV D  M%gfffV^p
fN%ffpDpf|L
Vfjp V ff6f8v
>|Lpf
MLNC C
ZNtT8BV^VNLpMf
vsff|p0jg wp> ntfJ^|LjL
wpt^Jsf|fpfpf9ffL L|f,fZfgfpV0L|
gfpp;ffL|VfsgpLV
Lf6p%f0Lpp TfNpt^0jg,pnMfpps*w-;wpgg`%
vvpzfVf9fBp
f;p^f8 wp@TZNgfg;V Bfff%f'f^t^fg.f9sfT^fT%.V*gBp|fff
fw! Bvp,>@fT;^'MT wpLsfLfvffpMff
ff%gp|fpfgLfsLLpf
Lv LMfZ n |ffNZZTf^Lgf
ff!pfXMf'fVX!p Lfp# 7 9^Tfa'fLp@XXBp
vO ;vVNp;;LgjLVp%;v; vpffVTp'f@gv'f
[E W B
fL.ptpfftgfjV
!zg .f.fn fVTJD`fwLG|TpZf
|TpLfMvnf@pvaVjV fpjBvBf|fv|nT9f,fVT
fpv[g0pfjgff@LpsV gv,gMfff{MB
ffntpf%'p nf%wv v|f{ jftfjg'p^fgff9Vw
Np-gBJVpBfff'Lgf VfMgp%DG;v^L|{
L LfjfL|Bp wf@LGwp@pf,Bfv
vV@ZfM'LVfMBngf[ %fBsVf^-gV
fv@M'pZLXf;pL ff^pLgf^L%X
%fpLpL|nzLgfftpg fJ%g,BNvMapf'f
BfpVf0pZf*V^nfp v!;Zgfpf^pDfNB
f;ZTptf^@f
pwJv`fJVLvf'pjL
pf-B,f^JwJg^fV}'pff-B p& LfppBpL|n| 
ffpffpfTJj|% %, $ ^p.B Lwv|nT.
pwf@g^L|pBpfffL fL|f.p6fVTzpmfg
gfpq|'^;g!%^!f fNf>V 9 {|LZp
f8vwp`Jffa%fpXf@f fwfX|Lv%pV^{XVMpB
ppvpNwp^|'ptfNVM LV.gsf8v

T

T:

=T

X&

(-

S

(&&

I

5 +,

G/21

6!

Copyright National Academy of Sciences. All rights reserved.

5 '&,*)

Twenty-Fifth Symposium on Naval Hydrodynamics

! `gafgfp6|G|BfMf
pX|f!BgpMV>fMf>L
fpaJp'ff@Bvvp'
^;|vfNfMBBfN`p
fJpvp8pfg|fBgf|
|fV%L,.ft^6
w>fVf
> ,Q P 
-afg%f@af^>p Npf
;f, Vg;t! Bvp |Lp
fzV EWE|f9t9pL
LL8T>wpfB!'LLw!`
swJD`|Lpf-Z%ffng|
s-|'LLMBL|
VNL^N!`^pMNp
VpfJvpzT^
BV^6f jLL[vf| |
fg |f!p^f
spXff>f,VfVN|%p
p,|V0VfwwgZL
|LfffB|wps8ff|fBgf
gfL vp@L
gfNL%gfp;|VVf^!
p6|Bffzg vsTtL
f.f>|LpfffVfzg
vgf^VXLLX fpDXff'
V^TXN ftf
ps'L
LL8
jB,vpftfM^f,;T
LL{wfNpwwpGfpf%
`ZVppf^!|f!pfX
GGMZ;ps^f|Lv%p`V
 zL%DfZLLVM
JBLp%X^f!|LvX^f
pNVp> 'NfffBn>p
ff|Bnaf'B8'N
.f!fzzfjp@fsV
vff'B^Tg'f
L%Vsp9g;pff f
g.L[!Jpfz!
ftL|JJ,gfjg|sL|
p@fT!VfZV|'fzf
|Lpf@wLfpX s8f
fJg8fLLsgp%N
h

sfaL,-LLfp
V{LfVZpvL>fTgpV Nf
;fp^B 'VG%T ff>
%vg!f'ff,fVZf|
|sgwpV
!pfafm^,wpfVf
!f f@gL8fLL;V
fmfv;ppfZa`;|'
Lpfgg0f9LVpg
VT Nsppgv|nT9|Lv%
0 [g;pf!TV-p
|Lf%^gg9f-s|L
LVfp;LL w|
Zfp{LLJpf;^fw
f|gfL0f;pf0Zff
^,%pfgNBfvff
wvf|wpv|LfwLf%
f|f8ff'f@sL@Lf
^f|LvftgL %V^9zL|p
^^pg%Zp|
-ft%'fafpfpfZ g
,fL@fVpfpDLL
Vf;Vff'Vf!pfJL
N|fgw^Lg
vff'|Lv%
J R  S @ 

w0Mfq[ggBfBTLg|
| Vffp6 ^LgfVf @mL
%J f9 G! f|tMLg
[ wJf|,ffVBLB J
Bf}pff%0fffp
Mp' !gnJD`

r.

r.

.s

.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

@ N





L%v ;_ TqDgBfLjDVmL|
!gsp |ITVT  a` LVn
%L
; G
gBNXBVpgfZB'D;\ LT n Z)
>

 LZB;L
V{ 8B>pTLNDV] Tp
!n!; BLpB L
@f j  Z  wp0T
Zb T|0B p!B @Jp L !fg
n LLc Tp BLXD {Bp
@pGt G  pLf
Uf}Bp;>BpV BD@>p L
Lpfzp>%JDVW VM` 
v L%M
@n T T p B  T|
! `% DV 
Lp%3 !gvf p  |
% f pfp %fV
d Tp MfL|Np f
pe YV!f{pB
@Va;  pM j GB T @
|fM pgf ! ^ # !M @
B] Tp vf! f Yp!!pp
!p; B,^X @VXZ Bv
Tp%f0w @V^@G T ff%,p
DVV!^L^f B'fTB t^f Xf
!fLvpwVfBB  !{@L LBBv%p'% g TV
%LLB !f
%pg h
Ug %p D 
|fB,ML ` VML
|fB^8  f V 
pB
%pZ Xwga'vp!MX
{
V > jGf T|^DG*! >T
> B;V  |LL^`L B
d Tpn!  JpV%p%L%
  G: T|'ffg
ZBw>f| L,LpfI TVB%X
!sXD > G%
V; ; Tw>Bp%fZpa`f



^TZ ZLt%@GB T|Bf |L


LtBBnLpfLVf6[ BTV
LLfB;Vs%pfT ` %
DV @ @fL Lwp L
vp%L
qZBpqBBf ; Uf
p DV nBBf!W V
LpXVt%Vp>M@!
>BLLB%p BMf !f
w p|J { ; XM % ; '
T fLf %f%-`p LB;L
VfMDBp^DVI TVB Tf| T
!@L!' Y%L !fn{BBB L
wpLN Zp %s 
TVJBf,Lp LLMT`p L
B;VI TVn pv Xp L
T w @gfsT T p G !v
XZ T|f%%f>f B;V
f[ TVDs%> >{%LB
TpV;> ' {>p-pgNZ
U  Lpg0s`f[L B
^f' ^LZfT ` fW V|!p
>f LLpfgf VpLp
p g{BaL
MgVfzG  T v'wL
ff gf'%`gM%X> ;TV
LpapsBB'>s!Vp!BfLLB'
Z
f> ZpZ>fM \ T>p L;B;L
VfG Laf%@f>%`DV
MV. TV!n! Vp%p
BfLg%ZMpn`% '  >T fV VL pB;pL
Vf] TpqwsMXDs!B
%L% L
f%j^ T>Vwp>f Lj BVfj
M|`V!f!t`p LsB;L
Vf fL sp LL
%TLwZffp
f%;VpfG Nppz !gf
$.

.

w

4.

.

c.

:.

w

.

.

.

r.

6.

2

.

.

$.

u

w

S

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

% Z TBfL LLD
vf|;B;pfpD`% DVw
TM>Z!Bm TVVM BL
% L%sBV
Lff%  T @ pM|pf Lv
^fN f % pf>wMwg
Bv'%I TV %f- G {%
Lppf N6 ZZ%f| j^p[
9G p TM%jf 0fn p
'a^J;B !q YV L
f%. Tp%^LV|w@ \ Yp G>g!!
%MMLn @\ YV @fB
M;{%L
Lp ' T pBfp |ZTB @f
pa%ff B{a`p LNB;V
D`= TVD  > pG LB
D' nZaXp p;a@r T|w Tp
VffM|pfgfLf f
v^,DJBZwfB%D
LTL] Tp vfXT% %BV 
@ @g@Gp ffZ T p
nMLf !'`%Bf`L
f.>p L9B;V. TVXD 
%Vv%s L
- BpNpff; \ T B>!;v
v!s%f|fNpX`%M
DV TVL` > ppGB L
-Z 'Dpv '%pJZvpf%
T9`ff%@Tf  n 
%f`%] TpDM BBf Ta%L
s
Ln mZ BBf ZZp,f%
. T>p sB;pfpfZDV6g^
| fpf!BB TV fLXL %
L%pp
|2 >f| LLT
NM)Lgf{

|!g !fMD
>f B^!;v] Tp !BX!{X!
{Bae i ffpMj
>L@@  T|%fBfsB| 
Mff fLLd TVB;Vp LL
p>%DV VfB^ @Mn
p j! k >kXG{ - a !B !fn
gB ;pV w|pp` f: T|| TV>!B >Gf 
D  vpGB L
w + l" T|Bgfpf f f L;L
fJBJp fLI TV > wf
pMp' !gnspw> p|
vBTLLgLwpnp%p%
G VpGp NVf: T sf L
ffXpBfn'LL @fL LLaX|
L9f f%Z.`%'|@ TV
` > % LGpBs
 p  Zm p
 Uf!f9`gfpBfp LLW V
npv Vv
Va;%ft>p 9Ga T>f L
B;V.`%JDVqVzBf%'
|fm TV!n! pv V%L%p
ff[XNB6![LgfL; T 
fpff |L;>f L
B;V. TVD  ppLBs%L
% n T|w p%fp.>pf
p|^s>f LB;pfI Tf
`f|I o`f|Np^fMg {
' aff!pf BBf- B
B fp{Bs
Lp >f LLpft|'
gN`V *Bf Lv
p
LL ; ^p0Vwp ;; T%Vgf
@Vf[ | f.B;V
p -%!`f% MX `V TV
VpfM;pf; L" l  
l fps-gg-Mv!pff
p LL
8

j.

2

=.

r.

s.

.

.

.

u

`.

(

`.

<

.

8.

X

.

<.

S

w

v.

s2

.

.

I.

w

.

c.

s

w

w

(.

.

.

s.

Copyright National Academy of Sciences. All rights reserved.

"

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics


St. Johns, Newfoundland and Labrador, CANADA, 8-13 August 2004

Improvements to AUV Control Surface Hydrodynamic


Modelling for use in Control
P. Ostafichuk, S. Calisal, D. Cherchas
(University of British Columbia, Canada)
ABSTRACT

Re L

Reynolds number based on vehicle length UL

Improvements to the modelling of control surfaces typical of autonomous underwater vehicles (AUVs) are
examined. An instrumented scale model is used in wind
tunnel and towing tank experiments to measure performance of bowplanes and sternplanes under a range of
operating conditions. The experiment findings are compared to semi-empirical predictions available in the literature. The existing and improved control surface
models are then used in a control and dynamics simulation program.
Although significant differences were noted
between the control surface models from existing predictions and experiment, vehicle performance in simulation was not significantly changed. The findings
suggest that it is not necessary to include detailed models of control surface performance for preliminary control system design and vehicle simulation. Changes to
control surface geometry and layout on the vehicle were
also examined in simulation and were found to have a
significant influence on vehicle performance.

plane thickness; time


freestream or total velocity

NOMENCLATURE
wave amplitude
aspect ratio, 2b exp C or 2b C

b , b exp

plane semi-span, exposed plane span

C
cj

chord length
error sensitivity to parameter j

CL

coefficient of lift

D
E, e j

hull diameter
depth to hull centreline
error, error for parameter j

J 1, J 2

performance indices

vehicle length
timestep, total number of timesteps
pitch rate
hull radius

n, N
q
R

u, v, w
w wave

velocity in x , y , z direction
wave velocity in z direction

x plane

x position of plane

Z' bp

non-dimensional bowplane z-force

Z' bp

slope of non-dimensional bowplane z-force

angle of incidence
bowplane tip vortex strength

bp , sp

bowplane deflection, sternplane deflection

commanded plane deflection

eff

effective plane deflection

control surface taper ratio; wavelength


plane sweep angle

ae

bowplane

sternplane

sp

bp

Z bp

Figure 1: Conventions shown on DOLPHIN AUV

Copyright National Academy of Sciences. All rights reserved.

Z sp

Twenty-Fifth Symposium on Naval Hydrodynamics

INTRODUCTION
Control and stability of an autonomous underwater
vehicle (AUV) depends on appropriate management of
the vehicle control surfaces. In order to compensate for
errors in position or pose, the vehicle controller determines the required orientations of control surfaces to
achieve corrective action. During vehicle and controller
development, particularly in the initial stages, sea trials
are often impractical and prohibitively expensive. Simulation is often the only means available by which to
examine performance. As AUVs are complex and nonlinear, the validity of such simulations is expected to
depend on the accuracy and realism of the vehicle
model used.
There is a wealth of information available to
predict the performance of AUV control surfaces for
simple conditions (see, for example, Dempsey, 1977,
Aucher, 1981, Whicker and Fehlner, 1958, or Lyons and
Bisgood, 1950). For more complicated conditions, it is
common practice to treat an AUV as a set of decoupled
subsystems (Lea et al., 1999); the coupling and unmodelled effects that remain as assumed to be small and are
treated as disturbances (Jalving, 1994, Healey and
Lienard, 1993). In terms of the control system, greater
precision is required in model accuracy as the closedloop bandwidth is increased; conversely, the requirements on model accuracy are relaxed if the closed-loop
bandwidth is decreased (see, for example, strm and
Wittenmark, 1997).
The current work focuses on the inter-relationship between control surface modelling and simulation
for an AUV. Systematic experiments were conducted in
wind tunnel and towing tank facilities with the objective
of improving the understanding of control surface
hydrodynamics. Particular attention was given to nonlinear control surface performance characteristics and
coupling effects. Flow interactions between control surfaces, the effect of the body in yaw and trim, and the
free surface influence were examined in detail. Numerical models characterizing the control surface behaviour
were created and used in a control and dynamics simulation package in the Simulink environment in MATLAB.
Through simulation, the significance of
including or neglecting the new control surface hydrodynamic information was determined. In addition, simulations were used to examine the performance of a
range of control surface configurations. Although not
discussed in this paper, the new control surface models
were also used in an attempt to develop of a more robust
control system. Existing tools used in control surface
hydrodynamic modelling are discussed in the next section. Following sections describe the design and results
of physical experiments and the development and find-

ings of a numerical simulation program for studying


AUV control and dynamics.

EXISTING METHODS IN MODELLING


Various methods are described in the literature
for modelling the performance of low aspect ratio control surfaces. Whicker and Fehlner (1958) give the following relationship for the coefficient of lift for an
isolated low aspect NACA 0015 ratio plane with square
tips
1.8a e
0.1 + 1.6
C L = --------------------------------------------------------- + -------------------------
ae
a e2
1.8 + cos 4 + --------------cos4

(1)

where is the flow angle of incidence, a e is the effective aspect ratio, is the quarter-chord sweep angle,
and is the taper ratio. The effective aspect ratio for an
isolated plane is defined as
2b exp
a e = ------------C

(2)

where b exp is the exposed span (the root-to-tip distance)


and C is the mean chord length. Aucher (1981) provides a similar estimate of force prediction given by
t 2
2a e 1 3 ----
C
C L = --------------------------------------------------------- + 2.1 3
a e2
2.6 + cos 2 + --------------cos2

(3)

where t C is the thickness to chord ratio.


Both of the above predictions of low aspect
ratio wing performance must be modified to account for
the influence of the hull on the planes. Based on a bowplane located on the centreline of a cylindrical hull, the
effective aspect ratio in the prediction of Whicker and
Fehlner (1958) can be adjusted to account for the effect
of hull image vortices. Specifically, the aspect ratio, a e ,
used in Equation (1) can be redefined as
1
R b
a e = --- 1 + --- ---2
b C

(4)

where R is the hull radius and b is the plane semi-span


(hull centreline to plane tip distance). In a similar man-

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

ner, Aucher (1981) suggests augmenting Equation (3)


by the approximation
t 2

R2
*
R
C L = --- 1 ------ * 1 ------ 1 + --- 1 3 ---- (5)
C
2
11
b
b2

where t is thickness, C is chord, and * is given by


R2

b
* = ---- 1 -----2-
C
b

(6)

In addition, for a body-mounted plane the


angle of flow incidence is influenced by the hull angle
and local changes to the flow velocity near the planes.
The local changes may be due to waves, vehicle motion,
or the effect of the hull in redirecting the surrounding
flow field. As noted by Field (2000), the resulting incidence angle, , is described by
w wave qx plane
+ + --------------------------------------- + local
U

(7)

where is the deflection of the plane relative to the


w

qx

w
wave
plane
- is the trim angle, --------------------------------------- is the angle
hull, --U

of the freestream flow to the body at the plane location,


and local is the local variation in flow due to the hull
geometry. The formulation of Equation (7) is shown
graphically in Figure 2.

plane

U
w w wave qx plane

local

resultant
flow vector

Figure 2: Angle of incidence ( ) of flow onto a plane due


deflection ( ), forward motion ( u ), vertical motion ( w ),
vertical wave velocity ( w wave ), pitch rate ( q ) and the
longitudinal position of the plane ( x plane ).

For a cylindrical hull with non-zero trim,


results from inclined cylinders can be used to estimate
plane performance. The flow around an inclined circular cylinder is known to change with radial distance
(see, for example, Aucher, 1981). The corresponding

expression for the local angle of incidence due to a trim


angle, , at a radial distance from the hull centreline, y ,
is
R 2
local ( y ) = 1 + ---
y

yR

(8)

where R is the hull radius. In calm water with constant


trim, Equation (7) can be simplified by integrating the
span-wise force distribution across the plane. If the
slope of the bowplane force to deflection curve is
approximated as linear then Equation (8) can be integrated directly from y = R to y = b . The overall effect
is the same as an effective deflection, eff , in level flight
of

eff

R 2

= + 1 + 1 + --- dy
y

(9)

Since the term in parentheses is constant for a given


geometry, the effect of trim should be equivalent to a
lateral shift in the bowplane force curves proportional to
the trim angle.
Crossflow arising from a number of conditions
such as turning or transverse currents and waves results
in vehicle yaw. Under typical conditions, the total yaw
angle at the bowplanes can reach as much as 20 degrees
(Dominguez, 2000). With non-zero yaw angles, the port
and starboard bowplanes are subjected to different flow
fields. Specifically, on the leeward side of an inclined
circular cylinder there are known to be two counterrotating vortices (see Kubota et al., 1992, for example).
These vortices are roughly aligned parallel to the hull
centreline and affect the flow field around the leeward
side bowplane.
A trailing vortex (tip vortex) is also produced
by the bowplanes as a result of the force generation.
This vortex is known to create changes in the flow field
aft of the bowplanes in the form of downwash. Lloyd
(1974), Hopkin et al. (1990), and Watt et al. (1997) have
examined the implications of the tip vortex from a low
aspect ratio appendage on the performance of a second
low aspect ratio appendage located downstream. The
results of Watt et al. in particular suggest that there
should be a bowplane-sternplane interaction of approximately 25% to 30% on DOLPHIN-like vehicles.
Finally, the free surface influences the flow
around the AUV thereby affecting performance. An and
Smith (1998) have shown that AUV depth following
abilities are significantly compromised at sea state 2 and
above, even at depths of up to three hull diameters.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Considering specifically the influence on submarinelike control surfaces there is limited information available. In calm water, Waldin et al. (1955) experimentally
demonstrated that there is negligible surface effect on a
rectangular plate at an angle of incidence for depths
greater than roughly one chord-length. Surface waves
on the other hand produce a cyclic flow field that diminishes in magnitude with depth. Since the wave induced
velocity is typically small in comparison to the forward
speed of the vehicle, it is common practice to approximate the wave effect on the planes by simply computing
the wave induced change to the angle of incidence (see,
for example, Field, 2000, or Bystrm, 1988). For a single frequency sinusoidal wave, particles trace out circles
with a constant speed of

tions considered is summarized in Table 3. All positions


were measured from the quarter-chord line to the nose
and dihedral was defined as positive with plane tips
inclined toward the free surface.
Table 1: DOLPHIN Specifications
Parameter

Value

overall length, L

8.534 m

hull diameter, D

1m

dry mass

4500 kg

centre of gravity (x, y, z from nose)

(-0.489, 0, 0.022) L

power

350 HP

top speed (while equipped with a tow- 9.3 m/s (18 knots)
fish housed at the keel)

2d

2g- e --------
U wave = a --------

(10)

where a is the wave amplitude, is the wavelength,


and d is the depth. The periodic vertical component of
velocity is
2x
w wave = U wave cos --------- 2g
---------- t

Table 2: Plane Geometries


Name

b exp L

ae

standard (bowplane)

0.0613

3.07

standard (sternplane)

0.0613

3.07

long

0.0810

4.06

short

0.0530

2.66

(11)
Table 3: Plane Locations

and is used as shown in Equation (7). Distinct wave


geometries are superimposed to reconstruct the effect of
a wave spectrum. For this work, a Bretschneider spectrum with significant wave height of 0.88 m (sea state 3)
was used.

Name

EXPERIMENT DESIGN
Experiments were conducted using a scale model in a
towing tank and a low speed wind tunnel both at the
University of British Columbia in Vancouver, Canada.
The 60 m towing tank had a carriage capable of moving
at speeds up to 3.5 m/s and a wave maker. The open circuit wind tunnel had a 1.5 x 2.4 m cross-section and
maximum windspeed of approximately 20 m/s. A
sophisticated one-quarter scale model of the International Submarine Engineering (ISE) DOLPHIN Mark II
AUV was designed and built for this research (see
Figure 1). Basic specifications of the full-scale DOLPHIN are outlined in Table 1.
All-movable, NACA 0025 planes were used
with the model. The planes were interchangeable and
were constructed in a range of different geometries (an
abbreviated list is outlined in Table 2). The model was
constructed such that the plane positions on the vehicle
could also be changed; a shortened list of plane posi-

dihedral

0.261

0.010

standard sternplane

0.906

-0.002

anhedral bowplane

0.261

-30

Y-tail (sternplane)

0.882

0.008

45

keelplanes

0.557

0.225

standard bowplane

The model included specially designed force


transducers in each of the four planes on the vehicle.
The transducers provided lift, drag and spanwise centre
of pressure measurement for each plane. Transducer
signals were filtered and amplified using signal conditioning hardware located on-board the model and were
recorded with a PC-based data acquisition system. The
PC was also used to control plane deflection through
stepper motor actuators.

EXPERIMENT RESULTS
Five main areas were considered through experiment:
the performance of isolated planes; the performance of
body mounted planes; the effect of body orientation on
plane performance; the flow interaction between bowplanes and sternplanes; and the effect of the free surface
on plane performance.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Experiment data collected for an isolated plane


is shown against predictions in Figure 3. The plane
force was non-dimensionalized using the coefficient of
lift, C L , by

rection of Equation (5) was applied to Equation (3). In


addition, the measured bowplane force, Z bp , was non-

Force
C L = ----------------------------1--- 2
U b exp C
2

Z bp
Z' bp = -------------------1
--- U 2 L 2
2

Due to the sign convention used (see Figure 1), C L is


negative for positive deflections. The experiments were
conducted at a chord-based Reynolds number of
5

and then scaled to 1.4 10 6 based on the


method suggested by Barlow et al. (1999). The predictions of Whicker and Fehlner (from Equation (1)) and
Aucher (from Equation (3)) both showed good agreement with the experimental data to the point of stall. As
the results of Whicker and Fehlner were based on a
NACA 0015 foil, Equation (1) was scaled by the ratio of
lift curve slope of the NACA 0025 section to the 0015
section (see Bullivant, 1941, and Jacobs and Sherman,
1937). The RMS errors for the range of deflection up to
stall were 2.6% for Whicker and Fehlner and 1.7% for
Aucher. The shortcoming of the two semi-empirical
predictions is the inability to predict behaviour near
plane stall. With the physical deflection limits of the
planes on DOLPHIN equal to 25, stall is not a significant concern for operation at typical speeds (equivalent
to chord based Reynolds numbers of 1.4 10 6 to
3 10

2.0 10 6 ).

(13)

The experiments were conducted at a Reynolds number


of 2.5 million based on vehicle length and scaled to
35 million, again with the method of Barlow (1999).
Between stall angles, the RMS errors were 2.9% and
3.8% for Whicker and Fehlner and for Aucher respectively.
0.003
0.002
0.001
Z'bp

(12)

dimensionalized by the convention suggested by Feldman (1979). Namely,

0
Experimental Data
-0.001
Aucher (1981)
-0.002

Whicker and Felhner


(1958)

-0.003
-30

-20

-10

0
bp [deg]

10

20

30

Figure 4: Force on a body-mounted plane: Re L = 35 10 6 ,


a e = 3.07 (based on exposed span)

Force to deflection behaviour was measured


for trim angles, , from -10 to +10 in 2 increments.
The force on the bowplanes is shown in Figure 5. The
influence of trim on bowplanes was predicted using the
effective deflection given in Equation (9). For the plane
span and hull radius used in this work, this deflection
was

Experimental Data

-0.2

Aucher (1981)
-0.4

Whicker and Fehlner (1958)

CL

-0.6
-0.8
-1
-1.2

eff = + 1.5

(14)

-1.4
0

10

20
[deg]

30

40

Figure 3: Force on an isolated NACA 0025 appendage:


Re C = 1.4 10 6 , a e = 2.66 (based on exposed span)

Likewise, there was good agreement between


experiment data and prediction for a body mounted
plane as shown in Figure 4. In this case, the correction
of Equation (4) was applied to Equation (1) and the cor-

The bowplane force is expressed in terms of the effective deflection as the independent variable in Figure 6.
As shown, the data for different trim angles are well
described by a single curve. Thus, by using the effective
deflection from above, the representation of the bowplane performance in trim was greatly simplified; this is
desirable for modelling and control system design due
to both simplicity and predictive capability.

Copyright National Academy of Sciences. All rights reserved.

0.002

0.002

0.001

0.001

Z'bp

Z'bp

Twenty-Fifth Symposium on Naval Hydrodynamics

0.000
Trim angle [deg]
-10
-8
-6
-4
-2
0

-0.001

-0.002

0.000
Yaw angle [deg]

10
8
6
4
2

-0.001

20

-20

16

-16

8
0

-8
Error scale

-0.002
-30

-20

-10

10

20

30

-30

-20

-10

bp [deg]

10

20

30

bp [deg]

Figure 5: Effect of trim on bowplane force: Re L = 2.5 10 6

Figure 7: Yaw angle influence on bowplane performance. For


negative yaw angles, the port plane is on the leeward side of
the vehicle: Re L = 2.5 10 6

0.002

Z'bp

0.001

0.000
Trim angle [deg]
-10
2
-8
4
-6
6
-4
8
-2
10
0

-0.001

-0.002
-30

-20

-10

10

20

30

bp, eff [deg]

Figure 6: Effect of trim shown using effective bowplane


deflection: ReL = 2.5 10 6

In addition to trim, the effect of yaw on the


plane performance was also considered. For yaw angles
from -20 to +20 in 4 increments, the force to deflection was measured. As shown in Figure 7, the performance of the port bowplane improved for negative yaw
angles (corresponding to the plane being on the leeward
side of the vehicle). Specifically, the slope of the force
to deflection curve was larger as was the stall angle and
maximum force. There was a corresponding decrease in
performance for positive yaw angles (with the plane on
the windward side of the vehicle).
The difference in performance of the windward
and leeward planes is a concern in terms of vehicle
operation. Namely, for the same deflection, in yaw the
leeward bowplane generates greater force than the windward plane and therefore produces a net roll moment on
the vehicle. As such, care must be taken to independently control port and starboard planes in conditions
where large local yaw angles are encountered. As was
shown in Figure 4, for zero yaw, the planes could safely
be operated up to the physical deflection limit of 25

without significant concern of stall; however, with the


degraded performance of the windward plane with yaw
and premature stall angle, it is possible that stall could
be observed on the windward plane in operation in yaw.
The behaviour of the leeward plane in yaw can
be described by the two counter-rotating vortices
located on the leeward side of the hull. Such vortices
would tend to draw momentum from the freestream and
increase the local flow velocity near the planes. As a
result, the planes would effectively operate in a higher
velocity flow and higher Reynolds number which in
turn would increase both the magnitude of forces as well
as the stall angle. In contrast, with increasing yaw
angle, the windward planes experience greater spanwise crossflow and obstruction from the body and hence
decreased chord-wise flow velocity and diminished performance. A simple potential flow model of the hull
vortex structure accurately predicted the behaviour of
the leeward planes up to 16 degrees of yaw (Ostafichuk,
2004).
Flow induced interactions from the bowplanes
to the sternplanes were examined as part of the wind
tunnel and towing tank research. Helium bubble path
visualization was used to map the path of the bowplane
tip vortices in relation to the sternplanes (see Hale et al.,
1971, for further information on this technique). The
sample flow visualization images in Figure 8 show the
entrainment of neutrally buoyant helium-filled soap
bubbles in the bowplane tip vortex and the path of this
vortex relative to the vehicle. In particular, the second
image demonstrates that the bowplane tip vortex persists along the length of the vehicle and that the path is
in close proximity to the sternplane.
Forces were measured on the sternplanes while
the bowplane deflection was varied and the sternplane
deflection was held at zero. An example of the force

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

0.0004

0.0002

Z'sp

measured on the sternplanes as a function of bowplane


deflection is shown in Figure 9. The magnitude of this
interaction force for the standard vehicle configuration
is approximately 22% with the force on the sternplanes
directed opposite to the force generated by the bowplanes. With bowplanes oriented with 30 degrees anhedral the average interaction force diminished to roughly
5%, also shown in Figure 9. From a modelling and controls perspective, the interaction is important since the
induced sternplane force not only opposes the desired
force of the bowplanes but it also generates a pitching
moment due to the large distance from the sternplanes to
the vehicle centre of gravity. As shown, the interaction
is also highly sensitive to control surface layout.

0.0000
Standard Planes
-0.0002

Anhedral Bowplanes
Predicted Interaction

-0.0004
-20

-15

-10

-5

10

15

20

bp [deg]

Figure 9: The interaction force on sternplanes resulting from


the bowplanes is shown versus bowplane deflection for the
standard vehicle configuration and for a configuration with
anhedral bowplanes. The sternplanes were held at zero
deflection. A comparison is made between experiment and
predicted values from a potential flow model (Ostafichuk,
2004). Re L = 2.5 10 6

The value of value of v in Equation (15) was deter(a)

mined from the measured C L and the known geometric

(b)
Figure 8: Sample helium bubble flow visualization images
showing (a) the migration of bubbles to the vortex core,
Re L = 2.5 10 6 , = 0 , bp = 16 ; and (b) the path of the
tip vortex (identified by the arrow) in relation ot the vehicle
Re L = 2.5 10 6 , = 6 , bp = 8

In Figure 9, the predicted interactions based on


a potential flow model are also shown. The model consisted of a semi-infinite tip vortex for each bowplane
with matching image vortices to represent the hull. The
strength of the tip vortices, v , was approximated as
being equal to the average bound circulation on the
plane
C L Ub exp
v --------------------ae

(15)

parameters. As suggested by Lloyd (1974), the path of


the tip vortices was linearly approximated based on the
tangential velocity induced by the vortex system at the
longitudinal location of the bowplanes. The resulting
induced velocity at each point along the sternplane span
was then calculated. With the previously measured
sternplane lift-curve slope, the net sternplane force was
computed by integrating the force contributions along
the plane span. For complete details of the development
of this model, the reader is referred to (Ostafichuk,
2004).
The agreement between the measured and predicted bowplane-sternplane interaction shown in
Figure 9 is exceptionally good, particularly for deflections prior to stall. The RMS error between prediction
and measurement for the standard and anhedral configurations was respectively 0.039 and 0.014 (measured in
units of Z' ). In total, the interaction was measured for
twelve different configurations of plane geometry and
placement on the vehicle; the average RMS error for all
configurations was 0.032 (indicating that, on average,
the agreement was better than the standard configuration shown in Figure 9).
In addition, the bowplane-sternplane interaction was measured for trim angles from 10 and yaw
angles from 20. As would be expected, the interaction changed significantly as the body orientation was
varied from the straight and level condition. For illustration, the effect of trim angle on the interaction force is
shown in Figure 10. Specifically, the figure shows the

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

ratio of the slope of the sternplane interaction force


(from Figure 9, for example) to the slope of the bowplane force (from Figure 4, for example). (The maximum interaction occurred for a trim angle of
approximately 1.5 degrees because the sternplane location was at an elevation above that of the bowplanes.)
In developing a control system to account for plane
interaction it is therefore important to include the variations in interaction with trim and yaw.

(16)

Also shown in the figure is a prediction of the wave


induced force amplitude based on linear wave theory.
The form of the prediction is
4d

---------2g
Z' wave, g = C 0 ---------- a 2 e

(17)

where is the wavelength, a is the wave amplitude,


and C0 is a constant that depends on vehicle geometry

0.3

Z'sp
Z'bp

Z wave
Z' wave, g = ---------------1
--- gL 3
2

0.2

and the plane force coefficient ( C 0 = 6.96 10 6 in

0.1

this case). The agreement between prediction and measurement is good for d D 1 . At d D = 0.5 , hull
immersion was noted and as a result the prediction overestimated the wave induced force.

0.0
-10

-5

0
[deg]

10

15

Figure 10: The effect of trim, , on bowplane-sternplane


interaction force is shown using the ratio of sternplane lift
curve slope, Z' sp , to the corresponding bowplane lift curve
slope, Z' bp . The interaction quickly diminishes for large trim
angles. The experiment data is represented by a polynomial
curve fit.

30.0
Experimental Data

25.0

Predicted Value
20.0
Z'wave,g

-15

15.0
10.0
5.0

Finally, testing was conducted in a towing tank


to establish the effect of waves and the free surface on
the performance of the planes. Depths from 0.5 to
3.5 hull diameters were considered (measured from the
hull centreline to the calm water free surface). In calm
water, there was no measurable difference in plane performance for the range of depths considered (as was
suggested by Wadlin et al., 1955). In the presence of
waves, the induced cyclic force on the planes was measured for varying vehicle depth, speed, and wave geometry. The amplitude of the wave induced force for a
stationary vehicle is shown in Figure 11 as a function of
the depth, d , non-dimensionalized by hull diameter, D .
In this case, the wave had a sinusoidal profile with a
wavelength of 1.14 vehicle lengths and an amplitude of
0.043 vehicle lengths. As the forward speed was zero
for these measurements, the force amplitude on the
planes was non-dimensionalized in an alternate form
from Equation (13) using

0.0
0

0.5

1.5

2
d/D

2.5

3.5

Figure 11: The amplitude of the wave induced force on the


planes non-dimensionalized using gravity, Z' g , is shown as a
function of the depth in hull diameters, d D . Regular waves
with = 1.14 L and a = 0.043 L were used. Experiment data
is compared to the prediction from linear wave theory.

In a similar manner as above, the amplitude of


the wave induced force on the planes was measured for
changing vehicle speed. Tests were conducted for two
depths and for Froude numbers based on vehicle length
from roughly 0.1 to 0.75. As shown in Figure 12, the
relative influence of the waves on the planes diminished
with increasing speed. In this case, the forces were nondimensionalized in the standard form. Referring to
Equation (13), the decrease in wave influence can be
explained by the form of the non-dimensionalization in
that the same wave orbital velocity was normalized by a
larger value of U with increased speed. Similarly, the

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

maximum incidence angle of the waves onto the planes


decreased as the forward speed increased. As was
observed with depth, the predicted wave force amplitude based on linear wave theory closely agreed with
measured values.

WD

Navigation
Module

r+

Controller

Sensors
and Filters

Submarine
Model

Data
Recorder

1.0
Experimental Data
0.8

SN

Predicted Value

d/D = 1.5

Figure 13: The block diagram used in simulations is shown.


The navigation module determines the reference trajectory (r)
which is compared to the measured output (y) to give the error
(e). The controller determines the required plane deflections
( c ) for corrective action based on the error. The system

Z'wave

0.6

0.4

operates in the presence of wave disturbances (WD) and sensor


noise (SN).

d/D = 3.5

0.2

0.0
0

0.1

0.2

0.3

0.4
Fr

0.5

0.6

0.7

0.8

Figure 12: The amplitude of the wave induced force on the


planes, Z' wave , is shown as a function of Froude number, Fr ,
for two depths.
Regular waves with = 1.14 L and
a = 0.043 L were used. Experiment data is compared to the
prediction from linear wave theory.

Overall, the tow tank tests demonstrate that it is


possible to accurately predict the wave induced forces
on planes as a function of both forward speed and depth
if the wave and vehicle geometries and the plane performance characteristics are known.

SIMULATOR DEVELOPMENT
Simulations were conducted to determine the importance of accurate control surface hydrodynamic modelling and to examine the effect of control surface
geometry and location on vehicle performance. The
vehicle control and dynamics simulation package was
developed in the Simulink environment of MATLAB based on work originally conducted by Field
(2000). The layout of the simulator is shown in
Figure 13. In short, the navigation module determined
the appropriate trajectory for path following based on
the current vehicle position and velocity; the controller
computed the required plane deflections for corrective
action to an error in position or velocity; the submarine
model represented the dynamics and hydrodynamics of
the submarine; and realistic measurements of the vehicle state were modelled with noise and filters and then
used to complete the feedback loop.
The information describing the hydrodynamic
performance on control surfaces was contained in the
submarine model block. The structure of this block is

shown in Figure 14. The commanded plane deflections


output from the controller, c , were adjusted to account
for the dynamics of the control surface actuation system
(rate and saturation). The resulting physical deflections,
, were combined with the change in incidence angle
due to the vehicle orientation, motion, and wave effects
to give the net incidence angle, . (This is the same
angle expressed Equation (7).) With , the plane forces
were computed based on methods outlined below and
were combined with the hydrodynamic, hydrostatic,
gravitational, and wave induced body forces to determine the net force on the vehicle. The product of this
force vector and the inverse mass matrix gave the vehicle acceleration which was in turn integrated to give the
vehicle output state.
Initial simulations were based on control surface hydrodynamic models drawn from semi-empirical
information in the literature (as outlined in the Existing
Methods in Modelling section); this represented the
baseline condition. The force to deflection behaviour of
the planes and the influence of waves on the planes were
drawn from experimentally measured relationships but
could equally have been based on existing methods due
to the close agreement shown in the Experiment
Results section. The effects of trim, yaw, bowplanesternplane interaction, and plane stall derived from the
experiments were included into the baseline control surface model. The effects were included separately to
determine the influence of each on overall vehicle performance; they were also included simultaneously to
produce the best representation of control surface
behaviour.
The test manoeuvre shown in Figure 15 was
used in sea state three with waves travelling in the x
direction (the sea state was constructed from a
Bretschneider spectrum). The manoeuvre was used to

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Wave
Influence on
Flow Angle
Transport
Delay
Wave Forces
on body

Actuator
Dynamics

Plane
Saturation

Plane
Forces

Inverse Mass
Matrix

Integrator

Body Forces
Body
Influence on
Flow Angle

Figure 14: The submarine block diagram is shown. The plane forces are computed based on the computed angle of incidence, ,
and are combined with body forces. The vehicle acceleration is computed and is integrated to determine the vehicle state.

measure the ability of the AUV to maintain straight and


level flight in the presence of wave disturbances and
changes in depth and heading. The commanded forward
speed of the vehicle was 6 m/s through the entire
manoeuvre.
60 m

For a single time-step, n , the weighted error was


defined by
En = ( cy ey + cz ez + c e + c e + c e )

(18)

where e j represents the error in state j compared to the

120 m
x

reference value and c j represents the error weight for


that state. The relative error weights correspond to the
sensitivity of sonar image distortion to error in each
state and they are shown in Table 4.

depth change
R

2R
~145 m run-out

Table 4: Error Weights in Performance Indices


Parameter

Symbol

Error Weight

lateral position

c y = 20 m 1

vertical position

c z = 20 m 1

roll angle

c = 36 rad 1

Figure 15: Test Manoeuvre

trim angle

c = 36 rad 1

With the output state of the vehicle described


by 12 parameters (six positions and six translations) it
was impractical to quantify simulated performance with
traditional quantities such as settling time, peak over
shoot, or steady state error. Instead, relative performance was quantified by two performance indices representing the quality of a sonar image in a hypothetical
scanning operation. The indices were based on the
weighted error, E , over the manoeuvre shown above.

yaw angle

c = 44 rad 1

x
3.0 m
3.5 m
z

The first performance index, J 1 , was based on


the average error over the straight portions of the
manoeuvre
1
J 1 = ---N

n=1

0 during the turn


E
otherwise
n

Copyright National Academy of Sciences. All rights reserved.

(19)

Twenty-Fifth Symposium on Naval Hydrodynamics

where N was the total number of time-steps not including the turn (N = 2700 for results presented in this
paper). The second performance index was the significant amplitude from the error response spectrum for
straight line motion in head seas of sea state three. That
is,

Table 5: Influence of Plane Modelling on Performance

J2 = 2

lined in the Experiment Design section) were each


independently incorporated to the standard model.
Finally, the condition where all the new hydrodynamic
effects were included, thus forming the best estimate of
control surface performance, was considered. The
resulting performance is detailed in Table 5.

S E ( n )

(20)

n=0

where S E is the square of the response amplitudes from


the Fast Fourier Transform of the weighted error, E , and
is the size of the frequency bands in the spectrum.
The J 1 index measured path following performance
while the J 2 index measured wave disturbance rejection; for both J 1 and J 2 , better performance is indicated
by lower values.

SIMULATION RESULTS
Four aspects in particular were considered through the
simulations:
the influence of an inaccurate model of control surface hydrodynamic performance on vehicle behaviour in simulation
the influence of an inaccurate model of control surface hydrodynamic performance on the control system design
the influence of changes in control surface geometry
to vehicle performance
the influence of changes to control surface placement on the vehicle to vehicle performance
The results described in this paper are based on
an LQG/LTR controller as outlined by Field (2002). It
must be stressed that the effect of control surfaces on the
overall vehicle performance is jointly dependent on the
hydrodynamic characteristics and the utilization (controller design). Although not presented here, additional
work has been conducted with PD, sliding mode, and
fuzzy logic to examine the control system influence.
The reader is referred to (Ostafichuk, 2004) for further
information.
Six different cases involving the accuracy of
the control surface hydrodynamic model used in simulation were examined. These simulations were conducted
in calm water only. The baseline condition was a standard model developed from existing information in the
literature (as given by Aucher, 1981, and Whicker and
Fehlner, 1958, for example). The effects of trim, yaw,
plane stall, and bowplane-sternplane interaction (as out-

Condition

J1

Max ( E n ) 10 3

standard model

2.63

48.9

trim effect

2.35

39.7

yaw effect

2.48

51.8

plane stall

2.51

42.7

plane interaction

2.20

39.4

all effects

2.27

58.1

Interestingly, the performance of the vehicle


improves slightly on average when the new hydrodynamic effects are included in the model of control surface performance. In other words, the performance of
the actual vehicle would be expected to be better than
predictions obtained using a simulation based on simplified models of control surface hydrodynamic behaviour.
The large maximum error found when all the new
hydrodynamic effects were included was a result of a
large error in heading at the exit of the turn.
The vehicle controller was designed based on
an assumed model of the vehicle performance. This
model is a linearized and simplified version of the actual
vehicle (which is represented by the submarine model
in Figure 14). In Table 6, results are shown for the
influence of the control surface hydrodynamic model
used in controller design. Above, the controller was
designed based on the standard modelwithout knowledge of the new hydrodynamic performance informationand only the simulator block representing the
physical vehicle was modified. The current simulations
were conducted in the same manner as above in that one
effect was considered at a time. As shown in the table,
the only condition that affected vehicle performance
was the bowplane-sternplane interaction; including the
interaction effect in the controller design degraded performance. This is because during a commanded change
in depth, the interaction force on the sternplanes generated a pitching moment that tended to orient the vehicle
in the direction of the depth change. When this aspect
of performance was included in the control design, the
pitching moment was negated which therefore caused
the vehicle to travel at a more level trim. The end result
was that it required longer to complete the depth change
manoeuvre and thus increased the average error.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Table 6: Influence of Modelling in Controller on Performance

Table 8: Plane Position Influence on Performance

Condition

J1

Max ( E n ) 10 3

Condition

J1

J2

Max ( E n ) 10 3

standard model

2.27

58.1

standard planes

15.4

479

64.3

trim effect

2.27

58.1

anhedral bowplanes

16.0

231

55.3

yaw effect

2.27

58.1

Y-tail

13.6

430

32.3

plane stall

2.27

58.1

keel planes

33.8

338

172.3

plane interaction

2.42

63.9

18.5

430

50.6

all effects

2.42

63.9

long anhedral bowplanes with Y-tail

There were significant variations in performance when the plane geometry and placement were
varied. Results for an abbreviated list of test cases is
summarized in Table 7 and Table 8. The plane geometries considered included the standard bowplanes and
sternplanes; the long bowplanes with the standard sternplanes; the short bowplanes with the standard sternplanes; and the standard bowplanes with the long
sternplanes. The plane locations considered included
the anhedral bowplanes with the standard sternplanes;
the standard bowplanes with a Y-tail (dihedral sternplanes); and the keel planes with the standard sternplanes. (See Table 2 and Table 3 for the geometric
details of these configurations.) In all cases, simulations
were conducted at sea state three.
Table 7: Plane Geometry Influence on Performance
Condition

J1

J2

Max ( E n ) 10 3

standard planes

15.4

479

64.3

long bowplanes

12.6

222

39.1

short bowplanes

20.4

397

71.2

long sternplanes

14.1

308

49.8

The best path following, indicated by the J1


performance index, was observed with the long bowplanes. The long bowplanes also had the smallest maximum error as well as the best wave disturbance rejection
properties, indicated by the J2 index. Of the different
plane positions considered, the best path following (J1)
was found with the Y-Tail. The Y-Tail wave disturbance
rejection performance (J2) was slightly better than the
standard plane configuration; the best J2 performance
was noted with the anhedral bowplanes.
In short, it appears that significant benefit
would be derived from increasing the length of the current bowplanes for this particular vehicle. The combination of all the best configurationslong anhedral

bowplanes with the Y-tailwas also examined. Unfortunately this combination did not show any significant
improvement in performance compared to the standard
configuration.

CONCLUSIONS
The performance of all-movable control surfaces on an
AUV was examined through experiment and simulation.
The experiments showed that existing methods in the
literature used to predict the performance of isolated and
body mounted planes are reasonably accurate. These
methods do not model stall behaviour so the physical
operating range must be sufficiently small. The wave
influence on plane performance measured in experiments also closely agreed with analytically derived predictions from linear wave theory. Significant changes in
plane performance were noted with changing body orientation (trim and yaw). In addition, there was strong
interaction (downwash) measured on the sternplanes as
a result of the bowplane tip vortices. The interaction
was accurately predicted with a simple potential flow
model and was found to be strongly dependent on control surface layout and body orientation.
The descriptions of hydrodynamic performance of the planes drawn from the literature and from
the experimental work were incorporated in a control
and dynamics simulation program in MATLAB. Simulations showed that the differences in the models used
to represent control surfaces behaviour had only limited
effect on the overall vehicle performance. The control
was designed based on an assumed model of plane
hydrodynamics but again only limited differences in
vehicle performance were noted as the plane model used
in controller design was changed. Finally, the influence
of changes to control surface geometry and placement
on the test vehicle was examined with the simulations.
Significant changes in vehicle performance were noted;
in particular, lengthened bowplanes and planes oriented
with dihedral (or anhedral) typically showed either

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

improved path following, improved wave disturbance


rejection properties, or both.
As noted previously, the vehicle performance
is dependent not only on the hydrodynamic behaviour of
the control surfaces but also on the utilization through
the control system. The design of the controller is an
integral part of the vehicle and should be considered in
conjunction with control surface design. Differences in
the simulation results should be expected if the vehicle
control system design was changed.

REFERENCES
An, E. and Smith, S.M., "An Experimental Self-Motion
Study of the Ocean Explorer AUV in Controlled Sea
States." IEEE Journal of Oceanic Engineering, Vol. 23,
No. 3, July 1998.
strm, K.J., and Wittenmark, B., Computer Controlled
Systems, 3rd ed., Prentice Hall, NJ, 1997, pp. 183-6.
Aucher, M. Submarine Dynamics, Ingenieur general
de l'armement, Excerpted from Sciences et Techniques
de L'Armement, Memorial de lArtillerie francaise,
Paris, France. 1981.
Barlow, J.B., Rae, W.H. Jr., and Pope, A., Low-Speed
Wind Tunnel Testing, 3rd Ed., John Wiley and Sons,
Toronto, 1999, pp. 315-20.
Bullivant, W.K., Tests of the NACA 0025 and 0035
Airfoils in the Full-Scale Wind Tunnel. NACA Report
No. 708, 1941.
Bystrm, L. "Adaptive Control of a Submarine in a
Snorting Condition in Waves." Proceedings of Warship
88, the International Symposium on Conventional Naval
Submarines, 3-5 May, 1988.
Dempsey, E.M., Static Stability Characteristics of a
Systematic Series of Stern Control Surfaces on a Body
of Revolution, Report 77-0085, August, 1977,
DTNSRDC, Bethesda, Maryland.
Dominguez, R. Effects of Yaw on Bowplane Performance on DOLPHIN. Dissertation, University of British Columbia, Canada, 2000.
Feldman, J. "DTNSRDC Revised Standard Submarine
Equations of Motion." report DTNSRDC/SPD-0393/09,
June 1979, David Taylor Naval Ship Research and
Development Center, Bethesda, Maryland.

Field, A.I. Simulation, Modelling, and Control of a


Near-Surface Underwater Vehicle, Dissertation. University of British Columbia, Canada, 2000.
Healey, A.J., and Lienard, D., Multivariable SlidingMode Control for Autonomous Diving and Steering of
Unmanned Underwater Vehicles, IEEE Journal of
Ocean Engineering, Vol. 18, No. 3, July 1993.
Hopkin, D., Davies, M., and Gartshore, I., "The Aerodynamics and Control of a Remotely-Piloted Underwater
Towed Vehicle." Canadian Aeronautics and Space Journal, Vol. 36, No. 3, Sept., 1990, pp. 122-9.
Jacobs, E.N., and Sherman, A., Airfoil Section Characteristics as Affected by Variations of the Reynolds Number. NACA Report 586, 1937.
Jalving, B., The NDRE-AUV Flight Control System,
IEEE Journal of Oceanic Engineering, Vol. 19, No. 4,
October 1994.
Kubota, H., Jia, M.Y., Watanaki, T., and Susuki, K.
Fluid Dynamics of High Angle of Attack. Proceedings of the IUTAM Symposium, Sept., 1992, pp. 27788.
Lea, R.K., Allen, R., and Merry, S.L., A Comparative
Study of Control Techniques for an Underwater Flight
Vehicles, International Journal of System Science, Vol.
30, No. 9, 1999, pp. 947-64.
Lyons, D.J., and Bisgood, P.L. An Analysis of the Lift
Slope of Aerofoils of Small Aspect Ratio: Including
Fins with Design Charts for Aerofoils and Control Surfaces. Reports and Memoranda No. 2308, 1950, Ministry of Supply, London.
Ostafichuk, P.M. AUV Hydrodynamics and Modelling
for Improved Control, Dissertation, University of British Columbia, Canada, 2004.
Wadlin, K.L., Ramsen, J.A., and Vaughan, V.L., Jr.,
"The Hydrodynamic Characteristics of Modified Rectangular Flat Plates Having Aspect Ratios of 1.00, 0.025,
and 0.125 and Operating Near a Free Water Surface."
NACA Report 1246, 1955.
Watt, G.D., Seto, M. and Brockett, T., "Hydrodynamic
Considerations for Semi-Submersible Minehunting
Vehicles." Proceedings of the Fourth Canadian Marine
Hydromechanics and Structures Conference, Ottawa,
25-27 June 1997.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Whicker, L.F., and Fehlner, L,F., Free-Stream Characteristics of a Family of Low-Aspect-Ratio, All-Movable
Control Surfaces for Application to Ship Design,

Report 933, December, 1958, David Taylor Model


Basin, Bethesda, Maryland.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics


St. Johns, Newfoundland and Labrador, CANADA, 813 August 2004

Submarine Rising Stability:


Quasi-Steady Theory and Unsteady Effects
George D. Watt (Defence Research and Development Canada Atlantic)
Hans-J
urgen Bohlmann (Howaldtswerke-Deutsche Werft, Germany)
ABSTRACT

x, u, X
p, K

Submarines that blow ballast at depth to rise in an


emergency can develop an underwater roll instability. Trials data from several boats are presented
showing the instability. Previous work on this problem focused on the destabilizing hydrodynamic effect
of the sail pointing into the crossflow during buoyant ascent. A new three degree-of-freedom stability analysis confirms this phenomenon as the most
likely source of the instability. The analysis also predicts a new yaw instability that occurs when only
the forward ballast tanks are blown or when negative pitch rates occur. Strong nonlinearities in flow
incidence, orientation, and rotation are accounted
for. Unsteady effects are partially accounted for.

q, M
r, N

y
z
W = mg
x, y, z
x0 , y 0 , z 0
xB , y B , z B
xF W , z F W
xG , y G , z G
x
X, Y, Z
XP
r , s , b

, ,

NOTATION
Standard notation is used (Feldman, 1979). Nondimensionalization is indicated by a prime 0 and
is based on `, , and U ; eg, m0 = m/( 21 `3 ),
g 0 = g`/U 2, q 0 = q`/U , etc.
B = V g Buoyancy.
BG

Height of the center of buoyancy above


the center of gravity.

g
F V , FH

Gravitational constant.
Vertical and horizontal plane forces or
moments.

m, m0
K, M, N
`
p, q, r
R
u, v, w

Mass, unblown mass within V .


Body axis moments.
Overall length of the submarine.
Body axis angular velocities.
Reynolds number, based on `.
Body axis velocities.

Overall speed: u2 + v 2 + w2
Volume of the external hydrodynamic
envelope, including main ballast tanks.

U
V

y, v, Y

z, w, Z

U
v

v 2 + w2

Weight within V .
Submarine body fixed axes.
Inertial (earth-fixed) axes.
Center of buoyancy (centroid of V ).
Sail center of pressure coordinates.
Center of gravity (centroid of m).
Axial coordinate of blown mass centroid.
Body axis forces.
Effect of propulsion on X force.
Rudder, stern, and foreplane deflections.
The small quantities: y 0 , , , v, p, r.
Yaw, pitch, and roll Euler angles.

Flow incidence: tan1 ( v 2 + w2 /u), always positive.

allowed to go negative when jumps


by ; for graphing purposes only.

Blown mass fraction: (B W )/B .


Fluid density.
Flow orientation: tan1 (v/w).

The following relations are useful:


u = U cos
w = U sin cos
1

Copyright National Academy of Sciences. All rights reserved.

v = U sin sin
p
v 2 + w2 = U sin

(1)

Twenty-Fifth Symposium on Naval Hydrodynamics

A dot as an over-script indicates differentiation with


respect to time; eg, u = du/dt. Force is often used
in a generic sense to mean force or moment.
1

second order products in lateral force, rolling moment, and yawing moment damping terms. The Ecriterion stability limit does not explain the diesel
submarine instabilities discussed in this paper.
Booths second F3 -criterion requires a numerical solution. It contains hydrostatic terms and,
Booth suggests, becomes relevant at high incidence
and when the blown mass centroid is well forward.
However, this and the E-criterion instability are just
two of five possible stability criteria that can be extracted from his analysis. This has not been pursued in favor of the analysis presented below, which
is thought to be more general. Booths effort was
admirable given that his analytical analysis would
have been done by hand, a limitation that modern
mathematical software has eliminated.
Booth (1977b) also considers the near vertical ascent case (very high flow incidence). He retains the
same coefficient model as for forward motion but,
appropriately, carries out special high incidence experiments to acquire coefficient values that better
describe this flow regime. As he acknowledges, his
quasi-steady stability analysis does not model the
unsteady forces associated with separation and vortex shedding at very high incidence.
In both of Booths papers, qualitative results
from experimental tank tests are presented. These
tests, with an unpropelled buoyant model, show that
roll is coupled strongly with yaw, a significant result.
Papoulias and McKinley (1994) also calculate the
stability of a submarine-like body during buoyant
ascent. Their analysis uses the same second order
coefficient model throughout the entire flow regime
which includes pitch and roll angles of 180 degrees.
Although such a model cannot be realistic, unsteady
forces aside, they suggest that it should nevertheless
provide qualitatively meaningful results. They discover an interesting inverted pendulum stability at
extreme pitch angles and discuss a roll instability
about this point, but this is of little interest to submariners who prefer to maintain low to moderate
pitch angles and are almost always able to do so. It
is not clear what the flow incidence angles are in this
analysis.
Haarhoff and Sharma (2000) consider the horizontal plane stability analysis of a surface ship. They
provide a good description of the mathematics and
physical insight for the terms in their equations.
The above methods analyze stability in 3 DOF
and are based on body axes state variables v, p, r
and roll angle , with the assumption that yaw angle is unnecessary because it does not explicitly
appear in the equations of motion. We show below

INTRODUCTION

Submarines that blow ballast at depth to rise in an


emergency can experience large underwater roll angles (Itard, 1999). Previous work (Watt, 2001) investigated the destabilizing hydrodynamic influence of
the sail, which points into the crossflow during buoyant ascent, as the source of the instability. This phenomenon does result in instability at high speeds and
low static stability levels, but does not explain the
early onset of the instability in several diesel-electric
submarine trials. This remains an unsolved problem.
In this paper, we examine previous attempts to
analyze buoyant ascent. Trials data are presented
showing the nature of the instability in various
diesel-electric submarines. Static model test data
spanning the flow incidence and orientation angles
seen during a buoyant ascent are presented and used
in a new stability analysis. Nonstandard modelling
techniques are required to adequately describe the
nonlinearity in the data. The analysis explains an
instability that occurs when only the forward ballast
tanks are blown or when negative pitch rates occur.
The stability limits presented below depend on modelling the nonlinearity in the forces, accounting for
unsteady effects associated with trailing sail vorticity, and accounting for heading ( ) in the stability
analysis, a term that is usually ignored.
2
2.1

PREVIOUS WORK
Stability Analyses

Classical submarine stability analyses neglect many


terms and linearize the equations of motion in all
state variables (Hoyt and Imlay, 1948, Lambert,
1956).
Booth (1977a) analyzes buoyant ascent stability by including nonlinear vertical plane incidence
terms as parameters independent of his horizontal
plane stability analysis. His three degree-of-freedom
(DOF) analysis, based on the lateral force, rolling
moment, and yawing moment equations of motion,
uses conventional coefficient modelling (Gertler and
Hagen, 1967) and considers predominantly forward
motion. It accounts for realistic pitch and incidence angles and is applicable to the current problem. Booth establishes two criteria for stability. His
E-criterion is a simple analytical formula setting
a stability limit on blown mass that is independent
of speed and static stability. It is determined by
2

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Given the above information, it is desirable for


a submarine to rise at incidence angles less than 30
degrees, so that its motions are at least predictable.

that should be included and, in fact, that its inclusion leads to a new stability limit.
These stability analyses are generally complex,
with limited analytical results available. Watt
(2001) does obtain a simple analytic expression modelling the balance in rolling moment between the stabilizing static and destabilizing sail hydrodynamic
forces in an ascending submarine. This 1 DOF analysis in extracts only a single stability limit and
ignores any coupling from yaw and sideslip motions.
Nevertheless, we show below that the result still provides a good estimate for this mode of instability.
The current 3 DOF analysis restricts itself to the
0 to 30 degree incidence range where model test data
are available that span the flow incidence and orientation angles (, ) of interest in the trials. That
is, we do not consider very high flow incidence where
unsteady vortex shedding makes maneuvering inherently unpredictable.

2.3

Unsteady Effects Under 30 Degrees


Incidence

Submarine maneuvering simulations are traditionally based on the Gertler and Hagen (1967) quasisteady equations of motion. However, the quasisteady assumption in these equations limits their
usefulness.
Tinker (1978) studied unsteady effects on an unappended, blunt nosed body of revolution and concluded there were no appreciable fluid memory effects that needed to be accounted for. But he did
not consider the interaction of body vortices with
any appendages, or between sail shed vorticity and
the tailplanes.
Feldman (1975,1978) adds quasi-unsteady terms
to Gertler and Hagens coefficient based model.
These account for time history effects resulting from
vorticity trailing from the sail in an unsteady flow.
This vorticity is generated at the sail by the crossflow at the sail. The vorticity convects downstream
generating out-of-plane force on the afterbody (force
normal to the local crossflow) and rolling moment on
the tail. These are proportional to the local vortex
strength, which is itself proportional to the crossflow
at the sail when the vorticity was generated, approximately d/u seconds previously, where d is the local
distance aft of the sail. So quasi-steady theory is
used to predict locally generated vorticity and force,
but vorticity is allowed to convect unsteadily.

2.2 Very High Incidence


Wichers Schreur (1990) documents the oscillatory
motions experienced by a buoyant, nominally horizontal circular cylinder rising vertically with no forward speed with and without simulated deck and
appendage geometry, albeit at low Reynolds numbers. He found that bare cylinders experienced unpredictable sway and yaw motions, attributed to unsteady vortex shedding in the wake. When vertical
vanes were attached atop the cylinder, roll amplitudes ranging from 30 to 150 degrees were obtained.
Itard (1999) presents buoyant ascent data from
a free swimming model where a trial run was suddenly stopped and ballast tanks blown (presumably
propulsion was also terminated). In the dead leaf
[flutter] ascent which followed, roll oscillations of
60 degrees were recorded before surfacing.
Binion and Stanewsky (1988) survey the literature on high incidence missile aerodynamics. They
consider Reynolds numbers just high enough to be
applicable to rising submarines. They note that
flow characteristics are a strong function of both incidence and Reynolds number. For missile bodies
(sharp noses, truncated tails) with length to diameter ratios similar to submarines, the wake is steady
with symmetric separation vortices at incidence angles up to 20 to 30 degrees, steady with asymmetric
vortices up to 60 or 70 degrees, and unsteady beyond that. The steady asymmetric vortex wake is
attributed to unavoidable, minor geometrical asymmetries in the pointed missile nose. An asymmetric
wake induces sway which would roll a submarine as
the sail resists the resulting lateral movement of the
hull.

Lloyd (1983) tries to do this in a more general


way. He models time dependence in trailing vorticity
by building a quasi-unsteady rational maneuvering
model in which vorticity trailing from all appendages
and the hull is convected past, and allowed to interact with, downstream structure.
The force generated by trailing sail vorticity is
substantial. This creates a problem because stability derivatives are generally acquired in steady state
tests in which the trailing vorticity is fully developed.
But prior to instability, during a buoyant ascent in
the vertical plane, there is minimal crossflow at the
sail. When instability does occur, the vorticity from
the resulting crossflow convects downstream only at
the nominal speed of the boat. Therefore, incipient
instability should be assessed by stability derivatives
that have the effects of trailing vorticity removed.
In his rolling moment stability analysis, Watt
(2001) removes the effect of the sail vortex/tail interaction from his stability derivative by basing it
3

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

on steady state model tests in which the tailplanes


are absent. It is possible to approximate unsteady
effects with information gleaned from steady state
experiments.

Boat

Mass (t) 20003000


Figures
1

10001500
2

10001500
36

Table 1 Trial submarine characteristics.


3

TRIALS DATA

The nature of the rising instability, which manifests


itself in operations as a roll instability, is demonstrated by full scale trials data from the three relatively modern diesel-electric submarines listed in
Table 1.
Figures 1 through 6 show the trials results. Quantities measured directly are u, , , , r , s , b , z0 .
Since y 0 could not be measured, it is assumed to
be zero and flow incidence is calculated from u,
z0 , and :
u
+ z0 tan
cos


z0
= tan1

x 0

is attained before enough flood water has drained to


restore stability. The boat then recovers, undergoing
a damped oscillation about zero roll angle.
Despite the care taken conducting this maneuver
(pitch up attitude prior to blowing, moderate blowing), an uncomfortably large roll angle still occurred
on the surface. The temporary hydrostatic instability a submarine experiences when surfacing is well
understood and normally not a problem; however,
the instability is aggravated by a large emergence
roll angle which, in Trial A, is the result of a submerged roll instability that is not understood.

x 0 =

(2)

Trials B, C4, C5, and C6 are all initiated from


depths of 80 or 90 m. Heading data are unavailable
for Trial B so rudder deflection is shown instead.
Unlike Trial A, Trials B and C use an emergency
blow capability that quickly empties the blown ballast tanks.

The axial velocity u is measured by an electromagnetic (EM) log located on the top of the hull on the
center plane. For zero roll, the local velocity at the
log is u. As roll progresses, the local velocity will be
somewhat larger because of interference from crossflow over the hull. By assuming the measurement is
always u, the calculation of errs on the low side.
In Figure 1, the largest boat is at depth when
it pitches up to the desired angle for the rising
maneuver. It then blows two of its four main ballast
tanks. The downward force from the sternplanes and
the forward momentum initially generate a positive
w velocity before the blow progresses enough that
buoyancy makes w negative. Although propeller
rpm is constant throughout the maneuver, the axial
velocity u increases because of the buoyancy component in the axial direction, and from the thrust resulting from the hull sailing in the crossflow. Flow
incidence increases as the blowing progresses but
tapers off after blowing stops. The air continues to
expand in the ballast tanks, but incidence is kept in
check by the increasing forward speed. The maneuver was nominally carried out in a vertical plane (the
rudder was fixed) but the heading still varies.
Roll angle is small through most of the rise,
until just after the ballast tanks empty. At this time,
the continually expanding air in the tanks begins
escaping and possibly interacts with the sail. When
the submarine rises above the surface, it temporarily
loses static stability until water has drained from
the sail and deck casing. A large surface roll angle

Trial B blows all ballast tanks simultaneously but


starts with a pitch down attitude ( = 10 degrees).
The vertical plane control surfaces are zeroed within
the first ten seconds of the maneuver and kept there.
The pitch up angle is achieved because the forward
ballast tanks have more volume than the aft ones.
Trials C blow only the forward tanks. Trials C1
to C3 all begin from an initial depth of 40 m and experience no roll instability, so only Trial C3 is shown.
Control surfaces are used to try to curtail the pitch
angle in this maneuver from t = 20 s onwards.
Similarly, Trials C4, C5, and C6 all use pitch angle curtailment. The curtailment is moderate in C4,
beginning at about t = 20 s. It is strong in the remaining two trials, beginning 10 to 15 seconds into
the maneuver.
The following observations can be made:
1) The incidence angles experienced in Trial A are
substantially lower than in Trials B and C. This is
attributed to the delayed and more moderate blowing in Trial A.
2) Because of the rapid changes taking place in Trials B and C, it is unclear at what incidence angle
stability is lost.
3) Pitch rate is nominally zero in Trial A but
changes rapidly in Trials B and C.
4

Copyright National Academy of Sciences. All rights reserved.

= 4%
U

Complete
Blowing
Begin
Blowing
Ballast

Figure 1 Boat A. Trial A.

30

20

10

0
10

20

20

40

30

60

40

80

50

10

10

20

20

40
z0

30

60

sensor limit

40
0

10

20
time (sec)

80

30

40

Figure 2 Boat B. Key for Figures 2 to 6.

20

10

10

10

20

20

40

30

60

40

80

10

20

20

40
0

10
20
time (sec)

10

20
30
time (sec)

100

40

Figure 5 Boat C. Trial C5.

, , , (degrees)

U,

20

100

50

z0

, , , (degrees)

40

30

0
30

30

U (m/s), (%)

20
time (sec)

z0 (m)

, , , (degrees)

20

10

Figure 4 Boat C. Trial C4.

U (m/s), (%)

z0 (m)

, , , r (degrees)

sail
broach

nose
broach

30

30

Figure 3 Boat C. Trial C3.


4) In Trials A, C5, and C6, expanding air begins
escaping from the ballast tanks at about the time
the roll instability begins. This is not the case for
Trials B, C3, or C4.
espe5) There is a correlation between and ,
cially in Trials A, B, and C6.

30

20

10

10

20

20

40

30

60

40

80

50

100

60
0

6) The sail for Boat A is somewhat further forward


than for Boats B and C.

10

20
30
time (sec)

40

Figure 6 Boat C. Trial C6.


5

Copyright National Academy of Sciences. All rights reserved.

50

120

U (m/s), (%)

z0 (m)

, , , (degrees)

Depth

Tanks
Empty

40

z0 (m)

0
Sail
Broaches

U (m/s), (%)

Twenty-Fifth Symposium on Naval Hydrodynamics

Twenty-Fifth Symposium on Naval Hydrodynamics

ways positive) goes through zero, these yaw sweeps


are equivalent to = 0 to 30 degrees for = 0
and 180, 30 and 150, 60 and 120, and 90 and
90 degrees. The solid lines are least squares fits to
the data of the anm coefficients in the quasi-Fourier
series representations of the vertical plane forces FV
(any of X, Z, or M ) and the horizontal plane forces
FH (Y , K, or N ):
FV0 = a00 cos2 +

m
X

a0m sinm

m=2
m even

Figure 7 STR generic model. R = 2023 million.

+ a11 cos sin cos

Conventional 6 DOF simulations of these maneuvers have been unable to reproduce any roll, yaw,
or sway instability, even with asymmetry added to
induce these motions when stability is lost.
4

n
X

n=1

m
X

(3a)

(3b)

anm sin cos n

m=2

0
= a11 cos sin sin
FH

MODELLING THE VISCOUS FORCES

n
X

n=1

The first step in analyzing the roll stability is to


model the hydrodynamic forces. The Gertler and
Hagen (1967) or Feldman (1979) models can be used
for this purpose. But these models have been developed for flow incidence angles up to only 18 degrees
or so. In this section, we present wind tunnel data
taken at incidence angles up to 30 degrees and show
that the conventional models break down in the 20
to 30 degree range where the forces are very nonlinear. We also present the models used in the stability
analysis of the next section
The wind tunnel data were acquired with the
Static Test Rig (STR) model test facility in the Institute for Aerospace Research 9 m wind tunnel in
Ottawa (Watt et al , 1991) using a 6 m long model.
Data were obtained on three support systems and
corrected for tare and interference effects, as described by Nguyen et al (1995). Figure 7 shows
the sting support, which was developed to obtain
roll sweep data at fixed incidence angles for the rising stability problem. This figure also shows the
generic model shape used in the experiments and
which the stability analysis in this paper is based
on. The model consists of an axisymmetric hull with
sail, symmetric + tail, and optional sailplanes and
deck.

m
X

m=2

anm sin sin n

where n and m vary from 2 to 5 and, of course,


the anm are different for each of the forces. The
dashed lines in Figure 8 show the conventional Feldman models for non-rotating flow and an axisymmetric hull:
X = X u2 + Xvv v 2 + Xww w2
p
Y = Yv uv + Yv|v| v v 2 + w2 + Yvw vw

Z = Z u2 + Zw uw + Z|w| u|w| + Zww |w


p
+ Zw|w| w v 2 + w2 + Zvv v 2

(4a)
(4b)
p

v 2 + w2 |
(4c)

K = Kv uv

M = M u2 + Mw uw + M|w|u|w| + Mww |w
p
+ Mw|w| w v 2 + w2 + Mvv v 2
p
N = Nv uv + Nv|v| v v 2 + w2 + Nvw vw

(4d)
p

v 2 + w2 |
(4e)
(4f )

These are fit to the data over the usual 18 degree test range. Except for the third terms in (4c)
and (4e), which result in unrealistic discontinuities
in slope, these terms are all included in (3). Watt
(2001) discusses how the standard rolling moment
model reduces to (4d) for the HST wind tunnel tests.
The HST data are useful for the four quadrant
perspective they supply. Figure 9 plots the HST fits
against for constant values ((3b) is odd in ,
so the = 0 to 90 degree region reflects the negative
region in Figure 8).
We show below that horizontal plane stability is
determined by the slopes through the origin of the
curves in Figure 9. These slopes are not well rep-

4.1 HST Yaw Sweep Data


The first STR data acquired were with the hull, sail,
and tail (HST) model configuration (no sailplanes
or deck). Yaw sweeps through 30 degrees were
carried out with the model rolled in 30 degree increments, as shown in Figure 8. Because flow orientation jumps by 180 degrees as the incidence (al6

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

0
2

30 20 10
6

Y 0 100

4
2
0
2
4
6

1
0
1
2
3

10 20 30

+
+
+
+
+

+
+ 
+ 
+ 
+ 
+ 
+
+

+

+



+

++
 ++
 +++


+
+

+
+
+
+
+
+
+




++
+
+

+++ 
++
+++


++
+
+
+

+
+
+
++



+

+

+
++
+
+


++
+
+

+












30 20 10
2
1

+
+
++
+
+
+
+
+

+

+


++



+
+
+


+
+
+
+

+

++
++

 ++
+++
++

30
60
90

1
2

30 20 10 0 10 20 30
degrees

2
1

150
120
90

++++
+
+
+
+
+
+
+
+

+



+

+
+
+
+


+
+


+
+
+


+


+++
+ ++
+
+
+
+
+
+
+
+






 

0
1





10 20 30

M 0 1000


+

+

+

+
 +

+

 +

 +
+  



+

+ 

+


+

+




+
+
+

+


Z 0 100

+
+
+ +
+
+
+

K 0 1000

X 0 1000

+
++
+
+
+
+
+

30 20 10
3
1
0
1
2

10 20 30

++++ 
++
++
++
+


+
+

+
+
 

 
+


N 0 1000


+

+


 ++
++
 +
+
+
++
++++

3
4

30 20 10 0 10 20 30
degrees

30 20 10 0 10 20 30
degrees

Figure 8 STR HST data; solid lines = fit with (3); dashed lines = fit with (4). The key applies to each plot.
2.5
2.0

K 0 1000

Y 0 100

24
18
12

30

1.5
1.0
0.5

N 0 1000

= 30

60 90 120 150 180


degrees

0.0

2
1
0
1
2

30

60 90 120 150 180


degrees

30

60 90 120 150 180


degrees

Figure 9 Cross plotting the HST fits from Figure 8; solid lines = fit with (3); dashed lines = fit with (4).
resented by the standard model for incidence angles
over 18 degrees. In fact, the model breaks down
at high incidence angles throughout the entire
range. Watt (2001) presents static rolling moment
data taken with the same generic model shape for
values up to 90 degrees (from a different facility).
Inherent unsteadiness aside, these data are even less
well represented by the standard model.

4.2

HSSDT Roll Sweep Data

This is not a criticism of the standard model


which was developed for use in a restricted incidence
range. Rather, it is warning that the standard model
cannot be reliably used for extrapolation. If a wider
incidence range is to be simulated, data must be
acquired over that range and the standard model
expanded to fit it.

The discontinuities in X and M at = 30 degrees are thought to be associated with tailplane


stall. The stalled X data are excluded from the fit.

For the rising stability problem, we concentrate on


the , = 0 to 30 degree range. STR roll sweep
data were acquired in this limited flow regime using
the sting roll mechanism of Figure 7 and the generic
HSSDT (hull, sail, sailplanes, deck, and tail) and
HSSD model configurations. The HSSDT data are
shown in Figure 10. These forces are fit with (3).

Two estimates of the error in these data are given


in Table 2. The standard error is the standard
deviation obtained by propagating RMS measurement and fitting error through the support strut
7

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

1.5

X 0 1000

1.0
0.5
0.0
0.5
1.0
1.5

4
4
4
4
4
4
4

4
4

44

4


44


44


444
44


44444444


+
44

+

+

+


+

+++

+

++

++

+
4
+
+++
44
++
+ + + + ++
+++++
44
4
4
44

444
444

444
44

44444444444
 

2.5
2.0
20

+ ++ + + +

+++++

++

1.5
1.0

 

++ +

+ + ++

+ ++ + + + + + + + +

10

0.5

0.0

4
444444
4444444444
44444444444444
  

      

+ ++ + + +
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+

10

20

30

3.0

30

Y 0 100

2.0

30

4
4
4
4
4
4
4

4
4

4

4

4

4

4



4
4
+


+
4
+

+



++
4
+


+
4


++
4
44
  +++
4 
44 
+
4 
+
44  
4
+

4

4  ++
44  
 ++
44  
4

++
44    + + + +
4
4 ++
+


+
44   + + +
4
+

+
4
+
44
4

4
+

4

4
+

4
+ +
 +
+
+
444
 
4

44

444

+
4

+ 4


+
444


4
+ 4
+4


+
++

4 4


+


4


4

4

4

+
4

+
+

4

+
+



4
+
+
4
+
4
+
+ + + + + ++ +
4
+
+
+
+
+

4

+
+
+
4
+

+

4
+
+


4


+
+


+


+
+

4
4
+

4
4


4
+

4
+

10

degrees

20

2.0

++ + + + + + + + + +

++ + +

 

++++++ +

++ + + +

++++

4444
44444444
444444
4444444444444

1.0

1.5

30



K 0 1000

Z 0 100

4
4 444444
4444
444
444
444


444


444

44
44


1.5

20

  
   

++ + + + + + + + + + +

+++++++++ ++++ +++ ++ +

444444444444444444444444444444 4

              


+ ++ + ++ + + + + + + + + + + + + + + + + + + + + + + + + +

0.0

1.0

0.5

10

0.5

0.0

10

30

20

30

4
4
4
4

4

4

4


4


4


4
++
4 

+
4 
++
4 
+
4 
+
4 
+
4 
+
4
4  ++
44
4  +
44
4
4  ++
4

4 +

44

4  ++
44


4

4 +

44   
4 + +
4
+


4

+


4
+ + 444    + + + + +
4
 +
4   ++ ++
4
 +++
 + + 44

4

+

44
4

++
4444
4 +44 
++
44444


+

+

+
44
+ +
4444          
4+

4

4
4

+

4
4





+44




+
+
+
+
4
4
4
4
4
4
+
+
4
4

4
+


+ ++ + + + + + + +
4

4

++
++

+
+
++


+
+
+
+
+


+
+

4
+

+
4
4

10

degrees

M 0 1000

0.4
0.6
0.8
1.0
1.2

0
44
4
+ + ++ + + +
4+ + + + +
+ ++ + +
+ + ++ + ++ + + + + + + +
   44
4


4


4
4



4
4


4



4
+


     4

4

4
4
4
4
4
44
4
++  
+4
+
+4
4444
4
4444
4
4
44
44444444
+ 4
+ 

444

+
+

+ 4 
4
+
+
+
 + + +
  44
+
4
++
++
++
+
+
4
+
+

++
 +4
+

+ ++

4
+
+ ++
4+ + + + +
 +


 +
4
 44
 +
4

+
4
   4
4
4
+
 
4
+  
4 
4

+ 

 4

4
++   4
4
++

44
+
4
4
44
+ +444

4
4444444
++++

1
10

N 0 1000

0.2

20

1.4

4
44 
44  +
4  +

4  +
+
 +
44
+

++
4

+
4

4 +
++ + +
+
4  ++

4
 +
+

44
4

++


+
4
+ + + + +

4
44
+
4

4
44

+

+
4
+

44

+

4
+

4
+ 444

+

       
4
+ 4
+


+
4
+ 44
 

+

4

+ 4    
+
4
+

+44

 
+4
4

+
+ 4
+4
    
4
4


+
+

4

+
4

4

4
+
+


4+

 + ++++ ++

+
4
+
+
4
+
4
4
+ +
4


4
4
+

4
+
 + +

+
+
+
4
+ + + + + + +
4
+ ++

+

+


+

+
4
+
4

+

+



4
4



 44

4
+

4
4
44
4
4
4
4
4
44
444
4 44444
4

30

2
10

10

30

20

20

degrees

0.2
0.0

degrees

10

30

2.5

2.0

20

20

30

10

degrees

20
degrees

Figure 10 HSSDT roll sweep data and fits; , = 0 to 30 degrees with increments of 2 degrees.
8

Copyright National Academy of Sciences. All rights reserved.

30

20
10

0
30

Twenty-Fifth Symposium on Naval Hydrodynamics

Coefficient
1000

Standard
Error

Error
Bound

X0

0.07

0.3

Y 0, Z0

0.41

0.8

0.03

0.13

0.12

0.2

M ,N

0.9
Cd 0.8
0.7
0

Figure 11

Table 2 STR error.

= Cd

xn b(x)w(x)|w(x)| dx + O 2
`

(7)

where b(x) is the local breadth of the hull, n = 0


for Z, and n = 1 for M . Again, these integrals each
can be written as the sum of two terms constructed
from the three integrals:
Z
Bm (w, q) = xm b(x)|w xq| dx m = 0, 1, 2. (8)
`

The drag coefficient Cd is obtained from the


HST Z force fit of Figure 8. The terms odd in w
are extracted
by taking the difference Z(, )

Z(, 0) /2 and then dropping the cos sin term
(ie, the uw term):
Cd () = c0 + c1 sin + c2 sin2

 `2
Az

(9)

R
where Az = ` b(x) dx. Shown in Figure 11, this
function models the variation in the drag coefficient
with incidence, though the variation is not large at
moderate to high incidence angles where it matters.
The ZHST function is used since it was fit to two
quadrant data, unlike the HSSDT data. Thus, its
coefficients for even and odd functions in w should
be well resolved. In any case, Az is exactly the same
for the HST and HSSDT configurations.
The addition of (5) and (7) to the equations of
motion duplicates Cd ()v|w| and Cd ()w|w| terms
already present in the Figure 10 fits. These functions
must be subtracted from the fits to ensure consistency at q, r = 0. This is done in the modified Y, N
forces Ymod , Nmod in 4.6 below.

xn h(x)v(x)|w(x)| dx + O 3

Cd () is used to model crossflow drag.

These terms are just parameters in the horizontal


plane stability analysis.
The vertical plane equations use slightly different
integrals for the Z and M forces:
Z
p
Cd xn b(x)w(x) v(x)2 + w(x)2 dx

4.3 Modelling Rotation


The above models apply only to translation. To
model rotation, unfortunately, we must make do
with either estimating the linear derivatives or, at
best, extracting them from conventional, small amplitude oscillatory experiments carried out at small
incidence angles. The impact of incidence on the rotational derivatives may be as large as on the translational forces, but currently we have no way of accounting for this.
For the stability analysis, linearized p, r models
are all that is required.
Pitch rate q is an important state to model for
Trials 2 through 6. Adding nonlinear rotation terms
to low quality linear derivatives may not add accuracy, but doing so at least shows how it can be done
and allows the sensitivity of the stability analysis to
these nonlinear terms to be evaluated.
Feldmans standard model generates second order rotation terms in the Y, Z, M, N forces by accounting for crossflow drag by integrating the local
crossflow velocity over the length of the hull. For our
stability analysis, these forces are linearized in v and
r. The integrals for the horizontal plane equations
become:
Z
p
Cd xn h(x)v(x) v(x)2 + w(x)2 dx
= Cd

30

n, two of which are identical, so only three integrals


are needed:
Z
Hm (w, q) = xm h(x)|w xq| dx m = 0, 1, 2. (6)

tare and interference correction procedure. The error bound is a simple summation of the magnitudes
of all sources of error in a single measurement. In all
cases, the RMS error of the fits shown in Figure 10
is less than the standard error in Table 2.

10
20
degrees

(5)

where Cd is a drag coefficient usually extracted from


the Zw|w| derivative, h(x) is the local height of the
hull, w(x) = w xq is the local normal velocity of
the hull, v(x) = v + xr is its local lateral velocity,
n = 0 for Y , and n = 1 for N . The v and r variables can be brought outside the integral allowing
(5) to be written as the sum of two terms for each
9

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

100

0.4

F (, )/
FHSSDT (, )/ =0

Extrapolated Feldman N
model, Boat B

0.2

0

Nopf

=0

0.6

0
Yopf

=0

100

STR Y data

N0
NHSSDT

Extrapolated Feldman Y
model, Boat B

10

20

10
20
degrees

30

Figure 13 Normalized STR generic model stability derivatives. The N derivatives are separated
to avoid a singularity; they are each normalized by
|NHSSDT /| at = 30.

30

degrees

Figure 12 The Y and N out-of-plane force contribution to the steady state stability derivatives for
Boat B and the generic STR hull form.
4.4

F = KHSSD

3
0

F = Y0

STR N data

When the ratio N/Y is plotted using (10), the effective point of application of the out-of-plane force
is seen to be x = 0.2` at = 0 and to gradually move forward through x = 0 at = 22 to 25
degrees, the location of the zero in Nopf . This is consistent with our understanding that vorticity trailing
from the sail progressively separates from the afterbody, from the tail first and moving forward, with
increasing incidence.
The deck on the HSSDT model is not large and
so the out-of-plane force contribution to rolling moment is small and is ignored. However, the trailing
sail vorticity does interact with the tail. As previously noted, the tail interaction is subtracted out
of the rolling moment stability derivatives by basing
them on HSSD steady state rolling moment data.
Figure 13 shows how the stability derivatives
change when the tail is removed and the out-of-plane
force is subtracted to create the F0 forces. There is
a reduction of about 30% to the N derivative at
moderate incidence angles, a common state during
buoyant ascent. The change to the K derivative
is initially large but falls off quickly as the trailing
vortex separates from the hull and tail. The change
to the Y derivative is largest at moderate incidence
angles. These changes would be much larger if they
were based on the standard model. The change in
K, for example, would result in the 80% increase at
= 0 remaining constant at all incidence angles.
The zero crossing in N0 / at 25 degrees
is, we show below, the source of an instability that
moves to lower values under certain conditions.

Quasi-Unsteady Effects

The objective of the stability analysis is to predict


incipient instability. The boat is assumed to be ascending upright in the vertical plane before instability occurs. Any crossflow at the sail prior to instability, and its associated trailing vorticity, is neglected. The stability analysis therefore disregards
the out-of-plane force contribution to the stability
derivatives. This is accomplished by subtracting the
out-of-plane force components from the steady state
(3b) models. This is equivalent to zeroing the CL
and Ki coefficients in Feldmans equations of motion.
To the best of our knowledge, and certainly in
Feldmans model, the out-of-plane force is unique in
being odd in both v and w. This force is the vw
term in a conventional second order Taylor series
expansion of the steady state translational forces. It
is not well resolved by the HSSDT single quadrant
data so it is obtained from HST data instead. The
terms odd in v and w in (3b) are the sin 2n terms:
0
= (0.1289 sin2 + 0.2018 sin3 ) sin 2
Yopf

+ (0.0287 sin2 + 0.0851 sin3 ) sin 4 (10a)


0
Nopf
= (0.02949 sin2 0.06377 sin3 ) sin 2

+ (0.00327 sin2 0.01561 sin3 ) sin 4 (10b)

Note the large sin3 coefficients. The effects of (10)


on the stability derivatives are shown by the blue
curves in Figure 12. For comparison, the nonlinear
Feldman model, implemented for Boat B based on
data acquired over incidence angles of 10 degrees,
is also shown. Clearly, nonlinear effects at incidence
angles over 15 degrees are important.

4.5 Full Unsteady Effects


In this subsection we speculate about unsteady effects not accounted for in the current model. In Figures 2 through 6, the time taken for vorticity generated at the sail to convect to the tail is about six
10

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

seconds. If the boat is oscillating in roll with a half


period of six seconds or so, as is approximately the
case in Figures 2, 5, and 6, then there is a cyclic
impulse on the tail delivered by the vorticity generated by the sail a half cycle previously. In a steady
flow, the afterbody vortex interacts with the tail to
moderate the rolling moment at the sail. But if the
boat is already rolling in the opposite direction when
the trailing vorticity reaches the tail, the interaction
would amplify the roll. This is resonance.
The argument against this is the small effect of
the trailing vortex/tail interaction on K/ in
Figure 13 at the incidence angles seen in the trials during the oscillations (if that is what they are).
However, the drop off in K/ with results from
steady state separation which may not have time to
fully develop during a roll oscillation.
In fact, all of the arguments of 4.4 assume
that fully developed flow structures occur instantaneously, which is incorrect. But this is the best we
can do at this time.

4.7

The , angles are a convenient way to conceptualize the forces for the rising stability problem. They
simplify roll sweep descriptions when acquiring test
data. They allow Fourier analysis to be used as a
guide for choosing high order fitting functions and
for ensuring that the flow description for a given
harmonic is complete. And they simplify the calculation of roll stability derivatives since is linear in
the perturbation variables while is independent
of them (within O(2 )).
The downside to using this flow description is
added complexity when sweeping through zero incidence, as illustrated in Figure 8 by the discontinuity in that results. When choosing Fourier terms
for fitting data, one must be careful that unnatural discontinuities in the flow characteristics are not
introduced by the discontinuity.
5

4.6 Viscous Forces Summarized


If the modified Y, N forces are:

Ymod = YHSSDT (, ) + Cd ()vH0 (w, 0)


2

sin 2n terms in YHST (, ) (11a)

Nmod = NHSSDT (, ) + Cd ()vH1 (w, 0)


2

sin 2n terms in NHST (, ) (11b)

5.1

(12a)
(12b)

Nvisc = Nmod (, ) + Nr ur + Np up


Cd () vH1 (w, q) + rH2 (w, q) (12c)


2
The vertical plane forces are independent of v, r, p
to within O(2 ):
Xvisc = XHSSDT (, ) + Xqq q

STABILITY ANALYSIS
Theory

Consider the stability of a submarine rising upright


( = 0) in a fixed vertical plane (y0 , = 0) with
arbitrary horizontal and vertical velocities ( x 0 , z0 ),
pitch angle ( ), and time derivatives of same. Stability is determined by perturbing y0 , , in the equations of motion to see if the perturbations grow or
die out. But these equations use body axes velocities so we first examine the impact of the above
assumptions on the kinematic relations between the
inertial coordinates and Euler angles and the body
axes velocities. That is, we linearize Feldmans auxiliary equations about y0 , , = 0, assuming the
time derivatives in these quantities are also small:

then the viscous horizontal plane forces are:

Yvisc = Ymod (, ) + Yr ur + Yp up


Cd () vH0 (w, q) + rH1 (w, q)


2
Kvisc = KHSSD (, ) + Kr ur + Kp up

Rising Stability Angles

u = x 0 cos z0 sin

(14a)

v = x 0 ( sin ) + y 0 + z0 cos

(14b)

w = x 0 sin + z0 cos

(14c)

p = sin

(14d)

= w x 0 + y 0

(13a)

Zvisc = ZHSSDT (, ) + Zq uq Cd () wB0 (w, q)


2

qB1 (w, q) wB0 (w, 0)
(13b)


Mvisc = MHSSDT (, )+Mq uq+ Cd () wB1 (w, q)


2

qB2 (w, q) wB1 (w, 0)
(13c)


q =

(14e)

r = cos

(14f )

all to within O(2 ). As expected, v, p, r are also


small quantities. More importantly, the vertical
plane motions u, w, q are, to first order, independent of the perturbations. The vertical plane equations of motion, the X, Z, M equations listed below, are also independent of the perturbations. This

An adjustment for scale effects is also made to the


a00 term in (3a) for X.

11

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

terms estimated from potential flow theory. The


acceleration (added mass) derivative estimates are
surprisingly good. And the rotation coefficient estimates, which potential flow theory predicts exist
(Imlay 1961, Watt 1988) but for which there is no
contribution in the viscous model, are the best available. Some of these potential flow terms are small
enough to be neglected.

means that u, w, q do not respond to the perturbations and therefore can be treated as merely parameters in the stability analysis. Since only the
horizontal plane equations take an active role in the
analysis, we call it a 3 DOF analysis. However, since
the vertical plane equations have been accounted for
and are used below to determine vertical plane velocities, this could be considered a 6 DOF analysis.
The linearized horizontal plane equations of motion are:

m v wp + ur + zG (qr p)
+ xG (qp + r)

The boat is also assumed to have good trim before


the ascent begins. That is, initially W = B and
xG = xB . After ballast blowing is initiated:

= Yv v + Yp p + Yr r + Ywp wp + Ypq pq
+ Yvisc (B W ) cos

B W = B

(15a)

= Kv v + Kp p + Kr r + Kwp wp + Kpq pq
+ Kvq vq + Kwr wr + Kqr qr
+ Kvisc + (zB B zG W ) cos

(15b)

= Nv v + Np p + Nr r + Nwp wp + Npq pq

Similarly, it is usual in horizontal plane stability


analyses to use v, , r as the perturbing quantities
because does not appear explicitly in the equations of motion (p and are eliminated from the
equations using (14d and f)). This simplifies the
problem even further, but ignores the influence
on v in (14b). We include the dependence to
maintain generality and this gives rise to stability
limits not previously seen (see Appendix).

(15c)

The linearized vertical plane equations contain no


perturbation quantities at all:

m u + wq xG q 2 + zG q
= Xu u + Xw w + Xq q + Xq uq + Xwq wq
+ Xvisc + Xs s u2 s2 + Xb b u2 b2
+ XP + (B W ) sin

Thus, the v, p, r variables in (15) are replaced


with (14b, d, and f), and their time derivatives with:

(16a)

v = w
+ w x0 x 0 + y0

cos
p = sin q

= Zu u + Zw w + Zq q

Iy q + m zG (u + wq) xG (w uq)
= Mu u + Mw w + Mq q

(16b)

(18a)
(18b)

sin q
q
r = cos q

+ Zvisc + ZS u2 s + Zb u2 b
(B W ) cos

(17)

Because the y0 perturbation does not appear in


(14) or in the equations of motion, it can be replaced
with y 0 with no loss of generality. This simplifies the
characteristic equation of the dynamical system by
reducing its order by one.

Iz r + (Iy Ix )pq + Izx (rq p)


+ mxG (v wp + ur)
+ Nvisc (xB B xG W ) cos

xB x
1

The mass m and the moment of inertia I dependence on and x is accounted for. Any vertical
or lateral displacements of the blown mass centroid
are ignored.

Ix p + (Iz Iy )qr Izx (r + pq) mzG (v wp + ur)

m w uq zG q 2 xG q

and xG =

(18c)

The forces in (12) are linearized using:

F (, ) =

+ Mvisc + MS u2 s + Mb u2 b

x 0
y
+ 0
w
w

(19)

since F (, 0) = 0, F/ (, 0) = 0, and =
v/w + O(3 ) (see (3b), (1), and (14b)).

+ (xB B xG W ) cos + (zB B zG W ) sin (16c)

Equations (18) and (19) contain x 0 , x0 , w,


, q
and q parameters, some of which are not normally
considered in horizontal plane stability analyses. For
trials analysis, values for these parameters are available by analyzing the trial measurements. For the

These equations assume that both the boat mass


distribution and its hydrodynamics are symmetrical
about the vertical plane of symmetry. The second
line of each equation (and third for (15b)) contain
12

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

generic analysis that follows, the x 0 term is written in terms of u, w, and using (14a and c). This
expression is differentiated in time to give x0 as a
function of u, w, , q, u,
and w.
Then, u and w are
obtained from the axial and normal force equations
of motion with Xw , Xq , s , b = 0.

5.2 Exact Numerical Solution


The maximum of the real parts of the five roots of
(21) are shown in Figure 14 for Trial A parameters and appropriately scaled generic model hydrodynamics. These numerically obtained surfaces are
exact solutions of (21) for the complete model presented above. The surface is truncated below zero so
that the white area indicates stability. Discontinu
ities appear in the surfaces when max <(s) shifts
focus from one root to another. There are three regions of interest in Figure 14a.
First, the boat is everywhere unstable near zero
speed. This is a buoyancy instability that results
from terms linear in B in A2 , A3 , A4 and quadratic
in B in A5 . It disappears as or go to zero, or increases as they get larger. It occurs at low incidence
angles even though all the Ai > 0 here.
The second region of interest is the large bulge
in the outer corner of the quadrant. This is the
rolling moment/ instability Watt (2001) initially
examined looking for the source of the buoyant ascent instability. It disappears if the rolling moment
equation and perturbation are eliminated from the
stability analysis (ie, the middle row and column in
(20) are discarded). All Ai > 0 for the low sloping
part of the bulge closest to the origin. The discontinuity in the bulge has not been closely investigated
but seems to occur as the A3 , A4 coefficients go negative. As the static stability BG of the boat is reduced, the bulge moves towards the origin.
The third region of interest in Figure 14a is
the (effectively) speed independent stability limit at
= 25 degrees. This limit is determined by a sign
change in A5 , which is convenient since A5 is one of
the simplest of the Ai coefficients. In the next subsection, we show that the basis for this limit is the
zero in N0 /.
The effects of q, q,
and x on stability are examined in the remaining parts of the Figure. Positive
q, q values result in a receding A5 stability limit and
movement of the instability towards the origin. We
show in the next subsection that this latter effect is
approximately accounted for by modelling the effect
q has on the local incidence at the sail, making it
dependent on the axial location of the sail.
Figure 14b shows an additional feature near zero
incidence at high speed. This results from, again, a
change in sign of A5 . As shown below, A5 is proportional to x0 :

This leaves u, w, , q, q.
We solve for stability in
U, space which is based on the u, w parameters.
The solutions are a function of the parameters , q, q
and the blown mass parameters , x . Appropriate
values are chosen based on the trials results.
We also assume self-propulsion at zero incidence
for all speeds. That is, XP = XHSSDT (0, 0).

With (15) linearized in the , we set  = 0 est


so that  = s0 est and  = s2 0 est , factor out 0 est
and get the characteristic equation for the system as
the determinant:

Yy0 s + Yy 0

Ky s + Ky
0
0
N s+N
y0
y 0

Ys2 + Y s + Y
Ks2 + K s + K
Ns2 + N s + N

Ys2 + Y s + Y

Ks2 + K s + K
Ns2 + N s + N

(20)

where the F are obtained from (15). The real parts


of the roots of the characteristic equation, a quintic
polynomial in s:
A0 s5 + A1 s4 + A2 s3 + A3 s2 + A4 s + A5 = 0 (21)
must be negative for the system to be stable. A necessary condition for stability is that all the Ai be
positive, but this is not sufficient. Rouths stability
criterion (Ogata, 1997) is a procedure for constructing additional expressions from the Ai that must
also be positive and, if so, guarantee stability. However, it is simplest to just solve (21) numerically and
directly examine the real parts of the five roots for
each case. This gives a good feeling for which criteria are important and analytical solutions of the
important criteria can then be obtained by making
suitable approximations.
The analysis is carried out symbolically using
mathematical software1 which also does numerical
analysis and plotting. This avoids mistakes in what
would otherwise be a tedious solution.

x0 = (u + wq) cos + (w uq) sin


1

(22)

which changes sign at relatively high uq values. High


pitch rates at high speed should be avoided.

www.maplesoft.com

13

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

a) q 0, q0 = 0
= 0.04
x = 0.05

d) q 0 = 0.5, q0 = 0
= 0.04
x = 0.05
2

max(<(s))

max(<(s))

2
Unstable

1
0

25
20
15
10

Stable

5
0

14

12

10

25
20
15
10

6
U m/s

max(<(s))

max(<(s))

12

10

6
U m/s

1
0

1
0

5
0 14

12

10

25
20
15
10

6
U m/s

5
0 14

c) q 0 = 0, q0 = 2
= 0.04
x = 0.05

12

10

6
U m/s

f ) q 0 = 0, q0 = 0
= 0.055
x = 0.35
2

max(<(s))

2
1
0
25
20
15
10

e) q 0 = 0, q0 = 2
= 0.04
x = 0.05

25
20
15
10

0 14

b) q 0 = 0.5, q0 = 0
= 0.04
x = 0.05

max(<(s))

1
0

5
0 14

12

10

25
20
15
10

6
U m/s

5
0 14

12

10

6
U m/s

Figure 14 Stability calculations based on Trial A parameters and Trial B and C pitch rates. The red lines
are the zeros in the simplified A5 coefficient. The blue lines are from a simplified 1 DOF rolling moment
analysis in which local flow incidence at the sail as a function of q is accounted for.
14

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Negative q, q values result in a receding instability and strong movement of the A5 instability towards = 0, especially at high speeds.
The Figure 14 calculations use = 0.04, except
for Figure 14f. When is doubled in the baseline
calculation of Figure 14a, there is a slight filling out
of the bulge and the A5 stability limit decreases to
about = 23 degrees. This is only a minor effect.
However, we show below that what really matters is
an increase in the product x , as shown in Figure
14f where and x values from Trials C are used.
Only the forward ballast tanks are blown in Trials
C so the blown mass centroid x is relatively far
forward. When this happens, the A5 stability limit
moves much closer to the origin.
Figure 14f provides an explanation for the instabilities in Trials C4, C5, and C6, all of which occur at
relatively high incidence, large negative pitch rates,
and with a large x value. However, the instabilities in Trials A and B remain unexplained. The
discrepancy in Trial A, for example, is equivalent to
either a 50% error in the BG value or an error half
as large in ship speed, which is consistent since the
hydrodynamic forces go as the square of the speed.

This gives:
A0 (m0 Yv0 )(Ix0 Kp0 )(Iz0 Nr0 ) cos

A5 x00 C0 + Cq q 0 + Cq q0

C0 = B 0 cos

Cq =


0
1 Nmod
0
+
+ Cd H1 (w, q)
(m
w0


0
1 Ymod
0
0
+ x m0
+ Cd H0 (w, q)
(27b)
w0

I y = Iz ,

Mq = Nr

Yv0 )

Cq = (m0 Yv0 )(Iz0 Nr0 )

0
1 KHSSD
w0

(27c)

and it is understood the derivatives are evaluated at


= 0. Since w = sin , the F//(w) terms
in (27) are positive quantities that are nonzero and
nonsingular as 0. The vertical scales in Figure
14 result from normalizing (21) by A0 .

(24)

The zero in the bracketed expression in (26b) is


plotted as red lines in Figure 14. So far, this expression has always provided a conservative estimate of
the exact zero in A5 .

to account for pitch rate effects that the steady state


experiments producing KHSSD could not measure.
This model is shown in Figure 14 as the blue line.
The model does not account for q so, in Figures 14c
and e, only the basic q = 0 model is shown (for
reference, as a dashed line).
A necessary requirement for stability is that the
A5 coefficient in (21) be greater than zero. A manageable and reasonably accurate analytic expression
approximating A5 is obtained with some assumptions. We neglect all the off-diagonal added mass
matrix terms, terms O(2 ), and set:

zB = BG,


0

1 KHSSD
(m0 Yv0 ) Nr0 u0 +Cd H20 (w, q)
0
w



+ x m00 (Yv0 Yr0 )u0 + Cd H10 (w, q)

Kr0 u0

=0

xB , xG , zG , Izx = 0



0
BG 0
1 Nmod
0
(m Yv0 )
+
C
H
(w,
q)
d 1
`
w0



0
BG
1 Ymod
0
+
C
H
(w,
q)
+ x m00
d 0
`
w0

0
1 KHSSD
(27a)
(m0 Yv0 )
w0

for stability in . The HSSD rolling moment is determined almost entirely by the sail so it is reasonable
to replace in KHSSD / in (23) with:
(w xF W q)
u + zF W q

(26b)

where:

5.3 Simplified Stability Limits


A 1 DOF stability analysis (Watt, 2001) using just
the rolling moment equation of motion with as the
perturbation parameter, results in the requirement:
v
u
BGB cos
u

U <u
(23)
0

t 1 3 KHSSD

`
2

F W = tan1

(26a)

For normal blowing of all main ballast tanks simultaneously, the x terms in (27) can probably
be neglected, greatly simplifying this expression. In
addition, when q, q = 0, the only term left is the
second line of (27a). Thus, the zero in N0 / in
Figure 13 is the basis of the zero in A5 .
The effect of pitch angle on the stability limit is
generally small, as seen in the cos terms
in (23) and
(26b). The limit in (23) reduces as cos which is
only a 15% reduction at = 45 relative to = 0
degrees. There is no effect on (26b) if q, q = 0. But
the effect on (22) is more interesting; it shows that

(25)
15

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

the tendency of x0 , and therefore A5 , to become


negative increases as uq sin .

REFERENCES
Binion, T.W., and Stanewsky, E., Observed Reynolds
Number Effects: Low Aspect Ratio Wings and Bodies,
Reynolds Number Effects in Transonic Flow, AGARD
AG303, December 1988.

The A5 stability limit is a new limit that results


from the use of as the perturbation parameter.
6

Booth, T.B., Stability of Buoyant Underwater Vehicles


Part I: Predominantly Forward Motion, International
Ship Building Progress, vol. 24, no. 279, November 1977.

DISCUSSION

The above analysis shows that, during a buoyant


ascent, horizontal plane stability is compromised by:

high incidence, especially at high speed,

both positive and negative q, q values.

blowing only the forward ballast tanks.

Booth, T.B., Stability of Buoyant Underwater Vehicles


Part II: Near Vertical Ascent, International Ship Building Progress, vol. 24, no. 280, December 1977.
Feldman, J.P., State-of-the-Art for Predicting the Hydrodynamic Characteristics of Submarines, Proceedings, Control Theory and Naval Applications Symposium, Naval Post-Graduate School, Monterey, July 1975.

Buoyancy does not alter the stability limits greatly


if ballast is blown evenly. Whether buoyancy compromises stability by moving the boat closer to these
limits is an issue yet to be addressed.

Feldman, J., DTNSRDC Revised Standard Submarine


Equations of Motion, DTNSRDC/SPD-0393-09, June
1979.

Trials results suggest that the actual region of


stability is smaller than the theory suggests. This
makes it difficult to make detailed recommendations
on avoiding instability. There are three areas for
future work:

Gertler, M. and Hagen, G.R., Standard Equations


of Motion for Submarine Simulation, NSRDC Report
2510, June 1967.
Haarhoff, S. and Sharma, S.D., A Note on the Influence
of Speed and Metacentric Height on the Yaw-rate Stability of Displacement Ships, Proceedings of the CPMCWorkshop on Ship Manoeuvrability at HSVA, Hamburg,
October 2000.

1) Continue to try to explain the events in Trials A


and B, through analysis and by gathering more
information on the trials. Does expanding air escaping from the ballast tanks as the boat rises
compromise stability? Are there other factors?
Conduct new trials with different boats.

Hoyt, E.D. and Imlay, F.H., The Influence of Metacentric Stability on the Dynamic Longitudinal Stability of
a Submarine, DTMB Report C-158, October 1948.

2) Vertical plane simulations can be carried out to


investigate strategies for avoiding high incidence
and, presumably, instability. Some of these have
already been done, with mixed results.

Imlay, F.H., The Complete Expressions for Added


Mass of a Rigid Body Moving in an Ideal Fluid, David
Taylor Model Basin, Report 1528, July 1961.
Itard, X., Recovery Procedure in Case of Flooding,
RINA Warship 99 International Symposium on Naval
Submarines 6, paper 24, London, June 1999.

3) Develop an unsteady simulation capability with


a CFD code to investigate the true effects of unsteadiness. This work is underway.
7

Lambert, J.D., The Effect of Changes in the Stability Derivatives on the Dynamic Behavior of a Torpedo,
ARC R&M no. 3143, March 1956.

CONCLUDING REMARKS

Lloyd, A.R.J.M., Progress Towards a Rational Method


of Predicting Submarine Manoeuvres, RINA International Symposium on Naval Submarines, paper 21, London, May 1983.

A 3 DOF horizontal plane stability analysis has been


carried out to investigate the modes of instability
a submarine in buoyant ascent is subject to. The
analysis confirms the existence of a instability discovered earlier. It also identifies several other modes
of instability which explain some instabilities seen in
full scale trials. High incidence force measurements
at unconventional flow orientations play an important role in establishing stability derivatives.

Nguyen, V.D., Drolet, Y., and Watt, G.D., Interference


of Various Support Strut Configurations in Wind Tunnel
Tests on a Model Submarine, 33rd Aerospace Sciences
Meeting and Exhibit, AIAA paper 950443, Reno, January 1995.
Ogata, K., Modern Control Engineering, 3rd edition,
Prentice Hall, 1997.

Not all the trial instabilities are explained by the


analysis. More work is required to identify sources
of error and expand our understanding of unsteady
effects during roll oscillations.

Papoulias, F.A., and McKinley, B.D., Inverted Pendulum Stabilization of Submarines in Free Positive Buoyancy Ascent, Journal of Ship Research, vol. 38, no. 1,
pp. 7182, March 1994.

16

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

New Analysis

Tinker, S.J., Fluid Memory Effects on the Trajectory


of a Submersible, International Ship Building Progress,
vol. 25, no. 290, 1978.

Perturbing in y0 and :

Watt, G.D., Estimates for the Added Mass of a MultiComponent, Deeply Submerged Vehicle, Defence Research Establishment Atlantic TM 88/213, October
1988.

v = y 0 x 0

v = y0 x0 x 0

r =

r =

(32)

Watt, G.D., Tanguay, B., and Cooper, K.R., Submarine


Hydrodynamics in the Wind Tunnel: The DREA Static
Test Rig, RINA Warship 91 International Symposium
on Naval Submarines 3, paper 19, London, May 1991.

instead of v and r gives the system an extra degreeof-freedom, resulting in a cubic characteristic equation:
A000 s3 + A001 s2 + A002 s + A003
(33)

Watt, G.D., A Quasi-Steady Evaluation of Submarine


Rising Stability, RTO-AVT Symposium on Advanced
Flow Management, Loen, Norway, May 2001.

The A000 , A001 , A002 coefficients here are identical to the


A00 , A01 , A02 coefficients in (30). So the requirement
that A002 > 0 reproduces the conventional limit (31).
But there are now two additional limits on stability.

Wichers Schreur, B.G.J., The Motion of Buoyant Bodies, Ph.D. Thesis, Trinity College, Cambridge, September 1990.

The first additional limit results from the A003 > 0


requirement. This limit is analogous to that derived
from the A5 > 0 requirement in the main text:

APPENDIX: Comparing the Conventional and


New Horizontal Plane Stability Analyses

We compare simplified conventional and new horizontal plane stability analyses to get a physical feeling for the meaning of the new stability limit. Conventional Gertler and Hagen (1967) hydrodynamic
coefficients are used in the equations of motion, the
rolling moment equation is ignored, and we set:
w, q, p, = 0

(34)

(35)

If x0 < 0, a condition on Nv /Yv contradictory to


(35) results:

m(v + ur + xG r)
= Yv v + Yr r + Yv uv + Yr ur



Nv mxG Nr mxG
Nv
< min
,
Yv
Yv m
Yr m

Iz r + mxG (v + ur) = Nv v + Nr r + Nv uv + Nr ur
(29)

(36)

It is impossible to simultaneously satisfy both (35)


and (36).
If x0 = 0, then (33) reduces to (30) and an interesting question (not answered here) arises: How
large an x0 magnitude is required for (34) to be of
practical concern?
Note that in many analyses, xG and off-diagonal
added mass matrix terms such as Nv are neglected.
The limit (34) then simplifies to Nv x0 < 0 since
Yv < 0 always.

Conventional Analysis
Consider the effect of perturbations in v and r by
setting v, r est . The characteristic equation is:


(m Yv )s Yv u
(mxG Yr )s + (m Yr )u



(mxG Nv )s Nv u (Iz Nr )s + (mxG Nr )u

(30)

A01

For stability, s must be negative.


and
are
always positive so s < 0 if A02 > 0, which gives:
Nv
N mxG
< r
Yv
Yr m

x0 > 0

Nv mxG
N
N mxG
< v < r
Yv m
Yv
Yr m

The lateral force and yawing moment equations then


reduce to:

A00

If x0 > 0, as is usually the case for buoyant ascents,


then the A002 and A003 limits require that Nv /Yv fall
within a fixed range:

(28)

= A00 s2 + A01 s + A02

Nv
mxG Nv

Yv
m Yv

The second additional stability limit results from


the application of Rouths criterion to the cubic
characteristic equation (33). This expression is algebraically complicated and is not reproduced here.
However, it cannot be ignored and in some cases will
be the critical condition determining stability.

(31)

That is, conventional stability is achieved when the


center of application of the lateral force resulting
from lateral velocity is aft of the center of application of lateral force resulting from yaw rotation and
minor inertial effects.
17

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Jerome P. Feldman
Naval Surface Warfare Center, Carderock Division,
USA
This paper provides an improved
understanding of the dynamic stability of a submarine
during buoyant ascent. It also provides a more
accurate method of predicting the motions of the
submarine performing this type of maneuver, as well
as a better understanding of the complex
hydrodynamic phenomena.
The hydrodynamic coefficient mathematical
model described in Reference (1) is now used at
NSWCCD primarily for performing trade-off studies
of candidate submarine designs, and then only to
predict mild maneuvers. A multi-vortex computer
simulation code is the current method for both
estimating the hydrodynamic forces and moments on
the submarine hull, appendages, and propulsor and
for performing computer simulations of the motions
of the submarine. The code is a physics-based, semiempirical method that uses two-dimensional potential
flow strip theory, discrete vortex representation of the
flow separation from the hull, and lifting line theory
for the appendages to determine all of the
hydrodynamic forces and moments except for the
longitudinal force. The trajectories of the vortices
shed from the hull and appendages are calculated, as
are the induced forces on the hull and appendages
that are downstream of the vortices.
Experimental data are required to develop an
accurate mathematical model of the submarine to
predict buoyant ascent. As indicated in Reference (2),
NSWCCD performs captive-model experiments to
measure the hydrodynamic forces and moments on
the entire submarine, as well as the forces and
moments on all of the appendages and the propeller.
The experiments are performed in the NSWCCD
straight-line basin using a yaw table. These tests are
performed over a range of angles of attack from -180
to +180 degrees and angles of drift from -20 to 180
degrees. Rotating arm experiments are performed
with a strut-supported model over a range of
nondimensional yawing angular velocities, angles of
drift, rudder angles, and propeller rpm's. Additional
rotating arm experiments are performed with a stingsupported model, but without a propeller, over a
range of combined rolling, pitching, and yawing
angular velocities and combined angles of attack and
angles of drift.
Almost all of the experiments that
NSWCCD has performed at large angles of attack or
drift have been with a fully appended model.
However, it appears that the stern appendages have a

significant effect on the shape of the force and


moment curves at an angle of attack of about 40
degrees. NSWCCD is now considering performing
these experiments with the unappended hull, as well
as fully appended to improve the prediction of the
coefficients at large angles.
For three-dimensional flow the effect of
Reynolds number on the boundary between
transitional and fully turbulent flow has been
investigated for missile-like bodies which have
pointed noses and blunt bases in Reference (3). This
reference indicates that the minimum Reynolds
number, based on the maximum diameter of the
missile, for which there is fully turbulent flow is 0.5
million at an angle of attack of 20 degrees, 0.8
million at 40 degrees, and 6.0 million at an angle of
attack of 90 degrees. However, experiments that have
been performed at NSWCCD appear to indicate that
the force coefficients do not vary significantly with
Reynolds number above a Reynolds of about 0.8
million based on maximum diameter. What were the
effects of Reynolds number on the results of the
Static Test Rig wind tunnel experiments?
The out-of-plane normal force coefficient
Zvv in Equation (4c) represents the pressure
difference over the hull due to the lift developed on
the sail when the submarine is underway at an angle
of drift. Equation (4d) should include the rolling
moment that depends on the product vw. Equation
(12b) indicates that the effect of rolling angular
velocity on the rolling moment is linear. However, at
zero forward speed the damping is primarily due to
the cross-flow drag developed by the bridge fairwater
(sail) and stern appendages. The contribution of the
sail is Kp|p| = CCFD(F)1/2AzF3 where CCFD(F) is the
cross-flow drag coefficient, A is the planform area,
and zF is the distance from the center of pressure to
the centerline. The coefficients Y, K, and N in
Equation (20) should be zero, and the characteristic
equation should be a quartic rather than a quintic.
The abscissa in each of the six plots that
make up Figure 8 show positive flow incidence
angles both to the left and right of zero. I thought
this was an error, but it appears that the left side of
the plot is to show flow orientation angles of -30, -60,
and -90 degrees and the right side of the plot is to
show flow orientation angles of 90, 120, and 150
degrees, both over a range of positive flow incidence
angles. However, in Figure 9 only positive flow
orientation angles are shown. It would help if the
legend for the flow orientation angles were included
in all of the six plots in Figure 8 rather than only the
rolling moment plot.
I think that the roll angle instability you are
investigating occurs only if the normal component of
velocity w (or the angle of attack) is negative. If w is

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

positive during the buoyant rise after the blowing of


ballast, the direction of the hydrodynamic rolling
moment that is developed on the sail tend to oppose
the roll angle of the submarine.
In Figures 8 and 9 you show the
destabilizing rolling moment as a function of the flow
orientation and flow incidence angles. However, the
moment is also a function of the angle of attack and
angle of drift. Is there a reason why the data was not
plotted against these two variables?
Since I do not have your 2001 paper, I do
not understand why in Equation (23) the cosine of the
pitch angle is not multiplied by the sine of the roll
angle, and why the destabilizing non-dimensional
rolling moment that is shown in Figures 8 and 9 as a
function of flow incidence angle and flow orientation
angle does not appear in the denominator instead of
the partial derivative of rolling moment with flow
orientation angle. Is there a relationship between the
roll angle of the submarine and both the flow
orientation and the flow incidence angles so that the
two variables can be reduced to a single variable?
REFERENCES
1. Feldman, J. P., "DTNSRDC Revised
Standard Submarine Equations of Motion,"
DTNSRDC/SPD-0393-09, June 1979.
2. Hess, D. E., T. C. Fu, and J. P. Feldman,
"Naval Maneuvering Research and the Need for
Shear Stress Measurements," AIAA Paper 2004-2605
for the 24th AIAA Aerodynamic Measurement
Technology and Ground Testing Conference, JuneJuly 2004.
3. Lamont, P.J., "Pressures Around an
Inclined Ogive Cylinder with Laminar, Transitional,
or Turbulent Separation," AIAA Journal, Volume 20,
Number 11, 1982.

experimentation but will be computationally


intensive.
We have looked for Reynolds number
effects by comparing the STR data with data acquired
in other facilities that have tested the same generic
model shape shown in this paper. See:
M. Mackay, The Standard Submarine Model: A
Survey of Static Hydrodynamic Experiments and
Semiempirical Predictions, DRDC Atlantic TR 2003
079, June 2003
Reynolds numbers range from 7 to 23 million based
on hull length, or 0.8 to 2.6 million based on
maximum diameter. The variation in the data falls
within experimental error.
This error includes
differences between facilities, models, balances, and
inconsistently applied support strut corrections. This
suggests Reynolds number effects are insignificant, at
least to within the ability of the above data to resolve
them.
However, recent tests in a compressible flow
wind tunnel, using a single model, balance, and
support system for each configuration tested,
provided Reynolds number variations from 7 to 40
million (based on hull length) while maintaining
reasonable resolution in the force measurements
(Reynolds number is changed by changing air
density, not onset velocity, so balance resolution goes
farther). These tests show that the linear transverse
force derivatives for unappended hulls decrease from
7 to 21 percent as the Reynolds number increases
from 7 to 20 million, depending on afterbody fullness
(slender hulls see greater change). Further, they
show that sternplane effectiveness changes are much
larger over the same Reynolds number range. Little
change is seen from R = 20 to 40 million.
Preliminary sternplane effectiveness results are
presented in:

AUTHORS RESPONSE
Thank you very much for your detailed
review.
Your discussion indicating that the standard
equations now play a reduced role in your analysis
of extreme maneuvers is consistent with our
experience that they do not predict the rising
instabilities in our trials. Your multi-vortex method
sounds interesting; presumably it is computationally
fast but requires much empirical input from the many
experiments you describe. For our part, we are
developing the capability to carry out a 6 DOF RANS
simulation to explain rising stability dynamics; this
will
minimize
the
necessity
for
costly

M. Mackay, H.J. Bohlmann, G.D. Watt, Modeling


Submarine Tailplane Efficiency, RTO SCI120
Meeting Proceedings, Paper #2, Berlin, May 2002.
The effect of the out-of-plane force on the
rolling moment is zero in (4d) because the HST
model has an axisymmetric hull, so the force acts
through the body x axis creating no moment. The
deck on our HSSDT model does result in an out-ofplane force contribution to rolling moment and this is
captured by fitting (3b) to the data. However, Watt
(2001) shows that the HSSDT out-of-plane moment
component is quite small. Thus, identifying and
extracting this component is not done in Section 4.4.
It is worth mentioning the missing out-ofplane moment in (4d) because Boats A C have

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

larger decks than our generic model. This highlights


a general shortcoming of the current analysis: We
are using the hydrodynamic characteristics of a
generic shape to try to explain trials results from
specific boats having varying hull geometries, sail
locations, etc.
The full rolling moment equation should
contain a Kp|p| term. However, as noted in Section
4.3, we only require a linearized model in p and r
because the stability analysis assumes these states are
small.
An appendix has been added to the paper
explaining why our stability analysis gives rise to a
quintic rather than a quartic characteristic equation
(Equations 20 and 21). Essentially, it is because axial
acceleration is not neglected as is done in the
classical horizontal plane stability analysis.
Acceleration is accounted for because the submarine
is constantly accelerating during buoyant ascent.
Unfortunately, there isnt room in Figure 8
to add the key to each of the six plots. Instead, an
explanatory sentence has been added to the caption.
Agreed, the roll stability analysis applies
only to negative w values (negative angles of
attack). This is accounted for in the analysis by
evaluating the stability derivatives at = 0 (see (23)
and the sentence immediately following (27)). When
w is positive, = 180 degrees.
Section 4.7 attempts an explanation of why
the forces and moments are modeled using flow
incidence and flow orientation instead of the
conventional angles of attack and drift.
This
description was initially used when first acquiring
experimental data on a turntable (directly controlling
) at various roll () orientations of the model.
Initially, the automated turntable sweeps had to take
place at constant because roll was not automated.
And it was both bothersome and intuitively less
informative to convert these angles to combinations
of angle of attack and drift. This is the Figure 8 data,
which is not readily presented in a conventional
format. Having become familiar with thinking about
flow incidence and orientation, its advantages
became apparent. High among these are the ability to
extend indefinitely the standard model (4) using
conventional Fourier analysis (3). In addition, this is
a natural way to describe the rising stability problem:
consider Figure 1 where the boat is rising at
approximately constant incidence while rolling.
What apparently determines roll stability there is the
rate at which K changes for a given . This led to
the development of an automated roll sweep
capability and acquisition of the Figure 10 data.
Subsequently, a 1 DOF rolling moment stability
analysis (Watt, 2001) assuming = and decoupling

from the analysis resulted in the roll stability limit


given by (23).
To address the Boat B and C trials data, as
well as the lack of success of (23) in predicting the
stability limit, the current stability analysis was
carried out. This required a more formal derivation
of the assumptions used in the 1 DOF analysis. This
derivation has not been presented as clearly as it
should have been, which has resulted in your last
question about the relationship between , and
.
If F(,) represents any of the horizontal
plane forces or moments (Y, K, N) then, for the
stability analysis, F must be linearized in the
perturbations = y0, , . From Equations (1) and
(14) we note that:
0 + O( 2 ),

v
+ O ( 3 )
w

where 0 is independent of the perturbations and v


is linear in them. Because 0 is independent of , 0
is just a parameter in the stability analysis; it is the
flow incidence immediately prior to instability. No
distinction is made between and 0 in Section 5.
The linearized horizontal plane forces can
now be written in the form:
F
( 0, 0)( 0 )

F
( 0 ,0)
+

F (, ) F ( 0 ,0) +

But, from (3b), the first term and the partial


derivative in on the RHS of this expansion are
zero. In addition, 0 is O(2). Only the last
term remains, where:

x&
y&
v
= 0 + 0
w
w
w

using (14b). This rederives (19), but it is more


explicit in that it shows that the stability derivative
F is evaluated at = 0 and = 0.
The last equation above justifies the =
relation used in the 1 DOF stability analysis where
, y& 0 are assumed to be zero. This analysis of the
rolling moment equation of motion effectively
cancels sin from the static stability term with sin
from the hydrodynamic rolling moment (cf,
Equation 3b) in the ratio under the radical sign in
(23).

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Jerome P. Feldman
Naval Surface Warfare Center, Carderock Division,
USA
I am not sure why you converted the
translational and rotational velocities from a body
coordinate system to velocities in a fixed, inertial
coordinate system. Why not perform your analysis
entirely in the body coordinate system? What effect
does this have on the stability analysis? If you had
not made the conversion, would there have been a
quadratic characteristic equation?
Also, why did you leave terms multiplying
qr, qp, wp, and pq in Equation (15a), and terms
multiplying qr, pq, wp, wp, pq, vq, wr, and qr in
Equation (15b)? The product of two small quantities
is so small that it can be neglected when you linearize
the equations.

The only terms neglected in (15) were those


that are second order in the perturbation quantities,
the ( listed in the nomenclature. The vertical plane
velocities u, w, q are not assumed to be small. As
noted in the text following (14), these are just
parameters in the analysis. And, indeed, this is why
(, q, and EMBED Equation.3
in Figure 14
do not have to be small. Note that the simplified
expressions in (26) and (27) are based on the
simplifications (25) which, when applied, indirectly
do discard some of the terms you mention.
Because all the algebraic manipulations are
handled by the computer, we are not compelled to
implement unwanted simplifications just to get a
manageable result. Manageable, algebraic results are
certainly desirable, and these are presented in Section
5.3, but they have increased value when their
relationship to the exact (such as it is) solution is
known; this relationship is shown in Figure 14.

AUTHORS RESPONSE
Inertial coordinates were used in the stability
analysis because of our mind's eye view of what was
going on physically, as described in the first
paragraph of Section 5.1. A true, fixed vertical plane
is best described using inertial coordinates, or so it
would seem. This also addresses a concern we had
with the trials data where we had no measure of
lateral velocity and so did not know to what extent
the vehicle was moving away from a fixed vertical
plane. This approach also let us work directly with
the heading , and the trials data clearly show that
is undergoing some instability.
In the paper we suggest that by working
with inertial coordinates we introduce an extra degree
of freedom (r and r& are replaced with , & , and &&).
But, in the simplified analysis presented in the
Appendix, by itself this does not give rise to a higher
order characteristic equation even though it does
result in nonzero Y , N coefficients in the equations
of motion (cf, Equation 20). In the Appendix
analysis, if &x&0 is set to zero, then the nonzero Y , N
coefficients combine with, and are cancelled by,
Y y&0 , N y&0 coefficients so that the A3 coefficient in
(33) is zero. In other words, it is only by accounting
for a nonzero &x&0 term that the characteristic equation
increases in order and a new stability limit is found.
Could the new stability limit have been
found in an analysis based on body axes velocities?
We haven't tried this, but the &x&0 term in (32)
suggests that an analysis based on is required.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics


St. Johns, Newfoundland, CANADA, 813 August, 2004

Validation of a Quasipotential Flow Model


for the Analysis of Marine Propellers Wake
L. Greco, F. Salvatore, F. Di Felice
(INSEAN, Italian Ship Model Basin, Italy)

ABSTRACT
A theoretical model for the analysis of marine propeller
wake flow is presented. The proposed methodology is
based on a boundary integral formulation for the velocity potential. A wakealignment technique is formulated
in which the near wake shape is determined as a part of
the flowfield solution, whereas a far wake model allows
to include the effects of the vortices extended to infinity.
A viscousflow correction to the potential field is determined through Lighthills transpiration velocity concept.
Numerical results by the proposed methodology are presented in order to verify the numerical scheme and to validate propeller slipstream flow predictions against model
propeller measurements by using PIV and LDV velocimetry techniques.

INTRODUCTION
The development and validation of a theoretical model for
the hydrodynamic analysis of marine propellers wake flow
is addressed.
The issue of wake flow modelling is related to the
correct evaluation of propeller performance. The mechanism of lift/thrust generation is intrinsically connected to
the amount of vorticity that is generated on the body surface and is shed downstream in the wake. In the theoretical analysis of isolated marine propeller flows, wake flow
modelling is of particular importance when the propeller
operates below design point, the wake pitch is small and
the trailing vorticity pattern tends to be packed close to the
propeller disc plane. Wake flow modelling is also important for cavity inception and tipvortex cavitation investigations and for the analysis of the mutual interactions
between propellers and other components as in the case of
propeller/rudder systems of conventional arrangements or
in the study of propeller/strut flow for podded propulsors

or propeller/propeller interactions for complex propulsors.


Physical models of the propeller slipstream should
address the evolution of a thin vortical layer shed at the
propeller blade trailing edge in a viscous turbulent flow.
However, propellers at model and full scale are characterized by very high Reynolds numbers and hence viscosity
induced vorticity diffusion plays a minor role in the flow
region within twothree diameters downstream the propeller disc plane. In this region, the most relevant features
are vortical sheet deformation and rollup towards the tip,
slipstream contraction and vorticity concentration in the
tipvortex.
In order to develop designoriented computational
tools that provide reliable predictions at low computational costs, a theoretical approach based on inviscid
flow assumptions may be considered. In this context,
potentialflow methods are used in which the wake is represented as a zerothickness vortical layer and a discontinuity surface for the velocity potential (trailing wake).
Early investigations on the effects of wake shape variations on propeller thrust and torque predictions by potential flow methods are given by Kerwin and Lee (1978).
Semiempirical procedures to adjust the wake pitch and
contraction to flow conditions and propeller loading are
discussed e.g. by Hoshino (1989) and by Kawakita
(1992). Fully theoretical methods to determine the wake
shape as a part of the potential flow solution by a vortex lattice method are proposed by Greeley and Kerwin
(1982) and by Kinnas and Pyo (1999), where wake alignment effects on blade sheet cavitation modelling is studied. More recently, the performance of a wake surface
relaxation scheme combined with a boundary element
model of the propeller flow is investigated by Liu and Colbourne (2002).
The methodology presented here is based on a boundary element method for the velocity potential, and extends

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

previous work described in Giordani et al. (1999) and


Salvatore et al. (2003). The term quasipotential is introduced to distinguish the present formulation where a
viscosityinduced correction to the potential flow is introduced. Specifically, an approximated solution of the
boundary layer flow surrounding the propeller blades and
of the viscous wake is obtained coupling the potential
flow solver with a boundary layer solver by means
of Lighthills transpiration velocity concept (Lighthill,
1958). An estimate of the vortical layer thickness in the
wake and of the velocity profile in the vortex core may
be determined. The trailing wake evolution mechanism
is described by means of an alignment procedure that is
based on a Lagrangiantype scheme, and wake points velocity is determined through a boundary integral representation. The formulation also includes a novel far wake
model derived by the Helmholtz decomposition of the velocity field. This allows to take into account the effects of
the wake portion extended to infinity with remarkable advantages in terms both of accuracy and of computational
effort reduction.
In the present paper the proposed theoretical formulation is described with enphasis on both wake alignment
and far wake models. Numerical results are also presented
to discuss the numerical scheme performance and to assess the capability of the proposed approach to provide accurate predictions of the wake flow. In particular, numerical results are validated against measurements of the wake
flow behind the INSEAN E779A model propeller tested in
the cavitation tunnel of the Italian Navy (CEIMM) by using LaserDoppler and Particle Image velocimetry techniques.

THEORETICAL FORMULATION
The boundary integral formulation for the perturbation velocity potential is briefly reviewed, whereas descriptions
of both trailing wake alignment and far wake modelling
are presented in detail. Fully inviscid flow conditions are
considered first. Next, the issue of viscosity induced correction to the potential flow solution through the transpiration velocity concept is discussed.

flow are neglected, the perturbation velocity v may be expressed in terms of a scalar potential as v = . The
total velocity field in the propeller frame of reference results
q = vI + .

Assuming the flow is incompressible and recalling


v = , the continuity equation v = 0 reduces to
the Laplace equation for the velocity potential 2 = 0.
The momentum equation reduces to Bernoullis theorem that, in the propeller frame of reference, reads
1
1
p0
p

+ q2 +
+ gz0 = vI2 +
,
t
2

The propeller flow is studied by considering a right


handed frame of reference (Oxyz) fixed to the propeller
with the x-axis parallel to the propeller axis and downstream pointing (propeller frame of reference). The incoming flow velocity is vI = vA + x, where vA is
the upstream flow, and = 2nex is the angular velocity
of the propeller.
Basic assumption of the present model is that the velocity perturbation induced by the propeller to the incoming flow is irrotational. If viscosity effects in the upstream

(2)

where p is the pressure, q = kqk, vI = kvI k, whereas g


is the gravity acceleration and z0 denotes depth.
The Laplace equation for is solved in the potential
flow region delimited by the body surface SB , by the trailing wake SW and extended to infinity. Non cavitating flow
conditions are considered.
On the body surface the impermeability condition
yields q n = 0, or, recalling Eq. (1),

= vI n
n

on SB .

(3)

As stated above, the trailing wake represents a zero


thickness layer where the vorticity generated on the propeller surface is shed downstream at the blades trailing
edges. In the framework of potential flows, this vorticity
layer determines a discontinuity surface for the velocity
potential. Mass and momentum conservation laws are imposed across SW to obtain that both pressure and the normal component of the perturbation velocity /n are
continuous



p = 0;

(4)
= 0; on SW ,
n
where the symbol denotes discontinuity across the
wake surface sides. Combining the Bernoulli Eq. (2) and
p = 0, one obtaines that is constant following wake
particles
(x, t) = (xT E , t )

Inviscid flow modelling

(1)

on SW ,

(5)

where xT E is a wake point at the blade trailing edge and


denotes the convection time of particles moving on the
wake surface with velocity q = vI .
A further condition on is required in order to assure that no finite pressure jump may exist at the body
trailing edge (Kutta condition). Following Morino et al.
(1975), this condition is approximately achieved by imposing that at the wake trailing edge equals the difference between potentials at the two sides of the blade
trailing edge surface.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

The solution of the Laplace equation for is obtained


here through a boundary integral formulation. A classical
approach based on the third Green identity yields for an
arbitrary point x immersed into the fluid

I 

G
(x) =
G
dS(y)
(6)
n
n
SB


 
Z
G

G
dS(y),
+

n
n
SW
where n is the unit normal to SB and to SW . The quantity G = 1/4kx yk denotes the unit source in the
unbounded threedimensional space.
In the limit as x tends to the body surface, Equation (6) provides a boundary integral equation for . The
resulting equation is formally equivalent to Eq. (6) with
(x) replaced by 21 (x) if x is a regular point on SB .

Trailing wake alignment


The location of the wake surface in Eq. (6) is not known
a priori and may be either prescribed or determined as a
part of the flowfield solution. The latter approach is considered here in order to describe the vorticity convection
mechamism downstream the propeller disc.
A physicallyconsistent wake shape may be determined recalling that the trailing wake represents the vortical layer where all the particles that came in contact with
the body surface are shed downstream. As a consequence
of this, the wake surface is determined by imposing that
wake points must be aligned to the local flowfield. The
procedure is labelled as wake alignment and is achieved
in two phases. First, the perturbation velocity field in the
wake region is computed using a boundary integral representation for the velocity potential. This is obtained by
taking the gradient of Eq. (6), to obtain at any point
on the wake surface.


I 
G

x G x
dS(y) (7)
x (x, t) =
n
SB n



 
Z
G

x G x
dS(y).
+

n
n
SW
The symbol x denotes the gradient operator acting on
x. Next, wake points xW are moved parallel to the local
velocity field q = +vI by using a Lagrangian scheme
xW (t + t)

= xW (t)
(8)
Z t+t



+
xW , t + vI dt.

the following representation of the velocity induced by a


constant dipole distribution over an arbitrary open surface
Sn (see, e.g. , Campbell, 1973)
Z
I
G
1
r dy
,
(9)
x
dS(y) =
4 Sn
r3
Sn n
where r = y x and r = krk. According to Eq. (9), a
constant dipole distribution on Sn induces the same velocity as a closed vortex lying on the surface boundary Sn .
The quantity in the righhand side of Eq. (9) is the
vortexinduced velocity by the BiotSavart law. It is apparent that an infinite velocity is evaluated as r goes to
zero. This would determine stability problems in the numerical procedure related to the wake alignment technique. Furthermore, infinite velocity is unphysical in
that viscous flow phenomena inside the vortex core limit
the velocity magnitude. In order to overcome unphysical trends and to reduce numerical instabilities, the finite
vortex core concept is introduced by Suciu and Morino
(1978). This denotes a small flow region surrounding each
wake vortex where the intensity of the velocity induced at
a given field point is directly proportional to the distance
between the vortex axis and the field point.
Specifically, a vortex core radius r is defined as follows
p
(10)
r = r0 1 + r ,
where r0 is the vortex core radius at the wake trailing
edge, is the arclength in the streamwise direction and
r is a growth factor. Thus, the velocity vn induced by
a unit dipole distributed on Sn is recast in terms of Eq. (9)
and the vortex core concept as follows
I

r dy
1
,
r r
(11)
vn = 4 Sn r3

r/r vn ,
r < r
where vn is evaluated by the BiotSavart law for r = r .

Far wake modelling


In view of the numerical calculation of Eqs. (6) and (7),
three regions in the propeller slipstream are distinguished:
near wake: denotes the wake portion closest to the
propeller, where the trailing wake alignement technique is applied;
intermediate wake: a limited wake portion downstream the near wake and defined as a helical surface
with prescribed pitch;

where t denotes time.


The combination of Eqs. (7) and (8) in the wake alignment procedure requires a careful evaluation of wake
induced velocity. In order to focus the problem, consider

far wake: the wake portion downstream the intermediate wake and extended to infinity.
Both near and intermediate wake portions are explicitely
taken into account when solving Eqs. (6) and (7),

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Figure 1: Representative discretization grid on the wake surface: near wake portion (in blue), intermediate wake (in red)
and far wake disc (in black).

whereas a model adapted from a formulation proposed


by Gennaretti et al. (1998) is proposed here to account
for the far wake.
The basic idea is to split the perturbation velocity field into a contribution induced by body and
near/intermediate wake portions, and a contribution vF W
induced by the far wake surface SF W
= + vF W

(12)

Denoting by SN IW the wake surface in the near and intermediate regions, it may be found that the following integral representation for is valid at an arbitrary field point

I 

G
(x)

=
G
dS(y)
(13)
n
n
SB
 


Z
G

G
dS(y),

n
n
SN IW
with boundary conditions on the body surface as follows
(compare Eqs. (3) and (12))

increased, the wake vorticity tends to concentrate in the


tip vortex.1 The far wake influence on the potential solution may be approximated by taking into account only the
tip-vortex contribution and neglecting residual vorticity at
inner radial locations. Furthermore, tip vortices describe
helicoidal trajectories that are sufficiently packed as seen
from the blades. Thus, the vorticity field in the far wake
may be approximated by a semiinfinite cylindrical surface SC with distributed vorticity of intensity = /PF W
where is the tip vortex intensity and PF W is the pitch of
the helical surface described by the far wake.
Without lack of generality, one may assume that the
induced by the vortical surface SC is povelocity field v
tential and incompressible inside the cylinder, and equals
zero outside. Denoting by SD a disc that represents the upstream lid of the vortical tube, the Helmoltz velocity field
decomposition may be used to obtain (Batchelor, 1967)
Z
) dS(y)
vF W (x) = x
(G
nv
SD

Z

= (vI + vF W ) n
n

on SB ,

(14)

where vF W is the velocity field induced by the far wake,


whereas the conditions for on the wake are given by
Eqs. (4) and (5) unchanged.
Recalling Eq. (7) and the second of Eqs. (4), yields



I
G
dS(y).
vF W (x) =
x
n
SF W
Equation (13) can replace Eq. (6) once a suitable
model for vF W is given. In order to accomplish this, a
simple geometrical schematization for the vorticity field
in the far wake is proposed first. Experimental evidencies show that, as the distance from the propeller disc is

) dS(y), (15)
(G
nv
SD SC

is the inward unit normal to the cylinder surwhere n


face. From Eq. (15), the velocity field vF W is given by
a distribution of vortices on SD and by a distribution of
sources on SD and SC . The unknown intensity v may be
determined using a simple actuator disc model. In particular, recalling that the mass flow within the propeller
wake streamtube is constant, it is possible to evaluate v
on the disc surface once the propeller induced velocity is
1 The wake rollup process in the tip region is a consequence of the
vorticity concentration process; this issue is discussed later in the paper
where numerical results by the present model are compared to vorticity
field measurements.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

known. In the present approach, such induced velocity


is assumed to be parallel to the propeller axis, and hence
Eq. (15) yields
Z
vF W (x)

= x

v G dS(y).

(16)

SD

Viscous flow correction


The present model is valid in the limited case of very high
Reynolds number flows2 around propellers close to design
operating conditions. Under these assumptions, the viscosity induced vortical layer is thin and attached to the
propeller blade surface, whereas the viscous wake may be
regarded as a thin vortical layer surrounding the trailing
wake surface.
Viscosity effects are studied here in the context of integral boundary layer models. Boundary layer equations
are integrated along the normal direction to the blade surface and to the trailing wake in order to reduce the number
of equations and hence to limit the computational effort.
In the case of twodimensional flows, this approach leads
to the classical von Karman equation that is valid in an inertial frame of reference. In the present approach, a generalized von Karman equation is proposed to address three
dimensional flows in the rotating propeller frame of reference in order to include crossflow effects on the boundary layer growth. Specifically, the following equation in
tensor form is obtained (see Salvatore and Testa, 2002, for
details)
(17)

+ qv [t qv ] t + 2 qv t =

(qv q) d

t = 1/qv
0

T = 1/qv2

(qv q) q d;
0

fW = W e + W e

Thus, far wake effects on the propeller flow may be taken


into account once Eqs. (13) and (14) are used to replace
Eqs. (6) and (3), respectively. A similar modification is
used to include far wake effects into Eq. (7).
A sketch of the propeller wake surface divided into
near, intermediate and far disc portions is depicted in
Fig. 1. It should be observed that, in order to avoid intersections between the last portion of the intermediate wake
and the far disc, this latter is shifted downstream of a small
quantity dD .

(qv t ) + t qv2 T
t

are introduced

1
f
W

where qv denotes the total velocity field computed by using the corrected quasipotential flow (see below), and
qv = kqv k. In Eq. (17) the following tensor quantities
2 Model propeller flows are typically in the range of the Reynolds
number ReD = nD2 / 106 , where D is the propeller diameter and
the kinematic viscosity. Larger values can occour at full scale.

(18)

Quantities t and T represent, respectively, generalized


displacement thickness and momentum thickness. Classical onedimensional expressions are obtained by representing the above tensors in terms of components in a
given coordinate system. Moreover, the quantity fW denotes the wall tangential stress vector, and W , W are,
respectively, its components along chordwise and spanwise directions , .
In order to obtain a numerical solution of Eq. (17),
additional equations must be introduced to balance the
number of unknowns represented by t , T and fW
components. Such closure equations are typically formulated along a direction that is aligned to the local flowfield at the outer edge of the boundary layer. A robust and
straightforward approach proposed by Nishida (1996) is
used here. Equation (17) is projected onto a locally Cartesian surface coordinate system (OxN yN ), where xN is locally parallel to the chordwise direction.
Next, boundary layer quantities are referred to corresponding quantities in a flowfitted coordinate system
(OxE yE ), with xE aligned to qv . Denoting by R the rotation matrix between coordinate systems (OxN yN ) and
(OxE yE ), one has
(t )x
(T )x

yN

= R (t )x

yN

= R (T )x

yE

yE

;
RT ,

(19)

Different closure equations are imposed in the laminar and turbulent regions of the boundary layer. Downstream the blade leading edge the flow is assumed to
be laminar, and a twodimensional Thwaites collocation
method (Thwaites, 1949) is used. Transition to turbulent
flow is detected through Michels method (Michel, 1952).
The turbulent portion of the boundary layer and the viscous wake are solved by coupling the LagEntrainment
closure model by Green et al. (1973) for twodimensional
flows with a Johnstons triangular velocity profile model
to include threedimensional effects (Johnston, 1957).
Specifically, the following quantity is introduced


Ac = qYE / qv qXE = Ac (W )
(20)
where W is the angle between the flow at the solid wall
and the inviscid flow. Equation (20) allows to express
crossflow displacement thickness, momentum thickness
and wall tangential stress as a function of the corresponding quantities along xE (see Salvatore and Testa, 2002, for
details).

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

The boundary layer solution is matched with the inviscid flow solution by means of Lighthills transpiration velocity concept (Lighthill, 1958). Basic assumption is that viscosityinduced effects may be included into
a quasipotential flow model by suitably modifying the
boundary conditions on the body and on the trailing wake
in order to take into account the streamlines displacement
induced by the boundary layer thickness. A mathematically rigorous generalization of this approximated model
is presented by Morino et al. (1997). It may be shown that
modified boundary conditions Eq. (3), and the second of
Eqs. (4) yield

= vI n + v , on SB ;
n


= v , on SW .
n

(21)

where the quantity v denotes the transpiration velocity


v = S (qv t ) =




qv ? +
qv ?

(22)

where ? , ? are, respectively, displacement thicknesses


along chordwise and spanwise directions , .

Solution procedure
A propeller hydrodynamics computational tool has been
developed by combining the trailing wake alignment technique, the far wake model and the viscosity effect correction into a nonlinear boundary element methodology. The
solution procedure used in the present analysis is valid for
uniform upstream flow vA = vA ex and constant propeller
rotational velocity n, and hence the flowfield is stationary if observed in the propeller frame of reference. The
numerical approach is based on a twosteps iterative procedure where viscous flow correction is applied first, and
next wake alignment procedure is performed.
The core of the procedure is the solution of the boundary integral equation for the velocity potential on the
body surface. As stated above, this equation is obtained as
the limit of Eq. (6) as x tends to SB . Neglecting far wake
and boundary layer effects, this boundary integral equation is solved with /n on SB known through the impermeability condition (3), whereas conditions given by
Eqs. (4) and (5) are imposed on the wake.
By using a prescribed wake geometry, a firstguess
solution for may be obtained. The evaluated potential flow is used to solve the boundary layer equation (17)
and to determine the transpiration velocity distribution by
Eq. (22). Thus, a corrected potential flow solution may
be determined by solving Eq. (6), with Eqs. (3) and (4)
replaced by Eqs. (21). This process is iterated until convergence of the velocity distribution qv . It should be observed that the viscousflow correction is performed here

only to provide an estimate of global boundary layer quantities. An approximated numerical solution of Eq. (17) is
obtained using a striptheory approach, with explicit solution of the flow in the chordwise direction and crossflow
contributions assumed known from past iterative steps.
Next, the wake alignment model is switched on and
the evaluated potential field on the body surface is used
to estimate the perturbation velocity at wake points by using the boundary integral expression given by Eq. (7). In
particular, the vortex core radius r0 and the growth factor r in Eq. (10) are determined through the boundary
layer flow solution. The resulting total velocity field is
used to deform the initial guess wake surface according
to the flow alignment technique through Eq. (8). The updated wake surface is then plugged into Eq. (6) and a new
estimate of the potential field is obtained. The process is
repeated until convergence of the wake shape.
In case the far wake model is used, the above procedure is still valid once the far wake velocity is evaluated by
Eq. (16) and the modified inflow vI + vF W is used to replace vI . Thus, the most general expression for boundary
conditions over the body surface including both viscous
transpiration velocity and far wake induced velocity reads

= (vI + vF W ) n + v ,
n

on SB ,

(23)

whereas on the near/intermediate wake the second of


Eqs. (21) is unchanged.

NUMERICAL SCHEME VERIFICATION


The solution procedure described above is implemented
into a boundary element code. Body and wake surfaces
are discretized into hyperboloidal quadrilateral elements.
Computational grids are characterized by the number of
blade elements in chordwise direction MB (from leading
edge to trailing edge) and in spanwise direction NB , the
number of wake elements in streamwise direction per turn
MW and in radial direction NW , and the total number of
hub surface elements NH . The numerical analysis discussed in the present work addresses the INSEAN E779A
model propeller (see below) in uniform inflow conditions
at constant angular velocity.
Flow quantitites are supposed to be piecewise constant on each element. The boundary integral equation
derived from Eq. (6) is enforced at each element centroid
on the body surface, and hence its solution is reduced to
the solution of a linear system of equations. Source and
dipole integrals in Eqs. (6) and (13), as well as source and
dipole gradient integrals in Eqs. (7) and (16) are evaluated
by analytical formulas proposed by Morino et al. (1975).
Consider first the wake alignement procedure without
far wake contributions. In the present calculations, the initial guess wake is a helical surface that leaves the trailing
edge tangent to the mean line of each radial blade section

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

and with the pitch given by the propeller mean geometrical pitch. At each step of the wake alignment iterative
procedure, wake nodes locations are updated according to
Eq. (8), that, in discretized form reads
i t
xp+1
= xpi1 + q
i

(24)

where the superscript p denotes the iteration step, and i is


the wake node index in streamwise direction (i = 0 at the
i = (1 W ) qi1 +
blade trailing edge). In addition, q
W qi , whereas [0, 1] is a weight factor (typically,
= 1/2), and t is the averaged convection time.
In order to reduce the computational effort, the alignment technique is applied only on the near wake, where
most of the slipstream contraction is assumed to take
place. The intermediate wake surface is determined as a
helical surface where the radial pitch distribution equals
the local pitch distribution at the last portion of the near
wake (frozen rollup model).

Figure 3: Effect of grid refinement on the calculated tip


vortex radial position (top) and intersection with a trasversal plane at x/R = 0.65 (bottom).

Figure 2: Effect of grid refinement on propeller loads (non


cavitating flow at J = 0.88). Top: thrust coefficient KT ,
bottom: torque coefficient KQ .

The numerical scheme consistency, i.e. the effect of blade and wake grids refinement on the wake
alignment procedure is discussed by considering different discretizations for the analysis of the propeller flow
at advance coefficient J = vA /nD = 0.88, where D
is the propeller diameter. Specifically, five grids with
MB = NB = 12, 18, 24, 30, 36, and MW = 5 MB =
60, 90, 120, 150, 180, and NW = NB are considered.
Figure 2 depicts calculated propeller thrust and torque
coefficients, KT = T /n2 D4 and KQ = Q/n2 D5 . Numerical results by using both prescibed and flowaligned
wake surfaces are shown.
The effect of grid refinement on the flowaligned
wake shape is illustrated in Fig. 3. Only three representative grids among the grid set introduced above are considered: MB = 12 (coarse grid), MB = 24 (medium grid),
MB = 36 (fine grid), and MB = NB , MW = 5 MB .
Specifically, the predicted tipvortex radial location as a

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

function of the axial location x/D is shown in top Fig. 3.


In this case, the near wake is extended up to one diameter downstream the propeller disc. Oscillations of the
tip-vortex are observed in the near wake region as a consequence of the wake rollup process at the tip. In the
intermediate wake portion at x/D > 1.0, the rolledup
shape of the last free-wake portion is kept constant according to the frozen rollup model. Bottom Fig. 3 shows
grid refinement effects on the predicted intersection between the trailing wake surface and a transversal plane located downstream the propeller at x/R = 0.65. It should
be observed that there is a substantial agreement among
different grids predictions, even if a large number of discretization elements in the spanwise direction is required
to capture the wake rollup at tip, where most of the trailing vorticity concentrates.

rection. Next, RW rapidly drops to low values and the


iterative procedure convergences (in the present calculations, RW 105 is used as the convergence criterion).
This trend is typical of the wake shape updating procedure
based on Eq. (8).

Figure 5: Effect of initial guess wake geometry on wake


displacement after flow alignement procedure

Figure 4: Effect of grid refinement on the wake alignment


convergence history index RW .

Next, the convergence of the iterative procedure to


determine the flowaligned wake shape is addressed. In
particular, the root mean square RW of wake nodes displacements between two subsequent iteration steps is considered as the convergence index
v
uN tot
uX
2
u W  (k)
1
(k1)
(k)
RW
= tot t
xi xi
(25)
NW
i=1
tot
where k is the iteration step and NW
denotes the total
number of near wake discretization elements. Figure 4
shows convergence histories of the coarse, medium and
fine grids previously described. Two different phases in
the iterative process can be clearly distinguished: first,
quantity RW grows up to a maximum value which is
reached after a number of iterations that is roughly related
to the number of near wake elements in streamwise di-

Another issue concerning the verification of the iterative solution of a nonlinear problem is the analysis of the
influence of the initial guess wake geometry on the converged solution. Five initial guess geometries, that differ only by the pitch distribution, are considered, while
the meshing grid is kept fixed. Specifically, the helicoidal
surface whose pitch equals the propeller blade mean geometrical pitch 0 is chosen as the reference geometry, and
four surfaces with = k 0 with k = [0.6, 0.8, 1.2, 1.4]
are also considered. Figure 5 depicts the convergence history of the radial position of a selected wake panel centroid located, considering the initial guess wake shape, at
x/R = 0.2. As expected, the converged position of the
centroid, after about 60 iterations, presents a negligible
dependence from the initial guess wake shape.
Next, the performance of the proposed far wake
model on the propeller flow prediction is investigated.
Consider first prescribed wake calculations. Figure 6
shows calculated thrust and torque coefficients against the
extension of the near/intermediate wake, given in terms of
the total number of wake turns considered, Nspiral . Numerical results with and without the far wake model are
compared. The figure clearly explains the practical utility
of the far wake model. Predicted loads by using only two
wake turns and the far wake disc are equivalent to those
using five wake turns without the far wake model. Thus,
using the far wake model it is possible to limit the extension of the wake surface in the calculations and hence the

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

computational effort may be reduced.

Figure 6: Influence of the total number of wake turns


Nspiral on calculated loads by using a prescribed wake
geometry. Results with and without far wake disc compared. Top: thrust coefficient, bottom: torque coefficient.

The computational advantages related with the introduction of the far wake model become apparent in the
case of wake alignment. Figure 7 shows the effect of
reducing the intermediate wake extension LIW with the
near wake lenght LN W kept constant on propeller thrust
and torque calculations. Results by using the far wake
model are compared to those obtained without the far
wake model. It is important to notice that the inclusion
of the far wake model allows to reduce the intermediate
wake portion with small effects on both thrust and torque
predictions, whereas if the far wake is not included, the
rate of variation of calculated loads with the intermediate
wake lenght is higher.
The effect of the far wake disc distance dD from the
intermediate wake and the effect of grid refinement on the
far wake disc are also studied. Tables 1 and 2 show the ef-

Figure 7: Influence of intermediate wake lenght LIW on


calculated loads by using wake alignment with fixed near
wake lenght LN W = 2.0D. Results with and without
far wake disc compared. Top: thrust coefficient, bottom:
torque coefficient.

fect of these parameters on predicted loads. As expected,


the influence on KT and KQ is almost negligible, in that
far wake contributions are given by a constant sources distribution, as indicated by Eq. (16). Also the far wake disc
radius is found to have a negligible influence, if its value
is slightly varied with respect to the radius of the wake
streamtube of the intermediate wake portion (result not
shown here).
As already mentioned, a crucial feature of the present
wake alignment approach is the introduction of a vortex
core region to avoid infinite velocity induced by wake vortices. Top Fig. 8 shows the location at x/R = 0.65 of
the calculated trailing wake surface by using different values of the vortex core radius. It is apparent that too large
vortex core values tend to inhibit the wake rollup process (black curve in the figure), whereas too small vortex

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

No far wake disc


dD /D = 0.001
dD /D = 0.075
dD /D = 0.175
dD /D = 0.275

KT
0.1483
0.1476
0.1476
0.1477
0.1477

10KQ
0.2878
0.2867
0.2868
0.2869
0.2869

Table 1: Effect of far wake disc location on predicted


thrust and torque coefficients.

No far wake disc


Coarse grid
Medium grid
Fine grid

KT
0.1483
0.1476
0.1476
0.1476

10KQ
0.2878
0.2868
0.2868
0.2868

surfaces and an infinite number of wake turns are considered (see extrapolating polynomials in Figs. 6). Specifically, the MB = NB = 18 and MW = 90 discretization
combined with far disc model yields KT predictions that
differ about 1.3% from the extrapolated values. Finally,
the verification task yields an adequate procedure to determine the vortex core radius value. In particular, results
in Fig. 9 prove that the estimated boundary layer thickness falls in the range of r values that allow both stable
and resolved wake rollup. For the particular case considered here, J = 0.88, r = 1.5 102 is the boundary layer
thickness value that is obtained from the boundary layer
analysis. Such value is obtained by combining suction
and pressure sides thicknesses at the blade trailing edge
and averaging along span. Similarly, an estimated value
of the growth factor r is obtained.

Table 2: Effect of far wake disc discretiation on predicted


thrust and torque coefficients.
core regions can determine highly distorted surface shapes
that tend to unstabilize the wake alignment process (red
curve). This is confirmed by observing the influence of
the vortex core radius on the wake convergence history, as
shown in bottom Fig. 8. Nevertheless, predicted propeller
thrust and torque are scarcely sensitive to variations of the
vortex core radius except for values close to the stability
limits.
The combination of the wake alignement technique
with the boundary layer analysis allows to determine a
physicallyconsistent value of the vortex core radius. Figure 9 shows the predicted thickness of the boundary
layer on the blade surface and on the near wake portion
close to the blade trailing edge. Distributions of the displacement thickness and of the friction coefficient Cf
are also reported for completeness (subscript denotes
quantities evaluated along the curvilinear direction introduced above). It should be observed that results in Fig. 9
are determined by using a prescribed helicoidal wake. The
influence of the wake shape on such distributions is found
negligible.
The results discussed in this section are useful to establish the most appropriate set of parameters that are necessary to run performance and flowfield predictions. In
particular, Figs. 6 and 7 clearly show that two wake turns,
1.5 of near wake plus 0.5 of intermediate wake, and far
wake disc included, can be considered as a good compromise between accuracy and computational effort (condition corresponding to ratio LIW /LN W = 0.33 in Figs. 6
and 7). Moreover, Fig. 2 yields that MB = NB = 18
and MW = 90 should not be considered as a sufficiently
fine grid. However, the inclusion of the far wake model reduces the discrepancy of calculated thrust and torque coefficients with respect to values that may be extrapolated in
the limit as an infinite number of panels on blade and wake

Figure 8: Effect of the vortex core radius on the shape


of the wake intersection with transversal plane at x/R =
0.65 (top) and on the convergence history index RW (bottom).

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Figure 9: Predicted boundary layer quantities on blade and near wake, pressure side (label: Back) and suction side (label:
Face), at J = 0.88, ReD = 1.5 106 . Top: thickness; center: displacement thickness, bottom: friction coefficient.

VALIDATION OF NUMERICAL RESULTS


The capability of the proposed theoretical methodology
to predict the propeller slipstream flow is investigated by
comparing numerical results with experimental data. The
analysis of the perturbation velocity field downstream a
propeller has been considered by using an early version
of the present approach without viscosity correction and
far wake modelling, Giordani et al. (1999), and more recently by Salvatore et al. (2003), where only the viscous
flow correction was included. The enphasis in the present
work is on the prediction of the trailing vorticity pattern
in the slipstream and on the capability of the present far
wake model to improve cost/effectiveness of the propeller
perfomance numerical predictions.
For this validation exercise the INSEAN E779A
model propeller has been considered. It is a skewed four
bladed propeller with diameter D = 0.227 m, pitch-todiameter ratio of 1.1 and forward rake angle of 4 300 . This
propeller has been tested at INSEAN for several years and

a thorough characterization of the propeller flow by using


LDV and PIV techniques is available. Descriptions of the
experimental activity on this propeller are given, e.g. , in
Di Felice et al. (2000), Stella et al. (2000) and Di Florio
et al. (2004). A limited portion of experimental data as
well as the E779A model propeller geometry details have
been collected into an experimental database available to
the public domain at http://crm.insean.it/E779A.
Consider first the propeller working at design conditions (J = 0.88). Figure 10 shows the comparison between numerical computations and experiments for the
evaluation of the tip vortex radial position. The blade
profile is also depicted for reference. The comparison
highlights some inaccuracy in the prediction of the slipstream contraction at x/D < 1, whether a good agreement is achieved at higher distance. The wavy behaviour
of the numerical prediction can be explained considering
the rolling up of the wake sheet and the non univocal definition of the numerical tip vortex. In the experiments,
the tip vortex is identified by turbulence level peaks in the

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

flowfield, whereas in the calculations its position is identified with the outmost streamwise wake node line.
The detailed characterization of the slipstream flow
by means of LDV and PIV techniques provides a powerful means to investigate the accuracy of the wake flow
calculations. To this purpose, measured velocity in both
longitudinal and transversal planes are processed to determine the trailing wake pattern. The vorticity may be
evaluated by finite differentiation of the pointwise velocity
measurements whereas turbulence intensity is determined
as the standard deviation of axial velocity samples (root
mean square). In the flow region of interest here, the two
approaches provide comparable locations of the trailing
wake pattern.
Experimental data considered hereafter represent turbulence intensity levels of the axial velocity. Such data
processing is used in Fig. 11 to depict the intersection of the wake with transversal planes at different distances from the propeller disc plane. Specifically, flow
conditions at J = 0.88 and four planes at x/R =
0.2, 0.65, 1.15, 1.65 are considered. Experimental data
are compared to the locations of the calculated trailing
wake surface on the same planes. The pictures show a
good agreement between measured and predicted wake
locations. In particular, both wake pitch distribution in
the radial direction, wake rollup at the tip and slipstream
contraction are fully captured at small distance from the
propeller. Some minor discrepancies appear in the two
planes far downstream, x/R = 1.15, 1.65.
This analysis is further investigated considering the
wake flow in a longitudinal axial plane, as shown in
Fig. 12. In this case, three different flow conditions are
given: J = 0.748, 0.88, 1.012. Results for the same value
of J considered in Fig. 11 above (mid picture), reveal that
the prediction of the trailing wake pattern is good not only
close to the propeller but also at more than two diameters
downstream, througout the flow region where experimental data are available (0 < x/R < 4.6). For completeness, the whole wake surface explicitely considered in the
calculations is given, in order to appreciate the increasing
rollup of wake sections moving from the blade trailing
edge to downstream. This geometrical trend is associated
with vorticity concentration towards the tip, and reflects a
physical feature that is evidenced by the size and intensity
of the measured tipvortex core. It is worthwhile to notice
that at large x, the observed vortical field may be correctly
approximated by the semiinfinite cylindrical model that
has been proposed above to introduce the far wake disc
concept.
Similar conclusions may be drawn by analysing results for flow conditions at J = 0.748 and 1.012, in
Fig. 12. For these two cases, the flow region where experimental data are available is limited between x/R = 0
and 2.8. In the case J = 0.748, corresponding to higher

than design loading conditions, the agreement between


measured and predicted vorticity patterns is good whereas
some discrepancies occour in the case of light propeller
loading at J = 1.012. This latter result is partly unexpected and requires further investigations.
A limit of the present wake flow model that is apparent from Figs. 11 and 12 is the description of the hub
vortex flow region. Experimental data in that area are not
available due to lack of optical access. However, the calculated wake shape presents an unphysical trend with formation of hub vortices that translate parallel to the propeller axis. This result can be explained recalling that
hub vortices interact each other into a thick viscous wake
emanating from the propeller hub. In this region, turbulence and viscosity diffusion effects that are missed in the
present model are relevant.
Finally, the influence of wake modelling on the propeller loads evaluation is considered in Table 3. Measured
data corresponding to the three values of the advance coefficient considered above are compared to predicted values
by using either a prescribed helicoidal wake (P.W.) and
the flowaligned wake model here proposed (F.W.). As
expected, an improvement of calculated thrust and torque
coefficients is generally achieved if the flowaligned wake
model is considered.
J
KT , Measured
KT , Num. F.W.
KT , Num. P.W.
10KQ , Measured
10KQ , Num. F.W.
10KQ , Num. P.W.

0.748
.2285
.2182
.2306
.4209
.3818
.3997

0.880
.1615
.1670
.1721
.3182
.3133
.3211

1.012
.0973
.1116
.1123
.2141
.2366
.2381

Table 3: Measured thrust and torque coefficients compared to calculations using a prescribed wake model
(P.W.) and a flowaligned wake model (F.W.). INSEAN
E779A model propeller.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Figure 10: Radial location of the INSEAN E779A model propeller wake tipvortex at J = 0.88. Numerical predictions
compared to experimental data.

Figure 11: Measured axial velocity turbulence levels in the propeller slipstream compared to numerical predictions of the
trailing wake path (dashed black lines). INSEAN E779A model propeller at J = 0.88. Transversal planes at: x/R = 0.2
(top left), x/R = 0.65 (bottom left), x/R = 1.15 (top right), and x/R = 1.65 (bottom right).

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Figure 12: Measured axial velocity turbulence levels in a longitudinal axial plane downstream the propeller compared to
numerical predictions of the trailing wake path (dashed black lines). INSEAN E779A model propeller at J = 0.748 (top),
J = 0.88 (center), J = 1.012 (bottom).

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

CONCLUSIONS
A theoretical model for the analysis of marine propeller
wake flow by a nonlinear boundary integral formulation
for the velocity potential has been presented. Wake flow
modelling is performed by assuming that the trailing vorticity path is part of the flowfield solution. Contributions
from the vorticity field far downstream the propeller and
viscous boundary layer effects to the potential field are
taken into account through physically consistent models.
Numerical results by the proposed methodology are
investigated in order to evaluate the numerical scheme
performance and the capability to describe propeller wake
flow features. The analysis of predicted vorticity pattern
in the wake demonstrates that the trailing vorticity shedding process is accurately represented and global features
as the wake pitch distribution and the slipstream contraction are in good agreement with measured data in a large
flow region extending twothree diameters downstream
the propeller disc plane. Numerical results are validated
over a range of variation of the advance coefficient below
and above design conditions. The improvement of propeller loads calculations by using wake alignment is also
analysed against experimental data.
It may be concluded that the proposed methodology
represents an appealing tool for the analysis of isolated
propellers wake flow. The reduced computational effort
allows to apply this methodology to the analysis of propeller wake interactions with other components in operating conditions.
Further activity on the proposed methodology is
aimed to improve the numerical solution of the three
dimensional boundary layer equations. The extension of
the wake alignment procedure to the analysis of unsteady
flow conditions and additional validation of numerical results is also underway.

REFERENCES
Batchelor, G.K. (1967), An Introduction to Fluid Dyn
amics, Cambridge University Press, Cambridge, U.K.
Campbell, R.G. (1973), Foundations of Fluid Flow
Theory, Addison-Wesley, Reading, Massachusetts.
Di Felice, F. and Romano, G. P. and Elefante, M.
(2000), Propeller Wake Evolution by Means of PIV.
Proceedings of the 23rd ONR Symposium on Naval
Hydrodynamics, Val de Reuil (France).
Di Florio, D., Di Felice, F., Felli, F. and Romano, G.P.
(2004), Experimental Investigation of the Propeller Wake
at Different Loading Condition by PIV, to appear in
Journal of ship Research.
Gennaretti, M., Corsetti, E., and Morino, L. (1998),

Coupled
Free-Wake-Aerodynamics/Blade-Dynamics
Analyses of Rotors in Forward Flight, AIAA Paper
98-2241,
4th AIAA/CEAS Aeroacoustic Conference,
Toulouse, France.
Giordani, A., Salvatore, F. and Esposito, P.G. (1999),
Free Wake Analysis of a Marine Propeller in Uniform
Flow, XXI World Conference on BEM, Oxford, U.K.
Greeley, D.S. and Kerwin, J.E. (1982), Numerical Methods for Propeller Design and Analysis in Steady Flow,
SNAME Transactions, Vol. 90, pp. 416453.
Green, J.E., Weeks, D.J. and Brooman, W.F. (1973), Prediction of Turbulent Boundary Layers and Wakes in Compressible Flow by a LagEntrainment Method. A.R.C.R.
& M. Tech. Rep. 3791.
Hoshino, T. (1989), Hydrodynamic Analysis of Propellers in Steady Flow Using a Surface Panel Method,
2nd Report: Flow Field Around Propeller, Journal
of The Society of Naval Architects of Japan, Vol. 166,
pp. 7992.
Jessup, S.D. (1989), An experimental investigation of
viscous aspects of propeller blade flows, PhD thesis, The
Catholic Univ. of America.
Johnston, J.P., (1957), Three Dimensional turbulent
Boundary Layers. Doctoral thesis, M.I.T., Massachussetts (USA).
Kawakita, C. (1992), A Surface Panel Method for
Ducted Propellers With New Wake Model Based
on Velocity Measurement, Journal of the Society
of Naval Architects of Japan, Vol. 172, pp. 187203.
Kerwin, J.E. and Lee, C.S. (1978), Prediction of Steady
and Unsteady Marine Propeller Performance by Numerical Lifting-Surface Theory, SNAME Transactions,
Vol. 86.
Kinnas, S.A. and Pyo, S. (1999), Cavitating Propeller
Analysis Including the Effects of Wake Alignment,
Journal of Ship Research, Vol. 43, No. 1, pp. 3847.
Lighthill, M.J. (1958), On displacement thickness.
Journal of Fluid Mechanics, Vol. 4, pp. 383392.
Liu, P. and Colbourne, B. (2002), A Study of Wake Discretization in Relation to the Performance of a Propeller
Panel Method, Ocean Engineering International Journal,
2002.
Michel, R. (1952), Etude de la Transition sur les Profils
daile - Establissment dun Point de Transition et Calcul
de la Trainee de Profil en Incompressible. ONERA Tech.
Rep. 1/1578, France.
Morino, L, Chen, L.T. and Suciu, E. (1975), Steady
and Oscillatory Subsonic and Supersonic Aerodynam-

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

ics Around Complex Configurations, AIAA Journal, Vol.


13, pp. 368374.
Morino, L., Salvatore, F. and Gennaretti, M. (1997),
A New Velocity Decomposition for Viscous Flows:
Lighthills EquivalentSource Method Revisited.
Computer Methods in Applied Mechanics and Eng., Vol.
173, No. 34, pp. 317336.
Nishida, B.A., (1996), Fully Simultaneous Coupling of
the Full Potential Equations and the Integrated Boundary
Layer Equations in Three Dimensions. Doctoral thesis,
M.I.T., Massachussetts (USA).
Salvatore, F. and Testa, C. (2002). Theoretical Modelling of Marine Propeller Cavitation in Unsteady HighReynolds Number Flows. INSEAN Report 2002-077,
Rome (Italy).
Salvatore, F., Testa, C. and Greco, L. (2003), A
Viscous/Inviscid Coupled Formulation for Unsteady
Sheet Cavitation Modelling of Marine Propellers, Fifth
International Symposium on Cavitation, Osaka (Japan).
Stella, A. and Guj, G. and Di Felice, F. (2000), Propeller
Flow Field Analysis by Means of LDV Phase Sampling
Techniques. Experiments in Fluids, Vol. 28, pp. 110.
Suciu E. and Morino L. (1978) Non Linear Steady Incompressible Lifting-Surface Analysis with Wake Rollup, AIAA Journal, Vol. 15, No. 1.
Thwaites, B. (1949), Approximate Calculation of the
Laminar Boundary Layer. Aeronaut. Quart., Vol. 1, pp.
245280.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Ismail B. Celik
West Virginia University, USA
While the authors were trying to calibrate
their vortex potential core model they showed
experimental data from LDV and PIV measurements.
The peak velocity in these measurements showed
about 30% difference between the two techniques.
The numerical modelers (I am one of them) would
have tried to match either of the two data sets 100%.
Could the experimentalist author comment first on
such large differences, next on how they can warn the
numerical modelers that not all experimental data are
perfect!
AUTHORS REPLY
Prof. Celik is referring to results that were
shown at the symposium and are here included in the
reply to Dr. P. Liu, Fig. 4.
The figure shows comparison of BEM
calculations with PIV and LDV data of the velocity
field at x/R=1.15. Experimental data show important
differences in the estimation of the velocity defect of
the blade wake. This should always be expected
because every measurement techniques has
limitations. In the present case, such behavior is due
to the different characteristics of the measurement
volume of LDV and PIV. In fact LDV measurement
volume size is of the order of 0.1x0.1x4.0 mm3
whereas PIV measurement volume, in the present
case, is 100 times larger (2.5x2.5x1.5 mm3 ). In such
a case PIV is not able to resolve the thin blade wake
and is providing a mean value in the measurement
volume smoothing the blade wake velocity defect.
A detailed report of the PIV and
LDVperformance in measuring the E779A model
propeller wake is given in Felli et al. (2002).
REFERENCES
Felli, M., Di Florio, D., Di Felice, F. (2002),
Comparison between PIV and LDV technique in the
analysis of propeller wake. Journal of Visualization
Vol. 5, n 3, 2002, pp. 209-210.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Ki-Han Kim
Office of Naval Research, USA
The discrepancy between predicted and
measured KT and KQ at off-design conditions is
somewhat higher than that by other potential flow
prediction methods, including lifting-surface method.
Can the authors comment on the possible causes of
the discrepancy? How did the authors treat the blade
frictional drag?
AUTHORS REPLY
The issue of loads prediction is only briefly
described in the paper because enphasis is posed on
the validation of slipstream flow predictions.
For this purpose, the INSEAN E779A model
propeller was selected as a test case due to the vast
velocity field dataset available. At present, propeller
loads at only three different values of the advance
coefficient, J = 0.748, 0.88, 1.012, are available from
cavitation tunnel measurements. The comparison
between measurements and numerical results (paper
Table 3) shows that predicted thrust and torque are in
good agreement with measurements close to design
conditions, J = 0.88. At lower J, predictions
underestimate both thrust and torque.
This trend is partly confirmed by
considering another test-case, the DTRC 4119 model
propeller tested by Jessup (1989). Predicted and
measured thrust and torque coefficients are compared
in Fig. 1 below.

to a lack of accuracy of pressure calculations in the


blade trailing edge region. A simple Kutta condition
is used whose limits to address highly-loaded bodies
with strong crossflow effects at the trailing edge are
known from the literature. As a consequence, a nonzero pressure jump may occour and local blade load
contributions are underestimated. The alignment of
the wake surface to the local flow does not contribute
to alleviate this problem. Thrust and torque
predictions are also affected by an approximated
evaluation of viscous-flow effects. The PrandtlSchlichting formula for flat plate frictional drag is
used to correct inviscid loads obtained by integrating
pressure over the propeller surface. Different
formulas have been tested with very similar results.
Close to design conditions, results by the PrandtlSchlichting formula have been validated through
comparisons with skin friction distributions predicted
by solving the boundary layer flow (see paper for
details). The assessment of this technique for low J
conditions was beyond the limits of the present work.

REFERENCES
Jessup, S.D. (1989),
"An experimental investigation of viscous aspects of
propeller blade flows," PhD thesis, The Catholic
Univ. of America, Washington, DC, USA.

Figure 1

A careful analysis of present results yields


that load prediction errors should be mostly attributed

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

too big to produce large enough induced


velocities compared with experiment data. Was
the dimensional analysis performed in the cut-off
radius formulation to ensure the same propeller
geometry with a different diameter (e.g., D=0.3m
versus D=6m) to have about the same amount of
non-dimensional
induced
velocities?
Downstream velocity prediction becomes
important for offshore engineering. In an earlier
study (5), it showed that the axial velocity still
remains at a 50% of magnitude at 10 diameters
downstream which shows a great possibility of
damage caused by propeller wake momentum to
structures far downstream.
Modeling of the wake was done under
lightly loaded conditions, i.e. with hydrodynamic
angle of attack related ratios, J/[p/D]=0.68, 0.88
and 1.1, respectively. How does the wake model
behave under heavily loaded conditions, say,
J/[p/D]=0.0~0.5?
What is the capability of the method for
predictions of propulsive forces, especially for a
propeller with a large pitch to diameter ratio at
low J (say p/D=1.4 at J=0.0)? It would be nice to
see a plot for Kt and Kq for the two different
wake models in a wide range of advance
coefficient J from 0.0 to J=[p/D]. Did the wake
alignment model improve the pressure
coefficient prediction at the trailing edge as well
and how much?

DISCUSSION
Pengfei Liu
National Research Council, Canada
This paper showed an excellent combination
of numerical and physical modeling studies
conducted at INSEAN.
In the paper the authors used the Biot-Savart
law to find induced velocity in the process of realigning the wake vortex sheet. From our experience,
the predicted induced velocities are very sensitive to
the vortex filament core radius cut-off value. The
effect of this value on the induced velocities is much
greater than on the wake roll up geometry, though a
good prediction of induced velocities also depends
largely on the wake pitch modeling. In our previous
work, a modified cut-off value formulation (1) based
on the work by Maskew (2) was determined by
geometrical relation between a vortex filament and a
field point. Similar works on induced velocities
downstream of a propeller can also be found in the
proceedings of the 22nd ITTC RANS/Panel Method
Workshop (3).
In the paper, a unique equation of vortex
core radius cut-off value (equation 10) was
established. The value is a function of boundary layer
thickness which links physical phenomena to a
potential flow solution procedure which is reasonable.
Although with a tight space limit of a conference
paper it is impossible to address the issues and
present all the major findings, it would be great if the
authors could address the following concerns:

REFERENCES
1.

The detailed derivation of equation (10) was


not presented. It also cannot be found from
publications listed in the reference section. Could
the authors show how this equation is linked
with the boundary layer thickness and how it is
implemented in terms of numerical schemes? In
detail, why they used the square root and what
are the physical meanings of the independent
variables and their ranges of the values?
The numerical prediction of the wake shape
geometry indicates a good agreement with the
measured ones but the paper did not show a
comparison between the predicted magnitude of
the induced velocities and the measured ones.
How well their predicted magnitude of the
circumferential averaged velocities compares
with measurements, e.g., by using an LDA
device (4)? In the paper, the authors used the
radius cut-off value of around 1.5E-02. What is
the unit of the value? If it is 1.5E-02D, where D
is the diameter of the propeller, it is then 3.405E03 m. This value in our previous work (1) was

2.

3.

4.

5.

P. Liu, N. Bose, and B. Colbourne, Automated


marine propeller geometry generation of
arbitrary configurations and a wake model for far
field momentum prediction, International
Shipbuilding Progress, Vol. 48, No. 4, 2001, pp.
351-381.
B. Maskew, Program VSEARO Theory
Document, NASA Contactor Report 4023,
Ames Research Center Contract No. NSA211945, 1987, 100p.
The 22nd ITTC Propulsion Committee,
Proceedings of the 22nd ITTC RANS/Panel
Method Propeller Workshop, Grenoble, France,
1998.
S.D. Jessup, An Experimental Investigation of
Viscous Aspects of Propeller Blade Flow, PhD
Thesis, The Catholic University of America,
1989, 249p.
U. Nienhuis, Analysis of Thruster Effectively
for Dynamic Positioning and Low Speed
Maneuvering,
PhD
Thesis,
Technical
University of Delft, Netherlands, 1992, 191p.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

AUTHORS' REPLY
The authors wish to thank Dr. P. Liu for his
valuable comments and suggestions for further work.
Comments highlight some basic issues on propeller
slipstream flow predictions by BEM.
First, the proposed vortex-core model is
addressed. The present approach stems from defining
the vortex-core as the flow region surrounding each
vortex filament where the induced velocity departs
from Biot-Savart law for inviscid flows. Close to the
vortex, viscosity effects are not negligible and a
velocity defect is present. Equation (10) in the paper
provides a simple model of the vortex-core diameter
growth along a wake streamline, as proposed by
Morino and Bharadvaj (1985). This two-parameter
law provides a realistic description of the viscous
wake thickness downstream the blade trailing edge.
This is illustrated in Fig. 1 here below, where
quantity r from paper Eq. (10) is compared with
predicted boundary layer thickness distribution in a
limited wake portion close to the blade (see paper
Fig. 9). Flow conditions in Fig. 1 are J = 0.88,
ReD = 1.5E6. In particular, the vortex-core radius at

deep velocity defect across the angular location


= 50 and amplitude 8. Recalling r/R = 0.7, this
corresponds to a thickness of about 0.05D, in
qualitative agreement with thickness values from Eq.
(10) shown in Fig. 1.
The impact of the imposed vortex-core size
on velocity predictions is sketched in Fig. 3 here
below.

Figure 2
Blade T.E.

Figure 1

the blade trailing edge is choosen as twice the


calculated boundary layer thickness. All quantities in
Fig. 1 are adimensionalized with respect to the
propeller diameter.
A unique definition of the boundary layer
thickness is obtained here by using the relationship
among , displacement thickness * and momentum
thickness proposed by Green et al. (1973). It should
be observed that the relationships between vortexcore size and boundary layer thickness from direct
solution of the boundary layer flow was studied only
for the case J = 0.88. The estimated vortex-core size
may be compared with the viscous-wake thickness as
it is observed from velocity field measurements. A
representative result is depicted in Fig. 2, where
measured induced axial velocity at r/R = 0.70,
x/R = 0.2 is compared to numerical predictions (case
J = 0.88). The viscous wake location is stressed by a

Figure 3

Specifically, the effect of growth factor r


is considered. Comparisons between predicted and
measured velocity show that too high values of this
parameter determine a loss of accuracy of velocity
predictions. Conversely, too small values of r
yield unstable wake-alignment calculations (paper
Fig. 8).

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Due to lack of space, results of propellerinduced velocity predictions were not included in the
paper. Nevertheless, some results were shown during
the conference presentation and are given here in Fig.
4 for the measurement plane at x/R = 1.15.

Figure 4

Finally, Dr. Liu addresses blade pressure


and propeller performance predictions. These issues
are discussed in the authors' reply to Dr. K.-H. Kim.

REFERENCES
Morino, L. and Bharadvaj, B.K. (1985),
"Two Methods for Viscous and Inviscid Free Wake
Analysis of Helicopter Rotors," Report CCAD-TR
85-02-R, Boston University, Boston, MA, USA.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics


St. Johns, Newfoundland and Labrador, CANADA, 8-13 August 2004

Experimental Analysis of the Wake from a Dynamic


Positioning Thruster
S. El Lababidy1, N. Bose1, P. Liu2, F. Di Felice3, M. Felli3, F. Pereira3
(1Memorial University of Newfoundland, Canada; 2Institute for Ocean Technology
National Research Council Canada; 3Italian Ship Model Basin (INSEAN), Italy)
ABSTRACT

Greek Symbols
- : fluid density, Kg/m3
- : flow vorticity (s-1)
- : propeller open water efficiency,
() = (J.KT) / (2.KQ)

The knowledge of the effects of the


wakes from Dynamic Positioning (DP) thrusters
and the practical need to describe and predict
their impact is fundamental in the assessment of
design loading on slipstream-submerged
equipment associated with offshore floating
systems. In particular, with the increasing
importance and application of DP thrusters in the
offshore industry, the designer as well as the
operator must have more information about the
momentum in the wake from the thruster.
This paper presents wake measurements
from cavitation tunnel tests carried out on a
ducted propeller model at the Italian Ship Model
Basin (INSEAN), Rome, Italy. Through these
experiments, DP thruster wake velocity
components at different downstream axial planes
were obtained experimentally using a Stereo
Particle Image Velocimetry (SPIV) system in the
INSEAN cavitation tunnel, up to 15 diameters
downstream. These experiments were carried out
with and without the nozzle in the bollard pull
condition.
This paper presents the results from
these tests and discusses their importance for
structures placed in the slipstream.

INTRODUCTION
Application of Dynamic Positioning
(DP) systems is increasing in the offshore oil and
gas industry. As operations move into more
hostile and deeper waters, the need for DP
system has become of greater and greater
importance. There are different types of thrusters
used in DP systems, but the majority are
steerable over 3600 and mounted below the hull,
or placed in tunnels crossing the hull from one
side to the other (Fay, 1990). A historical review
of the application and of different types of DP
systems is provided by Morgan (1978). DP
systems have been available since the early
1960s, but they have become in much more
common use since the exploration and
development of offshore fields moved into deep
waters. Thrusters working at low advance ratio
are employed in a number of marine and
offshore DP systems, on FPSO (floating,
production, storage and offloading) systems,
shuttle tankers, and mobile offshore units.
Information on the hydrodynamic characteristics
of these thrusters is still limited. More studies are
required in order to gain knowledge on and
prove the reliability of such systems.
The tests described here were done to
obtain experimental information on the flow
field in the wake of these thrusters. The main
objective was to analyze the DP thruster near and
far wake experimentally when operating with
and without a nozzle under variable operating
conditions.

NOMENCLATURE
-

D: propeller diameter, m
J: advance coefficient, (J) = VA /(n.D)
KQ: torque coefficient, (KQ) = Q / (. n2.D5)
KT: thrust coefficient, (KT) = T / (. n2.D4)
n: angular speed of propeller shaft, rev/s
p: propeller pitch, m
Q: torque, N.m
T: thrust, N
VA: velocity of advance, m/s
X: axial downstream distance, m

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

cameras to view the flow from two perspectives.


The two images for a point in the flow are
combined to yield a three-dimensional velocity
vector. By combining the vector fields from the
two cameras, the three-dimensional velocity field
over a plane is computed (Calcagno et al. 2002).
The effects of a DP thrusters far wake
on structures in the slipstream were mentioned
by Liu et al. (2001) in their numerical work to
investigate the near and far wake, up to five
diameters downstream. The analysis was done
using the propeller panel code PROPELLA.
Nienhuis (1992) was probably the first person to
study the performance of a DP thruster during
tracking and low speed maneuvers. In his work,
thruster-thruster and thruster-hull interaction
were studied numerically and experimentally
using a 2D LDV system. Recently, El Lababidy
et al. (2004) measured the near wake flow of a
DP thruster model, when operating with and
without a nozzle at an advance coefficient, J, of
0.4, using a 2D LDV system in a cavitation
tunnel. The results of these experiments, with the
same propeller model as used here, indicated that
there was greater axial momentum and mass
defect from the wake for the open propeller than
for the ducted propeller.
DP thrusters are used in highly loaded
propeller conditions. Experiments should be
carried out at bollard pull (J = 0) and low J
operating conditions. We know of no research to
date that presents, describes and analyzes the
hydrodynamic characteristics of a DP thruster
wake when operating with and without a nozzle
in the bollard pull condition. In the present work
Stereo PIV (SPIV) was used to evaluate the
wake characteristics of a DP thruster model
when operating with and without a nozzle in the
bollard pull condition.

The flow in the wake of a DP thruster


involves several complex features, transition
zones, turbulence, presence of vortical structures,
deformation, effects of waves and currents,
vortex rollup and break down, shear layers, etc.
These do not easily allow a complete evaluation
of the flow features. However, experimental
measurements become partly feasible through
the introduction of the non-intrusive laser
techniques (laser Doppler velocimetry LDV and
particle image velocimetry PIV). LDV is a single
point measurement technique, which has been
widely used to evaluate complex flow fields
including propeller flows (Min 1978; Hoshino &
Oshima 1987; Jessup 1989; Stella et al. 2000;
Chesnakas & Jessup 1998, El Lababidy et al.
2004). However, as with any experimental
technique, LDV has some limitations:
- As a single point measurement technique, it
cannot give an overall picture of the
characteristics of large coherent structures,
which are generally encountered in complex
and separated flows.
- The fixed location and time-averaged nature
of the LDV technique can induce significant
errors in the measurement of the intensity of
unsteady vortical structures.
- It is a time consuming technique, as multiple
measurements must be taken to construct a
whole velocity field. This leads to increased
cost of experiments and difficulties in the
investigation of unsteady flows.
A sufficiently dense grid of LDV
measurements is required in order to enable it to
resolve the flow structure during a propeller
revolution in a non-uniform inflow (Felli et al.
2000, Di Felice et al. 2000). From this point of
view, the PIV technique that can take
simultaneous measurements over a given plane
in the fluid, offers many advantages over the
LDV technique. So, experimental measurements
can be fast and can be easily conducted by
acquiring images at each angular position of
blade (Felli et al. 2002). Since the initiation of
PIV techniques, significant progress and
improvements to the system have been made to
broaden its range of application, increasing the
capability and reliability of the measurement
approach. Nowadays the PIV technique is
considered to be a powerful whole field
measurement technique.
Stereoscopic PIV is an obvious
measuring technique for complex propeller flow
fields. It allows the construction of the three
components of velocity fields by using two

EXPERIMENTAL SET-UP
Tests were done at INSEAN, the Italian
Ship Model Basin, Rome, Italy, in their large
cavitation tunnel. This is a free surface channel
with a test section of 10 m in length, 3.6 m in
width and 2.25 m in depth. The maximum
allowable water speed is 5.2 m/s.
Tests were carried out at a propeller
pitch ratio (P/D) of 1.2 and at propeller speed of
20 rps. The propeller model was a ducted
controllable pitch propeller. The principal
dimensions and particulars of the propeller, the
model and the nozzle are given in Table (1)
below (Doucet, 1996).

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

configuration have been previously assessed on a


test bench (Felli 2003) and are about 3% for the
light sheet in plane components and about 2%
for the out of plane components.

Propeller Particulars
Number of Blades
4
Diameter
3.0 m
Hub to Propeller Diameter Ratio
0.367
Propeller Model Particulars
Number of Blades
4
Diameter
0.2 m
Expanded Area Ratio (EAR)
0.604
Nozzle Model Particulars
Length
0.100 m
Inside Diameter
0.202 m
Nozzle Contraction Ratio
1.2223
Table 1: Ducted propeller model and model
nozzle particulars (Doucet, 1996)

Figure 1: DP thruster model and SPIV in the


INSEAN large cavitation tunnel

Measurements were performed using an


underwater stereo-PIV probe. Fig. (1) shows the
probe and the thruster installed in the test
section. The probe, when completely assembled,
forms a streamlined torpedo-like tube with an
external diameter of 150 mm. The tube was
rigidly linked to a bench through two
hydrodynamically optimized struts. The whole
system could be traversed by a mechanism
controlled by a PC to sweep the measurement
plane up to X/D = 15. The stereoscopic system
consisted of two 2048 X 2048 pixels CCD
cameras, with a 12-bit resolution. The aperture,
the focus and the Scheimpflug angle of the
camera lenses were remotely controlled. The
laser light was delivered to the underwater sheet
optics through the struts and inside an articulated
arm. Mirrors within the arm allowed the beam to
be correctly driven to the output optics. The laser
optics consisted of a set of cylindrical and
spherical lenses, which respectively expanded
the beam into a sheet and focused it onto the
measurement plane. The laser subsystem
consisted of a pulse-doubled 15 Hz 200 mJ NdYAG laser rigidly attached to the probe support
bench. The camera mirror sections were open to
the water to avoid multimedia refractions and
hence to minimize the optical aberrations
through one single orthogonal water-air
interface. More details on the underwater SPIV
probe have been presented by Felli (2003).
The calibration process for the stereo
PIV system was done using 2D calibration plate.
The overlapped imaged areas of the two cameras
allow the measurement of the 3D velocity field
in an area of about 328 mm X 265 mm. The
measurement errors for the adopted optical

CO-ORDINATE SYSTEM
A Cartesian (0-x, y, z) co-ordinate
system was used. As shown in Fig. (2), the x-axis
is along the propeller shaft centerline and points
downstream, the y-axis is horizontal, and the zaxis is vertical pointing upwards. As shown in
Fig. (2) the longitudinal velocity is U and the
cross-flow velocity components are V and W.
W

y
U

x
Figure 2: Coordinate system
The mesh of the propeller and the Nozzle were
generated using PROPELLA (Liu and Bose,
1998; and Liu, 2002)
The analysis of wake velocity was done
firstly by constructing the instantaneous 3D
velocity vector field for each image (129
images/plane) from the combination of the 2D
encoded velocity vector fields of the left and
right camera. Then the mean 3D velocity field

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

velocity. Detected erroneous vectors are


eliminated and replaced by interpolation.
Nevertheless, spurious vectors might also be
validated biasing the statistics especially in the
proximity of the hub where the flow and
propeller velocity become closer. This effect is
relevant especially for the evaluation of second
order statistics.
The accuracy of the mean velocity field
depends on the number of acquired samples and
on the shape of the velocity probability
distribution function. By using the Student tdistribution (for which the confidence interval
for 95%, is 1.96*RMS /(N-1), with N = 129),
it is possible to estimate the uncertainty in a
velocity component to be about 1/6 of the
measured RMS velocity.

was constructed for each plane by averaging the


summation of the instantaneous 3D vector files
for each plane over 129.
MEASUREMENT UNCERTAINTY
A comprehensive discussion on the
uncertainty and the accuracy of the PIV
technique, especially for SPIV, is outside the
goals of the present work and this aspect is a
complex topic with many unresolved issues. The
technique used for the stereo reconstruction was
the method described by Solof (1997) and a
detailed analysis of the errors was presented by
Prasad (2000).
The
uncertainty
of
velocity
measurements from each single camera is mainly
due to the error in the particle displacement
evaluation, which can be normally considered to
be around 1/10 to 1/20 of a pixel for the image
analysis and subpixel interpolation algorithms
(Raffel et al., 1998). This was equivalent to
approximately 3 cm/s in terms of velocity for the
flow speeds considered here. This error is
present in the measurement of the instantaneous
flow field seen by each camera, but with the
stereo reconstruction the error increases and is up
to 10 cm/s for the longitudinal velocity
component.
This error, which is relatively large, is a
typical value for a SPIV measurement and
depends on many factors such as the optical
configuration (angle between the cameras),
optical aberrations, number of dots in the
calibration target, number of planes used in the
calibration, etc. The errors due to light
reflections from the hub and from the blade
edges and the nozzle were also important in flow
field regions mapped in proximity to reflection
spots.
The moving propeller in the background
of the measurement plane is another source of
error. To reduce these errors the images where
preprocessed in order to remove the background.
However the effect of the moving background
and of reflections could not be completely
eliminated. In fact the correlation peak is still
locked at the propeller or nozzle velocity in the
regions where there is a lack of particle traces or
where the reflections are dominating the PIV
images. This effect appears in the PIV images as
vectors of very high velocity values, which are
called spurious vectors. In the post-processing
phase, the validation procedure is very effective
to detect such spurious vectors due to the large
difference between the flow and the blade

PERFORMANCE CHARACTERISTICS
A series of performance tests on the
same propeller model were performed by El
Lababidy (2003) in a cavitation tunnel. Fig. (3)
shows the performance curves of a DP thruster
with and without the nozzle at propeller speed of
20 rps. The results of these tests indicated that
the thrust generated from the propeller only
without the nozzle is higher than that generated
with the nozzle working at the same advance
coefficient as would be expected.
VELOCITY FIELD CHARACTERISTICS
AROUND THE PROPELLER
All the results presented in the
following are obtained by phase averaging the
measurements for a blade position having the
trailing edge at the 12 oclock position.
Fig. (4) shows the distribution of the
propeller
wake
velocity
components,
longitudinal component (U) and cross
components (V and W), at 0.4 D downstream
axial position. Results are plotted for the DP
thruster with and without the nozzle at the
bollard pull condition (J=0). The results indicate
some common features between the distributions
of the flow velocity components with and
without the nozzle. The distribution of the axial
velocity shows low values near the hub
increasing radially outwards to reach a maximum
value at a radial position around r/R = 0.6-0.8 for
the measurements with the nozzle and around r/R
= 0.6 for the measurements without the nozzle.
The distribution of the cross velocity
components indicates similar magnitudes of
velocity over the two halves of the propeller

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

velocity component for both measurements with


and without the nozzle shows:
- The wake of the ducted propeller persists
further downstream than that of the open
propeller due to the concentration effect of
the nozzle wake.
- The wake contraction between the first two
planes, which are very close to each other
and to the propeller plane, causes some axial
acceleration which is indicated by the higher
axial velocity achieved in the second
measurement plane.
- A reduction in the longitudinal flow velocity
starts beyond the second measurement plane
(X/D = 0.5).
- The beginning of the slipstream broadening
is about one diameter from the propeller
disk.
- The slipstream fluctuates unstably in the
downstream region beyond X/D = 1.5 for the
measurements with the nozzle and X/D = 1.0
for the measurements without the nozzle.

disk, but opposite signs, as would be expected


due to the swirl that is dominating the flow field.
Some deformation occurs due to the high effect
of the strut wake, which is disturbing the inflow
to the propeller.
The effect of the nozzle in reducing the
loads acting on the propeller blades due to the
acceleration of the inflow velocity caused by the
contraction, which reduces locally the blade
angle of attack is also demonstrated in Fig. (4).
This can be seen by comparing the results of the
measurements with and without the nozzle.
Measurements without the nozzle indicate higher
axial and cross flow velocity components than
measurements with the nozzle.
When removing the nozzle a contraction
of the slipstream is observed. This is due to the
fact that only the inner part of the propeller blade
is able to generate lift while the outer blade
regions are stalled. This would be expected
because, for the adopted P/D in the present
measurement campaign, the thruster is operating
far from the maximum efficiency as shown in
Fig. (3). This is demonstrated by the distribution
of the total flow turbulence of the DP thruster
with and without the nozzle at X/D = 0.4 as
shown in Fig. (5) which indicates that
measurements with the nozzle have a larger
thrusting region, and lower turbulence, than
without the nozzle. Also, Fig. (5) indicates that
the turbulence intensity is stronger near the
propeller blade tip and hub wake region than
other wake regions.
The propeller blade wakes could not be
recognized from the mean flow velocity
components shown in Fig. (4). Fig. (6) shows the
vorticity distribution of the DP thruster with and
without the nozzle at the downstream axial
position of X/D = 0.4. The maximum values of
vorticity are found at the hub and propeller blade
tip regions. The trailing vorticity, shed from the
blade trailing edges, is easily identified and
consists of two circular layers of opposite sign
which overlap at about r/R = 0.7, at the blade
section of maximum loading. This result was
also obtained by Felli et al. (2002). It is difficult
to identify the tip vortices in both the
measurements with and without the nozzle, due
to the high loading condition of the test.

The near wake evaluation of the mean


axial flow velocity shows that measurements
with the nozzle have lower mean axial velocity
values at similar radial and downstream axial
positions than those without the nozzle. This is
because the ducted propeller is more lightly
loaded than the open propeller. It is also noted
that the contraction of the propeller slipstream
behind the propeller is less with the propeller in
the nozzle than with the open propeller. This is
due to the effect of the nozzle in reducing the
displacement effect of the propeller wake (El
Lababidy et al. 2004). Moreover, measurements
without the nozzle indicate higher reduction and
fluctuation in the axial velocity among the
measured axial planes when moving downstream
from the propeller than measurements with the
nozzle. However, the far wake evaluation clearly
shows that measurements with the nozzle have
higher axial velocity than measurements without
the nozzle. This implies that the propeller
without the nozzle has more rapid diffusion and
dissipation of the wake energy than the ducted
propeller. Therefore, the rate of wake energy
decay is higher in case of the propeller without
the nozzle than with the nozzle. That is why
measurements without the nozzle have higher
axial velocity in the near wake region and lower
axial velocity in the far wake region than
measurements with the nozzle. Also, this implies
that the breakdown of the open propeller wake
takes place earlier than the ducted propeller
wake.

LONGITUDINAL WAKE EVOLUTION


Fig. (7) and Fig. (8) show the near wake
and far wake distribution respectively of the DP
thruster mean axial velocity components. The
evolution along the longitudinal axis of the axial

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

of the accelerated fluid over a larger disc area


downstream of the propeller. Hence, if the
thruster is used as a dynamic positioning device
with potentially other structures placed in the
slipstream, a ducted propeller may be preferred
to spread the intensity of load from the
slipstream over a larger area of application.
From the experimental point of view,
stereo-PIV shows a number of advantages
compared with the well-assessed LDV
technique. In particular, and considering the
limited time usually available for these tests
combined with the management and technical
difficulties typical of operating a large testing
facility, the PIV technique can provide results
within a short period. The LDV technique
requires up to three-four times more testing time
to obtain the same information, which
consequently translates into additional costs of
facility occupancy. The measurement time is
drastically reduced with the stereo-PIV method,
where the plane of measurement is mapped
instantaneously and provides all three-velocity
components in one single step, while the LDV
technique requires multiple measurements to
scan the interrogation domain. In this sense, the
PIV approach offers the freedom of extending
wake survey to a larger number of areas of
interest, with very limited setup changes. The
major drawbacks of the PIV technique are a
reduced accuracy with respect to the LDV
technique and the huge quantity of information
gathered. One must address the critical problem
of storing, managing and processing this
information without compromising the test costs
by extended data processing time.

Moreover, Fig. (8) shows that the flow


loses it identity when moving downstream from
the propeller. As the distance behind the
propeller increases, the gravitational effect on the
displaced fluid becomes more dominant and
results in a vertical collapse and a horizontal
growth in the wake region.
The ducted and the open propeller wake
do not have a circular geometry in the near wake
region close to the propeller disk. This is due to
the effect of the propeller strut wake. The strut of
the DP thruster model has a conical shape with a
diameter larger than the propeller blade width
and with a fin from the bottom side as shown in
Fig,(1). This large geometry of the strut affects
the slipstream wake as well as the downstream
wake of the propeller. The effect of the strut
geometry in the downstream wake appears in
both measurements with and without the nozzle
as a deformation of the wake geometry from the
top left side at X/D = 0.4. This deformation
affects the downstream broadening and
spreading of the propeller wake. Fig. (7) and Fig.
(8) show that the effect of strut wake is higher in
the wake of the open propeller than the ducted
propeller.
CONCLUSION
The analysis of a DP thruster wake
when operating with and without a nozzle at the
bollard pull condition (J=0), in a large cavitation
channel, has been performed using a stereo-PIV
system.
An analysis of measurements of fluid
flow velocity in the thruster slipstream with and
without a nozzle in the bollard pull condition
indicates that there is greater axial velocity from
the wake for the open propeller than the ducted
propeller. The slipstream contraction of the wake
along the downstream direction occurs up to X/D
= 0.5. Thereafter the slipstream fluctuates
unstably in the downstream region and begins to
broaden at about one diameter from the propeller
disc for the measurements without the nozzle and
1.5 diameters from the propeller disc for the
measurements with the nozzle.
Measurements without the nozzle show
higher flow turbulence intensity in the near wake
region than measurements with the nozzle.
However, the rate of wake energy dissipation
and decay is more rapid in the case of the open
propeller than the ducted propeller. This
indicates that one effect of the duct, for a similar
level of total thrust, is to spread the thrust
generated from the increase in axial momentum

ACKNOWLEDGMENTS
We thank the Institute for Ocean
Technology of National Research Council
Canada, Memorial University of Newfoundland,
the Italian Ship Model Basin (INSEAN), Rome
Italy, and Petroleum Research Atlantic Canada
for the financial support on this experimental
work, and the Natural Sciences and Research
Council, Canada, for the provision of a PGS B
fellowship to the first author.
This Project was undertaken and
completed with a Grant and the financial
assistance of Petroleum Research Atlantic
Canada. (PRAC). Thanks are due to Norsk
Hydro Canada for their participation in the travel
support for Said El Lababidy to visit INSEAN
through a grant to Dr. Thormod Johansen, and to
the INSEAN large cavitation tunnel technical

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

References

Felli, M. (2003), A versatile Fully Submersible


Stereo-PIV Probe for Tow Tank Applications.
Fluid Measurements and Instrumentation
Symposium, Honolulu, Hawaii, USA.

Calcagno, G.; Di Felice, F.; Felli, M.; Pererira, F.


(2002), Propeller Wake Analysis Behind a Ship
by Stereo PIV, 24th Symposium on Naval
Hydrodynamics, Japan.

Hoshino, T.; Oshima, A. (1987), Measurement


of Around Propeller by Using 3-Component
Laser Doppler Velocimeter, Mitsubishi
Technical Review.

Chesnack, C.; Jessup, S. (1998), Experimental


Characterisation of Propeller Tip Flow, 22nd
Symposium
on
Naval
Hydrodynamics,
Washington D.C.

Jessup, S. D. (1989), An Experimental


Investigation of Viscous aspects of Propeller
Blade Flow, Ph.D. Thesis, The Catholic
University of America, Washington D.C

Di Felice, F.; Felli, M.; Ingenito, G. (2000),


Propeller Wake Analysis in Non Uniform
Inflow by LDV, Proceedings of the Propeller
and Shafting Symposium, Virginia Beach.

Liu, P.; Bose, N. (1998), An unsteady panel


method for highly skewed propellers in nonuniform inflow, ITTC Propeller RANS/Panel
Method Workshop, Grenoble.

Doucet, M. (1996), Cavitation Erosion


Experiments in Blocked Flow with Two Ice
Class Propeller Models, Master Thesis, Faculty
of Engineering and Applied Science, Memorial
University of Newfoundland, Canada.

Liu, P.; Bose, N.; Colbourne, B. (2001),


Automated Marine Propeller Geometry
Generation of Arbitrary Configurations and A
Wake Model for Far Field Momentum
Prediction, International Ship Building Progress
(ISP), Volume 48.

stuff, for their support and help during the


experimental work.

El Lababidy, S. (2003), Performance Tests of a


Dynamic Positioning Thruster with and without
Nozzle, IMD Cavitation Tunnel Tests with MV
Robert LeMeur Propeller Model, OERC Report
No. 2003-02, Memorial University of
Newfoundland, Canada.

Liu, P. (2002) Design and Implementation for


3D unsteady CFD Data Visualization Using
Object-Oriented
MFC
with
OpenGL,
Computational Fluid Dynamics Journal of Japan
(CFDJJ), vol. 11.
Morgan, M. J. (1978), Dynamic positioning of
offshore vessels, PPC Books Division,
Petroleum Pub. Co.

El Lababidy, S.; Bose, N.; Liu, P. (2004),


Evaluation of a Dynamic Positioning Thruster
Wake using Laser Doppler Velocimetry, 23rd
International Conference on Offshore Mechanics
and Arctic Engineering, Vancouver, British
Columbia, Canada.

Min, K.S. (1978), Numerical and Experimental


Methods for Prediction of field point velocities
around propeller blades, Massachusetts Institute
of Technology (MIT), Department of Ocean
Engineering, Report No. 78-12.

Fay, H. (1990), Dynamic Positioning Systems:


Principles,
Design
and
Applications,
Imprimerie Nouvelle, France.

Nienhuis, U. (1992), Analysis of Thruster


Effectivity for Dynamic Positioning and Low
Speed maneuvering, PhD Thesis, Delft
University, Netherlands.

Felli, M.; Di Felice, F.; Romano, G.P. (2000),


Installed Propeller wake analysis by LDV:
Phase Sampling Technique, 9th International
Symposium on Flow Visualization, Edinburgh.

Prasad, A.K. (2000), Stereoscopic Particle


Image Velocimetry, Experiments in Fluids,
Vol.29.

Felli, M.; Pereria, F.; Calcagno, G.; Di Felice, F.


(2002), Application of Stereo-PIV: Propeller
Wake Analysis in a Large Circulation Water
Channel,
International
Symposium
of
Application of Laser Anemometry to fluid
Mechanic, Lisbon, Portugal.

Raffel, M.; Willert, C.; Kompenhans, J. (1998),


"Particle Image Velocimetry", Springer.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

KT, 10 KQ, Eff

Solof, S. M.; Adrian, R. J.; Liu, Z. C.(1997),


Distortion Compensation for Generalized
Stereoscopic Particle Image Velocimetry, J. of
Measurement Science Technology, Vol.8.

Stella, A.; Guj, G.; Di Felice, F. (2000),


Propeller flow field analysis by means of LDV
phase sampling techniques, Experiments in
Fluids, Vol.28.

1.0
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0.0

KT Ducted
10 KQ Ducted
Eff Ducted

0.2

0.4

0.6

0.8

KTOpen
10 KQ Open
Eff Open

1.0

1.2

Figure 3: Propeller performance curve (propeller only)-with and without a nozzle (P/D =1.2 & n =20 rps)

Figure 4: Circumferential variation of velocity components around the DP thruster at J=0 and at X/D =0.4
8

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Figure 5: Distribution of velocity total turbulence around the DP thruster at J=0 and at X/D =0.4

With Nozzle

Without Nozzle

Figure 6: Vorticity distribution with and without the nozzle at J=0 and at X/D =0.4

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Figure 7: Near wake mean axial flow velocity evolution for the DP thruster with and without the nozzle
at bollard pull condition up to X/D =1.5 downstream

10

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Figure 8: Far wake mean axial flow velocity evolution for the DP thruster with and without the nozzle
at bollard pull condition up to X/D =15 downstream

11

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Christopher Chesnakas
Naval Surface Warfare Center, Carderock Division,
USA
In the plots of the flow downstream of the
thruster, no blade wakes or tip vortices are visible. Is
this due to lack of spatial resolution of the
measurements? Was any flow visualization
performed to examine the stability or durability of the
blade wake?
AUTHORS REPLY
We thank Dr. Chesnakas. The DP thruster
blade wake and tip vortices could not be recognized
from the DP thruster wake velocity distribution this is
due to the high loading condition of the tests.
Measurements were performed at propeller pitch
ratio of 1.2 and at bollard pull operating conditions.
On the other hand, the blade wake could be
recognized from the vorticity distribution of the DP
thruster as shown in Figure 6.
A serious of wake survey tests on the same
DP thruster model were performed at INSEAN large
cavitation tunnel at higher advance coefficient values
(0.4, 0.45 and 1.029). The blade wake and tip
vortices of the DP thruster wake became more visible
as the loading condition of the DP thruster is reduced.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Stuart Jessup
Naval Surface Warfare Center, Carderock Division,
USA
I commend the authors for another
application of PIV, which has been extended, to two
camera stereo imaging. I would like to ask a few
questions which I may have missed in the paper.
The performance data shown in Figure 3
indicates the thrust produced for the ducted rotor was
less that that the rotor tested alone. Did the thrust
shown for the ducted configuration include the thrust
produced by the duct? The thrust of the accelerating
nozzle may a significant contributor to the total thrust
of the unit.
The PIV data was taken at J=0. The tests
were conducted in the water channel. Our experience
with water tunnel tests is that true bollard conditions
are hard to attain due to tunnel recirculation. In
Figure 4, from the contour legend shown, it appears
that there maybe some nonzero flow outside of the
propulsor disk. Can a more precise value of J be
determined from the test data.
For the PIV tests, I did not see a stated test
rpm. Were both the ducted and unducted cases run at
the same rpm?
I believe there is a significant effect of the
nozzle on accelerating the flow, which will be done
in an axisymetric fashion. It will create possibly a
20% increase in flow through the propeller, without
adding blading turbulence. This is consistant with
Figure 5, where the turbulence inside the rotor disk is
much less than the rotor alone case.
Can the authors explain how the PIV
system was calibrated. Was a calibration plate used?
Was it traversed or was a 3-d plate used.
The individual PIV vectors maps were
averaged from the 129 individual images. Has any
consideration been given to averaging the correlation
maps prior to generating the vector maps.

We agree with Dr. Jessup that true bollard


conditions are hard to attain at small cavitation
tunnels due to tunnel re-circulation. However, the
size and the characteristics of the INSEAN large
cavitation tunnel tunnel (10 m in length, 3.6 m in
width and 2.25 m in depth) enable to investigate a
true bollard pull measurements. The figure below
shows that the flow is almost zero (0.001and less) at
r/R beyond 1.25 for the measurements without the
nozzle and r/R beyond 1.45 for the measurements
with the nozzle at X/D = 0.3.
It is mentioned in the paper that the SPIV
tests for the ducted and open DP thruster were
performed at a constant propeller speed of 20 rps.
Also, the calibration of the SPIV system carried out
using 2D calibration plate.
Regarding to the velocity vectors maps,
there were no consideration taken into account in
averaging the instantaneous velocity vectors (129
images/plane) in order to generate the mean velocity
vector map from each plane.

AUTHORS REPLY
First of all, we would like to thank Dr.
Stuart Jessup for his interesting observations and
detailed review of the paper.
The accelerating nozzle has a significant
effect in the total thrust of ducted thruster. It
contributes a bout the 50% of the ducted thruster total
thrust at bollard pull condition. The thrust
distribution of the DP thruster with and without the
nozzle in Figure 3 represents only the thrust of the
propeller without adding the nozzle thrust.

Copyright National Academy of Sciences. All rights reserved.

0.3
0.25
0.2
0.15
0.1
0.05

r/R

0
1

1.1 1.2 1.3 1.4 1.5 1.6 1.7

0.05
0.04
0.03
0.02
0.01

r/R

0
1 1.1 1.2 1.3 1.4 1.5 1.6 1.7

Twenty-Fifth Symposium on Naval Hydrodynamics



  "! $#$%&$#$
' ()
*)+-,./10 )243)5768'#'#9 '#9:1;%<#$.%)2=?>@>A4>B2CD<E)F>4'G'* HHI

JLKNM
prqts

O
O

PRQTSVUXWYPZ
fuoaP9nLQd[Xo

P[]\VS_^
[
^.\VSvUXSwPr\Vf

\Y`aP

QdceP

MxQTn^
oaQy\wP

fegh\V`aP
q{z|K

R'mV
Y ?BX ."'u?

ijWYflkmPn
n.Po
q}QTn~UXnQ\Y^fl[XS

V9l'&

]@.$

*/3u % 33%ee6R '%*( %e .%*< '(35*/


%3)& 3)(** %&'#'#$3% 3%&68%  '(3V'#R(
*<*."+

:1.3)%A@ ' 3)%u>43)  3*3%X:Aw57 .3u'3)#


3).%3 3).*<46*/3B '%&. 3)3)#Y4'(*< 3%
&/ V=79%&3T %3)&3*3)#+|7/'3e G.%;'&3)#
3).%3 3).*<X5?3%3( %%3#h.$*68%&.* '#
;3/'#*/3 '%&. 3)3%R  3.+7/3y;$*< 3)#]5
'3)# 3%<#$3*< 3#u'&G./.*R*l*/35| '#
3 ;3)w/G/T' * # *&6?( $*<*& 
%3)&*<2.*/3;'.d 69;*&/mG;' '#$() 
' **3+3* (32 % '&. '#T % 33%
 357*3%]*3)*u5?3%3X ('#'(*3)#**/3
 *<&#' $#$3"7.y
>w? '#( '%&3#5*&/
*&/'3  ;3*3)*#*<+B7/3G%33 3).*w5*/*/3
;'/'3)## * G
$#+.##$*&9*&/'3;$*< 3)#
3$ 3% 3).*< w#*<'&3)#*l #'*&3y#$ (3)#
(. $* *&''##$
' ()l((*&e '#
; *&/L*&/3a3
3% 3)* 9 #*&/'3( $*<*& 
%3)&*<%&34#$('3#+7/3w=?A%&3*%&3) %
#'(3
*&/'3'* *&3e5 '/3. 3'3)%& 53)2;$*
' **<*&3#$&(%&3) ' (3)3$*+
7/'3:A
3).%3 3).*<5;3@ %&
#$3)#*Y*&/3w(.  *
6*3%*&/'3% %&33)**.el*&/3.y$  .&'
9"!4 $#$%
#  ()+

7/'3d3d 6=?A.*.V *9 ]*&] %3)#(*


*/3  ' ;*V&*/3 33)(*.357 .3'3)#/'
'(%&3&3)#+7/3 ((*& 6-*/3 5?3Y'3)#y 6?
&/ *& '%&.3)*V;37@#('**.'#B&3*</G./
#3  '#'*%&;'3)'(3 $#$3)GE+4>@@$
() (*.' 6u*/3L5 %'#*&/3%&* *&G
'%&. 3)3%/'3 '%&.3T*&/'3%( ' ;'* d*#$33%
G.

#%3)&* w'#a F +.% *&/3() (*.] 6
*/3 '%&.  *3)**? 6" '%*( %- .%*< '(3
*/'**&/34
 3)%&($ 3*&/$# - ;'37* B*37*&/3
*&3)%.(*&;3*5?33) % 33)% #e/' "2e%<#$3)%
*;$*< G.

#%3)&*<+

. $@
7/3v '%&. 3)3%;3/''#
G.3v(%35
/'
 3%<*3)Y/G/ /. G3)3.''5T(.'#$*&')2
(</'%.(*&3)%&&3)#X;
X*/35?3 3 '#*&/3;'G3
.%*3 6]*&/3/' /
+
7/3'5*&57 %<#
*&/'3x %&. 3)3%;#3) #v*/3% 3)%68.%& '(3
;3)(. 3)768''(*. 6*&/3V 'G %? *& 6"*/3
% 33)%)2
#$3
*&
#3%3*

 2
*G3)*& B'#%.#$Y3)(*&3))+h7/3 % 33%
()(3)3)% *&3)*/3'5t#3 3'#$G.|*/3%.#$ 
.# 6*/3;#$3+
?3)/#*&/3 % 33%
(. 3'5
'3)#() r;3N3'(*&3)%&3#2
5/3)%&3*&/333)(*< 6e; *&//' |'# % 33%
& 3)%& 3.+7/3
53#$G3 6?*&/'3 '5;3/'#

$/ ''#$3%a % '&|('#$*&%&3 3*


*Y  '( *&'@( (' *&3)#5*/;$#$ 968%<(34
.%#3%*&%&3#$'(3y*&/'3T( $*<*.' w3%&*)+L7/'3
33(* 6*/3 8*7 '%&33).* % 33%?*/3w'5
  %$ *3)#B; #*&%;'$*&Gw(*.3?68%<(3))25/(</

3#*/3V*&/'%&'*4 '#*&/'3V*&%<32
R #$&$()*&3#
 **/3l % 33% '3+j@' 6835 ';( *&'
#3) 5*&/B() (*.'1('&#3%G4*&/3%&3  *%'G
'%&. 3)3%u*&/'35?3a 6T*/3&/ I+
7/
  %..(</&  *&3?*&/3w*&%3V*%'3*.*3%<(*&
;3*5?33)/  '# % 33)%;$* &w%3)%3) .*
( $*<*.' T%3)&%<(3+
!53)3%(%3)&G
3.%*9;3G #$3y*&3)*;&/*&/ ' %&(</
9 %<(*(3+
.%#3%d*&&  %&**&/l#33).  3*)268%*/3%
#*.e%3)%3)#+d%*&/'%3)&l$>/'
('#'(*3)#:Aa 3)&%3 3*<H'+E A;3)/#'#
H'+ HA68%&..*w 6-*/3B '%&. 3)3%w  3 6*&/'3Y=7
5*&/R '#5*/$*?%&'GB % 33%'#.;$*3#
*/33$(* 3#d]68(*& 6*/3G'%
.&*&6*/3l '%&. 3)3%y;.#$3))+7/3)' ;3)
*9# *&3=?A%3)&*V*w. y*&/3 ;'.w 6
G..;' '.**3 '(</]*/%&* #*&.%3R;$*
&*&(. ' %3V$('5'*&*&3))+

G3E
Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

:<; Y"'&m
4

7@ @-Y1&"m71

m"
4'1m $

>5
$#$3)j&/ $#$3EE.EE H.HE5*&/
( 3]% *&  F'E+ I/'.T;3)3  
$6.(*%&3#2
()(.%#$G*l*/3T#*<l6*&/3TY=7 %
#$3#;

Y'+?7/3 %'( ' @ ' %&*&('%6V*&/3y $#$3)
 #*/3V;$#$ 'R%&3wG3)9 ;1+EY'#G'+E
%3)& 3(*&3) .+7/3R#$& (3 3*BBG..35*&/.$*
'('#$G *&/'3Y%&'#'#$3%+
=7
>$#$3"'+-E.EE.E$H.HE
Y
:"3)G */ 3% 3'#$('%
 


?%&3#
*/


A@%G/*


A@& (3) 3*



 3**3)#9&%6+-%&3&/



 3**3)#9&%6+-%&3B%'##3%



?$(<(
3(3*

D

$+ C
E.+ H'E.
 H

H'+ FI'EC
E.+ 
I 
+  F.
H+E
H'+  H

;3E-
/ R $#$3)"G3) 3*%&


G./* $#$( *&' 6*&/3 &/ 353)%&3
3(3) % *&* %&. 3)3%/' 6*u5*/%&G.3%
#$ 3*3%w**&/3 *&3%;;"+V!45?3.3%*/3B33(*
 6
*/3)&3 
#()*.'V*/3- 3).'%&3#5y%3G..
 3#*&;3Y 61 %7  %&*'(3.+-7/3%368.%&3
*&/'3w.;$*'3)#3$ 3)%& 3*#'*3$ 3(*&3#*;3
 6? ' %&*&('%w*&3)%&3*w68%Y #*.y '%& .&3))+B
#'#$*.#$3%3*@%'##$3)%*/' *&/33'3#;

Y'/'V;33)y'&3)#d#$%G*/3%3)&*'(3  '#
% '&.]*&3*<+A@%&Gl*&/3d3)$(* ] 3)&%3D
3**&/3Y%'##$3)%5?.?%&3) .3)#+
 H.
> %&. 3)3%4+ E)F .
% 33)%7# 3*&3)%

 H+ H
*(</R% *&
!"
D
H+ 
*(</R 3 
$#&%('*)
D
H+  IH
>4%3)% *&
+-, "+/.
D
H+ CH.H
!4;#$ 3*3%%<*
0212"
D
H+E)C.H
4 ;3)% 6;.#$3)
3
D

 ;'34 %&. 3)3%7G.3 3*&%

 ;"+7T*/3 %'( ' @ ' %&*&('%6Y*&/3y%&G/*


E)F %&3
/''#$3)#h % 33)%a5*/
>4+
G3)"215/'3eG'+R*&/3&/' 36*/3;#3)Y
&/5"+7/3] % 33)%y/d;3)3m  
$6(*&%3)#
()(.%#$G**/39 % 33)%  H.968%&. Y'+
7/3 '%&. 3)3%4 '3 *&'3)6
# 587 
  H'+ H'9E 
 '*&%3)  6>+ ++

=<-4

>B.1

'%*&/3w:1A 3)&%3 3*<Y*%.3%<GV*75?.


.*&3)#y*&/3&/ "21()(.%#$G*&9G'+F9 #I'+
7/'3?3$ 3)%& 3*
&3*  5?3)#*&w .3*/3:1A
'%&.;3H+  |T D '@
# ?D#$%3)(*.y'#d 6?H'+ F 

D#$%&3(*&12w5*&/%'G]*&/3G.*'#$' 
#%3)(*&6*/3]/ "+ >j#$'B;3) *5? D(..%
;(<
5?%#'m( **3%G:A
&
*&3)
68%
4

57t'&3)#2 5/(</
54N*&
3&%3*5?
.%*/G.'V3$(* (. .3*& B*3.' +
7/'368$(  64*/3R&; 3)%&G.3)# %;3R5?.6  H'+  R2
5*&/*&/3;3) v& '.(Gu;3GH'+ H 9+ '%y*&/3
($G%<*. 3&%G| 3) A
6 H B

# 3*&3)%V'#6EC+  y3G*&/y53)%&3;$*< 3)#+
:&3%;3 uG%33) '#;'3y53)%&33)  .3)#2
5*&/*&/3G%33VG./*/'
G4G%3) *&3)%" 3)%68.%& '(3.+
(</Y&3%;3 X(..%"&
( *&3#*.3-3)$(*
( 3).*+ .%#$3)%*&Y 3).'%&3?*&/'37*&/%33#$ 3$D
&'3)$(* '3#Y3).(</V .*/#w*&;3 3&%3)#
*5(3.2'(3 *5?3)$(* y( 3).*<B(#
;3l%3)(%<#$3)#'3e%&'"+ 7/3)%&368%3T*&/3 .*
5?3%3e 3).'%&3#68%&. ;35 #a68%&. }*/3]#$3+
'%9 .*< #/' ';.3*/3 % 33)%R&/'6*
3&%&3) 3* *  G.396wI
 DR68% */3&#3
/3V;3)39 3%&68% 3#+
7/'3 3).%3 3).*<5?3%3(%&%3)#d$*e*&/39
>
*5'G*T5*/.##$*&'?&33#$'G'+  */e*&/'
  %..(</*57 &&;'3*'(%3).37*/3@# *V%<*3
#%&G */3Y 3)&%3 3*<(#$3%< ;' .2(</3
G
d& 3)3)#e' l 6*&/3R'*7#'*% *&3R 64 l*&T $D
'%&$  *&3) HB* 3+
A@%'GN*/3_ G.% ;'.3# 3).'%&3) 3)*j
* *< 46E$2 HHH 3)5?3%3(.3)(*&3)#%'#$
$#$3*3(</%"2#$&%&3)G. %<#$G5/'(</m 3)((
.3$(* l(.  '3*5?. %&3(.%#$3#+7/3
 B;3%
6 3) 3)&%3)#5*&/*/3G%33;3) 5?.
  %$ *3 *5(3.%&G.3-5*&/V*/3;'3.3)%
;3) 9+7/3(3(*3)#Y# *4 *"3)(</Y 3&%&G7 .*
57V&%&*&3##$3 3#$'GR.*/3  G.%@ .&*&d 6
*/3 % 33%Y;#$3Yd.%#3%V*&.;$*y*&/'33)D
(*3768%7*/3 3V
()(."+
'%d*&/'3' %&. 3)3)#/' *& 3;'.3# 3).'%&3D
3).*<5?3%37() %%&3)#B$*2%3)(%<#$'GE '2 H.HH@'3)68%
3(</R 3)&%'G *)+
7/'3#'*57(.3)(*&3# (
#$%&(( D
.%#' *&3d $*&3 92@5*/%Gu*&/3T$6Y*&/3
'%&. 3)3%&/'6*+T%<#$ #$%3)(*.*/3#*<R5?.
3&%&3#68.%F%  H'+ IH2H'+ .$2H'G+ H2H'+ C$24E.+ H.H
'#uE.+E$+@e3).(</e%.#$'*/3 3&%&Gy .*
x(%( 683%3*& u#%3)(*&j53)%&368%G'%

G3Y
Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

.&*.'7 6IH  H'2F+C 2J $+ 2E$+ H2.$+ $2$FH'2$I$24 .H2



2 H2E)H$2EH2E)F.2E H'2EK .2'#E)C.H#$3)G%332
5*&/T*&/'3 3&%G9 *;3*5?33)]E)CH9'#TF H
#$3)G%33) (</3).((.%#G. .+*&*YE)CHy .*<
5?3%3 3&%3)#w68%"*&/3?/ Y5*&/Y*%&'G %&. 3)3%+
7/3d
H DD .&*.R#$3'3)# #/' ' ;3*/3
% 33)%&/'6*+%*&/33$(* l 3).'%&3) 3)*
5*&/*/3' % 33#/ */3L 3).'%&3) 3)*
5?3%3%3)*&%(*&3)#*&m3#$3295*&/*&/3a(%<( D
683%3*%< G.3X;3)G3
*&3)'#$3#;
hE#$3G.%&3)3)
3)%B*/39
3*&%  3.2G.
GE H 3&%G
.*<+
7/3m 3).%3 3).*<5?3%3() %%&3)#.$*
H+ A m68%&.*l '#H+EAj;3)/#m*&/3 % 33%
 3 68.%Y*&/'3/' T5*&/T*&'%&G % 33%B'#d
*&/'3 %&. 3)3% '3?68%*&/'34/ 5*/*- %&. 3)3%+

L @
y


."$

NM-'OM- "$P

u( (' *&3T*&/'3l


(.'R'5 %#*/3l&/ "2
*&/'3 % 33)%l #m*&/3/ 5*&/ % 33%e*/3
4>@@3)' *&']5?3%3&3)#
 3%()  .2'D
G*&/'3(. 3)%(1& 6*57 %3Y '(< G.3)=?R
Qw '#
=?R
Q4D>w'='5Y+j%T#$3* .X*&/3
 3%() 
3*/$#B 3).37&33V ''# +7/'3?TD Sd$e $#$3)
 63).*3%G )1577'&3)#*& $#$3)*&%;3'(3+
.%9*&/3T
 3%( () (*.'R;
(<D*%&(*&'%&3#

 3%() G.%&#T/'.3;33)G3)3%<*3)#5*&/*/3
& 6*57 %3@&=?9D=?A+
>L*&3#$ T ' %&(</TB'&3)#*&() (*3*&/3 % D
33)%w. 3)y57*3%@*&3*<+B?3( 3 6-*/3 / G.3D
3)'1$5d'#w*&/3%&3'*G4&
3*&%('5Y2*/3

 3%() 7 3)&/() l;39 *3)#*&d'3 % 33%
;#$3TT  D &3G 3).*  6wy( '#$3%5*/l 3%$#$(
&#$39;# % e(.'#$*&' l*/(&3+]7/3R68
% 33)%G.3 3*&% ('&#$3%3)#]*&/'3y %3)&3*
 ' %&(</"+w7/3 .$*.y#$.  /'@;33d(</.&3
*&y3
*&3)'#EH % 33)%#$ 3*3%<68%*B 6*/3
% 33)%)2EC#$ 3*&3%<R%.#$ #$%&3(*&R '#FH
;3/'#*&/'3 %&. 3)3%+u7/3
 3%(  3/]57
#$
#$3#*&3%E .2EC IG.%&#(3)<4 # 
*&3% ' %&*@.'2  H@G%#(3<+7/34G.%&#*&. ..G
57(</3)e(</"2*&/' *B*&/'33% ' %&*)2&%%&.'#
D
GY*&/'3 % 33)%-;#$3.2(#  B;343#68%-*/3
'*&3#$ d B' *&d 6?*&/3 '5m %#*/3/'
5*&/9 % 33)%)+G+'*/3&%6.(3w 3)&/R.*/3
% 33)%;#$3V4/'5"+
7/3YY=757?
.3)*&G.*3)##3%3*
 3%() 
3)&/3))25*&/#3%3*G.%&#d*&. ..G3) '# 3/
&U ?)3))+>44 3)&/35?3%3RG3'3%<*&3#l&'(</*&/' **/3
3% '%&. 3)3% 3)&/(#;3'('#$3#+]l*/3
685G*&/'3)&3 3)&/3@5;3#$3) *3)W
# V/XYZ [B*&

V-XY\Z9]'+t7/3( %<3*TG%#^V-XY\Z2[(.'*e 69I


*&;$(<#$.  '2
5/(</%&3w(.3)(*&3#*&3).(</
*&/3)%4'B  *(</Gw*&3)%6.(3+7/3
 ;3)% 6
G.%&#(3@@ B(</y& 3%w(.   %3)#R*&*&/3 */3%
3/32w68.%R.3T%&3&u;3G2@*/'*9*/3;.'#
D
%& X 3)%T *T%3)&3)##$5**/35?
 **3 $*&3#5*&/*&/3 */3%d 3)&/3))+ %_V/XY\Z"`
G.3;$(<#$. & 3( 7**3*&l5?.B #
' ..33 'G*&/'3T(3Y& 3(*% *&l;35hHH'2
5/'3a
 V/XY\Z"b&33@G'
+ .-'<
# V/XY\Z9]Y*&/3@;.'#
D
%&  3)%]%3'3#+
c
 V/XY\Z2['d
# V-XY\Z"bm
&
3*&%  '357$(*3)# *u*&/3m*d5? D
*3%3.2 B' *&Gyy#$;'3;$#$ $#$325/3
e
 V/XY\Z"`X 'f
# V-XY\Z9]X*/3u %3
'& m(.  '$*&3#
68%339&%6.(33)3)*&/';33(.'#$3)%&3#+d7/'3
&%&6(3?3)3)*&57((*&3#B5*&/
>0  '3
($#$3RY:@+*57B **&*&3)  *&3)#e*&y
3*D
G*34*&/'3@#$&(%&3*&g ?) *&3)%&%%-;
3%&68% GB() ($D
*.'7#$3%3*4 3)&/3);'$**&
3)*&G. *&3@*&/3
$''3'(3Y 6 '&;3VG%#*&. ..G3))+
4+-(3

V/XY\Z2[
V/XY\Z"b
V/XY\Z"`
V/XY\Z9]

E)2
I $E)H.
C.H I.I.FH
'E)CI.CF
C.F I.HH

h4i
+ H
H'+ 
F'+ IH
E.+ H.H

 +(3
& 3(*%<*&
E  .
E CC
$EHH
EKE FC

 ;3YFJ4' 3)%&(" 3)&/"2Y=7


m;"+YF*/3(</ %<(*3%*&( 6*/3'#$
#' 
3/3768%3&#3Y 6*&/3Y=7R %3VG31+R*&/3
*/%#Y( Y*&/3 %3)3)%G3##$ 3'&.3)5?
#*'(j
3 h4i -kl J
m68.%3).(</G.%&#G3)"2
5*n
/ m
;3G*&/3?#$
' (-
(* 2 k8lpo q\r 79s@*&/'3&/3)%
.3$(*  '# qFr */3/3 %*&%3)V*V*&/357 +
*%&  '# &G37/'37;33)(#$3%3)##$%GV*&/3

 3)%&( B*."+>' 3/'3)3
*3'#'3?&/
3'G *&/*&4*/368%&..* 6$*/3&/ "2$E+   
*&4*/37#$3)
'#9#$5
5?%# '#$+   
e;3/''#*/3/' "+
'%Y*/3() (*.'V 67*/3'5 %#*/3&/
5*&/% *<*&Gm % 33)%]''*3).#$ L ' %&(</
57(</'.&3"+7/3 % 33)%G3) 3*%& -('&#3%3)#
68#3*+L7/3T&/ 9() (*3)#u*D
*.' % 2?5/39*/3 % 33)%d% * *&Gd68% 3
6%3683)%&3)'(3.+.%T*&/3&/ /
 *&/3]
 3%( 
3/]#$3' *&3t
# V/XYZK]d/' ;33)3#2?&33G'
+
+
7/'3u B*./'T;33)m 3%&68% 3)#&'(</*/'*
%**3)# L.*&68%]*&/'3X %&. 3)3%u *u
G..3G'% .&*.57.;$*'3)#u;368.%&3*&/3
''*3).#$  B' *&57V*%*3)#+7/Y ' %&(</
3)' ;'3 6.*3%() (*.R6-'*1 3)@'(3
*/3* 34#$3) 3)'#$3(  6*&/3&$*.-&  %3)3#

G3VF
Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

*&*&/'3 '%&. 3)3%VY 5?3)#9*9%&**3+


* %&*&G
*&/'3x'*&3).#$ t( $*<*.N68%
*&/*
 3)(.'#$3)%; &.3)u( $*<*.' *& 3.+
7/3X*3)# &  *&*&/3) (..*
3]*&
3%$#$(* a*& 3e%3)(</'3)#2w68%y5/(</a*&/%33
% 33)%%3.$*&'453)%&3'3)(3& % R*/@()&3+
7/3R % 33)%;#3/'.3;33% * *&3#;
d2
I D
3.3% *& 3*3 "+

u 1
$ev=w=<"a'  MmY1"
%  33%? 357*3%*&3*<5?3%3@('#'(*3)# **/3

>5*&/*/34 % 33)% EF.   H+1G+.C
*&/'3d. 3)57*3%R*&3*%3)&*9 %3T( '%&3#u5*&/
*&/'39#'*68% Y''#+= '%&G*/3
3).%3 3).*<68.%x#$ (3{(
3(3*x,  H'G+
;3*5?33
> '#lY'#$3
*.'6E+ x 
ywz  '#6y$+ 2I xv y|{ ( L;368'#"+7/3
(.%&%3)& '#$GX#$3)%&3)'(368.%e*&/'3)&3)e3l%&3
$+
C x '}
# + 
 xh()&3R*&/3# *y68%&. 
> '#
$7 %3V( '%&3#+

>#'(3(
36+
7/%'*(
36+
%<.'3w(36+

~
y z
w
y|{

E
H + H.H

H+ $E.E
H+ F. I


H +C I2

H+E)F I
H'+  ..

F
H + HH

H+ H.I
H+E I

;3VI-=?A%&3*)2$ % 33)% 3R5? *&3)%7*&3)*

(</ %<(*3%*&('#  G.%&3)3X5*/h ' %&$ *&3)


I+  x+'%*&/3V&/ R 33#|  I.**&*  $#$3
%3)&*< '(3
6   F+  JE)H48X '#%3)&#$' 
%3)&*< '(3 c
6  HC+   4E4
H 8 57 3)&%3)#+
3&39 *32?*&/' **&/333(] *;3(. D
 %3)#5*&/*&/'3]#'*G3)*T*/
3 w*&/3)
;%G
HH.H  %
&/. "2&'(3*/3 53)%&3y;$*< 3)#5*/
%'##3%'#T*B3)3)33)('#$*&1+943)&
68.%9*&/3#$3%3'(3da%&3* (3T(
3(3*< 
;3y#3*l#$&(%&3) ' '(3u/' 3 &e%&##$3%2
57*3%*&3 3%<*%&3 '#(%%3)(*.'&'(</.68%
*/3V*;'$(< G.3+
7/'3w $#$3)"5?.68%&3)3**%& L'#&#$%G*&3)*D
G+>*3%y5&3*%& |#$3'3#*;3 .&*&3+
G'+-E.EB*/3#$
' (*&%  G.3/'5d'#
( '%&3#]5*/*&/3d#*<eG..3;
Y4'+?7/'3
*%& G3)R.3%*&/3e& 3)3)#%G3T&/5|G
$#
G.33)% G.%&3)3 3*+
%L*&/3/' }& 33)# 6
  I..*<]#$
' (*&% 
6 H  DJ
H E D5?.
3&%&3#yY'
pH  D
H "E D +
  "
; + 4*&/3?
.3)*&G.*3)#B 3%<*&B .*G3)
T #  +EK  w (%%&3 '#'*|  I*<+
=7 $#$31(3
V
3)$(* $#$3)
J#
w 
4 *&3Y6%3.*&

"E 
'%&.'#$3V
 B;3%

D
3)
#7
 B;3%
/
D

$+E
+ 
H+  H
E
I E)H

 ;3B4@ 3%<*&R .*


7/3a
 3%() 9 B' *&L6*/368%33a%G
% 33)%#$3%3)*& *&3#*&/3 3)&%3)#*&/'%&'*
(
3(3)*X;
|  %
 *3 E)2
H x25/3*/3
*&.%3 (
3(3)*Y %3 %&3#$(*&3#d3% ()(%<*3 2
*&/'3Y3%%&.%?;3G 3)?*&/'d"E x3)3VG'
+ +

u 1
$ev= Y11tLem71
3*< '(3
*&3)*
5*/
*&/3
Y=7
$#$3)
eEEE.E$H.HE3)  3)#m5*&/%&##$3%/3];3)3
3%&68% 3#l*&/3*&5Gd* 'e 6
>+>57*3%
*&3) 3)% *&%3  6@"E 2KE D=a57V 3).%3)#y#$%'G*/3
%3)&*'(3d '# %&. u*&3)*)+L7/3y*&3*# *
R(%%&3(*3)#.%#3%*&](#$3%*u;$(< G3
33)(*)2.((%<#$'G**&/'3 3*&/'
#T6$(</
'*3% C+
yG'+E)H*&/3%3)&*'(3(%3V(.   %3)#5*&/
*&/'3Y'3)7G3);
Y4'2$;$*< 3)#68%457*3%
*&3) 3)% *&%36E)
H  D)=@+Y7/3 3)&%3 3*<y%&3
3' *&3#()(%<#$G**/3 3*&/$#6%#$35*&/
*&/'3=D
 %3)&*'(3T('%&.3d'#a
 3.+
  .3*/3a 3)#$ 3'&3)*&/3a5?3**3)#
&%6.(3d %3)e'('#G*&/'3y%'##$3)%   +  'E

57'3#+T7/3%3)&*< '(3(%3B&/5&  %

7/'3-5?3V*&/3- '%&. 3)3%  '3 6


*&/3 % 33#
Y=7 57 3&%3)#*&/'3*5GB*  6
>;

3 '6V:1Au'#w /37 %&.;3.+1G+$E4*&/3
 $57 .3B  *&*&3#l.3%*/39(%( 683%3'(368%
#3%3*?# 33)%.#$2 3&%3)# 5*&/; */
#3
(3))+-7/3# *&/5476 %& G
$#R G%33) 3*
;3*5?33):1A'#R /3V %;3w 3&%&3) 3*)+
G'+E)F]*/3 $5?3.#$3%3*d%.#$
( '%&3#u68.%#$3%3*
 3%( Y 3/3+ 7/'3
3/3Y#$3' *&3
# V-XY\Z"`R '
# V/XYZK]R%&3) %
#'(3*&/3
57 .33#l*&/39*&3%%<#Y8\%   HC+  #E.+ H
6*&/3R % 33)%B  '3 %3.(('% *&3 y*&/'*&/3
*&/3)%-*5? 3)&/3))+a
 V/XYZ9`B '<
# V/XY\Z9]Y*&/'34  3%
&$*.#$ ;# % ]/'#]*/33)3)*& 6
*/3468%33@%&6(34 '# *%&3*3)#0 68%&3)3 5?0+
7/'V  9#$()*32*/'*w*&/3 3)3) *&y 6-*/368%&3)3
&%&6(3@/'.7  '(**/3V5?3@&'3)7*?*&/3
'  3)% G. % .&*&')+*VV& 3)#2*/'*w*&/3
3)%&%%#$'39*&eT(3)%*< 
*& 6w*&/'3#$
' (
;''# % (#$*.**&/3@ %3)(%;3)#68%33&%&6(3
9 63&R  %&*'(3y68%R*/3d() (*.6*&/3
57 .3u*&/3 '%&. 3)3% '324( ' %3)#5*&/

G3wI
Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

 &&R 6*&/3V68%33&%&6(3V3)3) *&"+G'+"E)I'2


E '#E 9*/3R $2*G3)*& '#e%<#57 3
;* 3#.*&/34 3/#$3) *3)<
# V/XY\Z"`@-( ' %3)#
5*&/*&/33
3% 3)* Y# * 3).%3)#5*/*/3
/'3D %&.;3.+7/37G%3)&/56 %& G
$#G3)3%< 
 G.%&3)3 3*4;3*5?33)3$ 3% 3*< '#=?A#*<+
7/3G.%&3*&3*-#$(%3 ''(3)()  ;3.;'&3%3)#**/3
& 3%9%<#$2@5/3T*&/3' **<*&3G%33 3).*
*7*/3Y$*3%%<#$4& *&6.(*% +

u 1
$ev='-Y11tLeLm71
7/3#$
' (R*%& 6@*/3 % 33#Y=7l $#$3)
57 3&%&3#h68%*/3#$3)&G. .*2u
3)#$G
pH  DH E) I D+  '%&. 3)3)#('#$*&&G/*
#$3(%3)&3eX*&% {'#&'G3l( '%&3#*&u*/3
 '%&. 3)3)#&/ $#$3 (;3;'&3%3#+ '%
*&/'3:A 3).'%&3) 3)*@/53)3)%*&/3V=7 $#$3)
57$3)#B*&V*/34(%&% G3 *%& 3# (#$*.'
* */3#$3)&G]#$% 6*)+7/3((*&5?3%3(%D
%3#$*?5*/*&/'3@
 3%(  3)&/#$3' *&36
# V/XY\Z2[+
7/3%<*3 6%&3)$*.6*/34 % 33)%5?.(</.&3
()(.%#$G*& ;"++
G'+wE9 $2E)C'2YEl'#.T*&/'3y$ 2*G3)*& 
 #%<#3$(* (. .3*l %3 '**&3#m
*&/'3w '%&. 3)3%& *&%3) H+EA;3)/# % 33%
 3T68.%e*&/'3 3).%GX .*< *e*&/3%.#$ 
.&*.F%  HC+ w # *&/3w G. % .&*&|
 H  
H D 2
H   H D/
2 H  E)C.
H D'
# H   
H D9%3)& 3(*.3 +X7/3
# *Y57 .#$3@# 33)-5*&/*&/3w/ & 3)3)#
  $+E  |
<'#*&%< '68% 3)#*&u(
#$%&D
(R(
%<#$*&3 $*&3 9+%l*&/368'% 3).'%D
G .**/3a'*&3#$ ;3/%]6*&/'35
#$3) 3)'#$G*/34;#$3 *&"2.3#$GV*/'%&
3) $w*&/'3
35?3)#'*&*&3))2"( ;3 ;3)%&.3)#+
7/3
H D9 *&]#$3'3)# #&/ ';.3*/3
% 33)%/' 6*)2y5/3*&/
3 .
H D .&*.L .*<
*&e*&/3d*< %; %<#&#$3+7/3y G%33) 3*;3*53)3
((*& '#Y 3&%&3) 3*.'*3G.$#V68.%1*/3
3).%G .*<
* H  E)CH D  
# H   H D +'%
*&/'3l .*d
* H  
H De%&G.3%#(%3 ''(3)( X;3
#$3*&3)(*&3#l68.% *&/39%<#$ 7.3
(* .25/(</]#$()*3)
*&/*?*&/3w;G.3.%*3 5?.' *?&$(3* %3)&3#
;
*&/'3dG.%&#+>468%9*/3T .*n
* H  
H D*/3
*G3)*& B.3$(* a( 3).*dy*y %3)#$(*3)#
5?3+A3 *&3?*/3?'*&**.3?#3%3'(3);3*53)3
3).%3 3).*<7 '#9((*&2$3)%& 3)'(.%G D
GyG%33 3).**/3RG3'3%< (</ %<(*3%*&(B( 
;3 68'#"+7/3  '3)#y
 3%()  %$(3)#%&357
 ;'3B*&#3*&3)%& 3*/3#3 3'#$3)'( 6*&/3B'5X
*&/'3Y %&. 3)3%;'.#$3V *&R3)%& G.

#"+

]G'+-H2*&/'3R
35;3Gy68%&. ;3)/'#+ 3&3
' *&3.2l*/'*a*/33)$(* 3(*%<*&/3m'G.%3
%&3.*3% . *&3#j& '(32a/'5G#$3'&3%
.3)(*&%V3#*/' e.(*'  3&%3)#+7/3 */%'*
G.33)% *&3((3*&%(w&'(3w*&/3Y % 33%% * *&3
.*/3T .%*&#3TG.'* '#*&/3* %;. %<#
&#$3l5*&/X*/3l%G'5() '&'G]*&/3 % 33)%
* 3%<*35*&/ %
'Gy'('*.G3))+%
  %$ *3 N*/
3 
H D *&*&/3%&G.3)*
.3$(*&3)w( ;3 3)3"+V>4*&/'G./*&/3*& .%&*&(3
6 *&/33e;#3)# % 33)% %3 *d((%<*&3)
%3)&3#& '(32*&/3V .&*.6"*&/3V.%&*&(3)?(
(3 % l;3#$3).*3)#+a7/3 %G3)*$ 3)$(*
G.%.#$3*& '(3( ;3.;'&3%3)#e*&/357 .3
3)y #&/ 'B;.3*/3 % 33%&/'6*+R? */
;'G3%&*&(32 *&/'G./e * B %&.''(3#T5*/
. 3)% *&GR %&. 3)3%Y.V5*/$*21$'3)'(3*/3 3D
$(* '3)#dy*/Y%&3)Gd(#$3%< ;' .+7/Y(
;3.;'3)%&.3)#T;'&V 6?*&/3#557 %<#w .*'G
.3$(* 3(*&.%T #&/ 'e ;3*&/'3u % 33)%
5/'(</L%&3' 3)%& .&*. 6*/3;G3u.%&*&3
'#*&/'39* 'G3*43)$(* l( .3* 3#
;
*&/3e% *<*G] % 33%+|*&/3l5?3T 6*&/3
'%&. 3)3%;'.#$3)R*/3e3)$(* 3#a#*&%;3)#2
5*&/#3)(%3).3L*&/3a$ 9 '# '(%&3&3
*&/3*&%< .3%<&(. .3* 6y*&/33)$(*
.3)(*&%2.#3*&B %&3&&%33)'g ?) *& 33)(*-. *&/3
*%G3)#G36?*/3 '%&. 3)3%;.#$3))+eG+F
*/3y() (*3)#].3
(* '3)#]68.%*/3y 3;#3
.&*&dY&/5"+>w68%V*/3  */3%( ' %.'
;3*5?33) 3&%&3) 3* '#((*&2*&/3
G.33)%4'5 #3*R %3%&3) %
#'(3#u*3G.$#2
#3)& *3' *&* *&39#$3)%&3)'(325/(</%&3R .*
'%&.(3)#aX*/3l57 .3e 3 a;3/'#a*/3&/
'#d68%B*&/
3 
H D *&"+9eG'+E '#e I9*&/3
( '%&.T(%&%3#d$*68%*/3;'.#$3 .&*.
 I.H D +=. ' %'G9*&/3R3)$(* d'3#T68.%;#3
.&*&
  HDG+ H #L FT5*/L;#3
.&*&p
  I.
H D@G+$E7 '# I
*/37*& 3?/'*&.%&
6*/3V'5'3#9( R;3Y *(3)#+

7/3 3&%3)#R3)$(* '3#;3/'#9*/3B % 33%


 368.%y*/3;#$3] .&*.  . HDl/5

G3Y
Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

e/&/6

Ly @ M'1m
: A4 3).'%&3) 3)*Y68.%yG.3R&(%&3)5(*< 3%
1
3&&3Y=753)%&3-() %%&3)#@*1Y '%&. 3)3)#V #@'$D
% 33#9(#$*.' **&/3 
>+'%*/3
3)*&D
G. *&5*&/*%'GB % 33)%G %;'&3)# 3D
&%&3) 3*@53)%&3(%&%3#*)+7/3*& 3/'*&.%& 9 6
*&/'3-'5#$3) 3)'#$G.Y*&/3 % 33)%1;#$3 *&
57V%3)(.%#3)#+d.##$*&T%3)&*'(3.2 %&. 
 #u % 33)%9 357*3%*&3*<953)%&3d(#$'(*3)#
*&( '%&3w*&/'3;$*< 3)#9# * 5*&/  ;3w*&3)*
# *+4>@@$( (' *&'6*&/3'5 %#B*/3
&/ 5*&/%&**'G@ % 33%/'3?;33 3%&68% 3#+
7/3y 3&%3 3* %3'&3)#u68%#*.u %&D
.&3)6V*&/3d( (' *&')+X7/3y  3#]' 3)%&D
(1 3*/$#9@ ;3Y*%&3) %
#'(3V*/3( 3'$D
*&3).#$ @'5T'3#Y;3/'#*&/3*&%G4 % 33%3%
5?3+7/3y. *&* *&3( ' %.]&/54& 3
#$&(%&3) ' (3))23)& 3(  %&3B5*//G/G%<D
#$3*)+7/'3. * y 6?*&/3.;$*'3)#d%&3'*<Y
G3)3%< .3% G
$#+

E!v
K$v$/&JvRjRE86
3)'(</  %
'G 6= $*<*.' '#]A@ .D
' (68%
/ 54 h7/
3 w*&/3)
;%G
HH.H  %$/' "2
,.%& 6$/ 3)&3)%(</"2Y.+@2
I $2B4+BE2
 %<(</RHHF'2
 "
+ .FDCE

T**3 $*Y*9(*&%;$*&3*&R*/3 68%&*&/3)%V#33). $D


3*  6w=?A( (' *&' 6&/ '5*/e*&'%&G
% 33)%)2*&/'3$>]5$ 3? ;(*/3;$*< 3)#
3$ 3% 3).*< #*<'#5 %#$3*V%3)3*
*& *&3%3)*&3#R ' %&*&3))+

I!JRK$

7/3 *&/.%5.#Y68%&*&/3)%3$ %3)"*/3%G%<**'#$3


*&57 %<#1*&/3''#$3)& '*&3)%& X6 %-?#GV'#
.%(</
G Y68%768''#$G*&/'7 %3(*+

!JRK$

8$$<

WJJ&

&('5j
 B*.68.%=
3)*&'
 '#R!4G/$D
.35 %3 %&. 3)3%<

/ 93)(</'G. 3)&3)%(</"2
.+I
$2E C
F !
J
R$
_$
(vR
=?A # *&R68%@ %&3V % 33)%4'&'G
 R'*&%'(*%&3#R 3/R;'.3#4>@@R 3*/$#
%
(33#$'G.6Aw0 H.F2I *&/1+>w$D
,$t,.*#'-G.3)3%'Gh
 3%
=$683)%&3)'(3.2'!4."2'!57 
2 @$>B2,.
D
EE2'HH.F

 Ra

WJJ&

4'*&3#$ 
 3%( .3)*&G. *&6"*&/3w*%D
;3*7'5%&.'# */3(..*< 3%7&/ R $#$3
Y=7?5*/ '#5*/*4 %&. 3)3%2
%
(33#$'G.  !
6 w*&/3);'%&G HH.H27>  %D
/. .R 3%() $/ !
#%&$#$
' (
=?$QD>@='5
' 2$3%<$+E2

3 $*3 ;3)%@ H.H.
=?$QV  
'3%<$+ 2

3 $*3 ;3)%@ H.HF
JJ&Ij
75D3.*&
3)#'#$ D
&(.&*
*%&;'3)'(3
$#$3)?68.%3G.'33%G  (*.')2
>4>>4D,.'%&'2FC.2KE  I
C Y=@'*  
=aE  H'2
2     2
F DC2$$E8 IE)F

G3j
Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

G'%&3 E-?$#$ '6Y=7

G.%3YJ % 33)%7G3. 3*&%

G.%3J
F @$ 3)%& 3*&3* 68.%Y:1A4> 3)&%3D
3*)2'&#3w
35

G%3B
I $ 3% 3).*< 3*& 68%Y:1A> 3).'%&3D
3).*<26*4
3)5

G%34-4' 3)%&( 3)&/"2 % 33)%

G'%&3 J 3%()  3)&/"2'Y=7

G3
Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

KCS model scale

CR*10^3 , CTM*10^3

3
2
1
0
8

10

12

14

16

18 20
Vs [kts]

22

24

26

28

G'%&3 E)H3)&*'(3Y(%&.32Y=7

G%3! 24' 3)%&( 3)&/"2'&/ 95*&/9 % 33)%

Open water test: p1356 (KP505)

KCS model scale

-0.2

0.8

KRISO resistance test


SVA resistance test
SVA propulsion test

0.6

0.4

trim angle []

Kt , 10Kq , ETA0 [-]

CR KRISO
CT KRISO
CR SVA
CT SVA

Kt (SVA)
10Kq (SVA)
ETA0 (SVA)
Kt (SRI)
10Kq (SRI)
ETA0 (SRI)
Kt (KRISO)
10Kq (KRISO)
ETA0 (KRISO)

0.2

0
0

0.2

0.4
0.8
0.6
Advance Coefficient: J [-]

-0.1

0
8

12

16

20
Vs [kts]

24

28

G.%&3 E.EA
' (w*%& 92'V=7

G%3YCJ % 33)%7 395? *&3)%*3)*

KCS model scale

Open water test: p1356 (SVA)

0.5
Kt
SVA
10Kq SVA
Kt
CFD
10Kq CFD

0.6
Wake: w=1-u/U [-]

Kt , 10Kq

0.4
0.3
0.2
0.1
0
0.5

0.5
r/R=0.40
r/R=0.70
r/R=1.00
r/R=0.40
r/R=0.70
r/R=1.00

0.4
0.3

probe
probe
probe
LDV
LDV
LDV

0.2
0.1

0.6

0.7
0.8
0.9
Advance Coefficient: J [-]

G.%33&%3)#B'#( *&3)# % 33% 3

57*&3)%7*&3*

0
0

45

90
135
Angular position []

180

G%3EJ3&%3)#L $R57 .3 % 33)%


'32'Y=7

G3VC
Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

KCS model scale: radial wake in propeller plane

KCS model scale: axial wake in propeller plane


r/R=0.40 mesh1
r/R=0.70 mesh1
r/R=1.00 mesh1
r/R=0.40 mesh2
r/R=0.70 mesh2
r/R=1.00 mesh2
r/R=0.40 mesh3
r/R=0.70 mesh3
r/R=1.00 mesh3
r/R=0.40 mesh4
r/R=0.70 mesh4
r/R=1.00 mesh4

0.5
0.4
0.3

Radial wake: Vr/Vm [-]

Axial wake: 1-Vx/Vm [-]

0.1

0.2
0.1
0
0

45

90
135
Angular position []

G.%3E)J
F 1=.  '$*&3#Y$ 57 3.#$3%3*"G.%&#
R '%&. 3)3%  '3

probe
probe
probe
CFD
CFD
CFD

-0.1

-0.2
0

180

r/R=0.40
r/R=0.70
r/R=1.00
r/R=0.40
r/R=0.70
r/R=1.00

45

90
135
Angular position []

180

G%3E 3).'%&3#y'#T( *&3)#d%<#$ 57 .3


9 % 33)% 3

0.6

Vx/Vm -1, Vt/Vm, Vr/Vm

Axial wake: 1-Vx/Vm [-]

KCS model scale: axial wake in propeller plane

0.5
0.4

r/R=0.40
r/R=0.70
r/R=1.00
r/R=0.40
r/R=0.70
r/R=1.00

0.3
0.2

probe
probe
probe
CFD
CFD
CFD

0.1
0
0

Vx/Vm -1 LDV
Vt/Vm
LDV
Vr/Vm LDV
Vx/Vm-1 CFD
Vt/Vm CFD
Vr/Vm CFD

0.5
0.4
0.3
0.2
0.1
0
-0.1
-0.2

45

90
135
Angular position []

180

G.%3YEI3&%&3#'#(. $*&3#$ 57 .3

% 33)%  3

90
180
270
Angular position of propeller []

360

G%3E9 4V3
(* d(.  '3*)21 .&*.H  H D2
%\  HC+
2$H+EA;3/''#9 '%&. 3)3%  '3

KCS model scale: tangential wake in propeller plane

0.4
r/R=0.40
r/R=0.70
r/R=1.00
r/R=0.40
r/R=0.70
r/R=1.00

probe
probe
probe
CFD
CFD
CFD

Vx/Vm-1 , Vt/Vm , Vr/Vm

Tangential wake: Vt/Vm [-]

0.1

-0.1

-0.2
0

45

90
135
Angular position []

180

G.%3E43&%&3#'#( $*3)#*G3)*& 

57 3wR '%&. 3)3%  '3

Vx/Vm-1
Vt/Vm
Vr/Vm
Vx/Vm-1
Vt/Vm
Vr/Vm

0.3
0.2

LDV
LDV
LDV
CFD
CFD
CFD

0.1
0
-0.1
-0.2
0

90
180
270
Angular position of propeller []

360

G%3BECJ3$(* ( 3)*)2 .&*&|H  H D2


%\  HC+
2$H+EA;3/''#9 '%&. 3)3%  '3

G3j
Copyright National Academy of Sciences. All rights reserved.

0.5
0.4
0.3

Vx/Vm-1
Vt/Vm
Vr/Vm
Vx/Vm-1
Vt/Vm
Vr/Vm

LDV
LDV
LDV
CFD
CFD
CFD

Vx/Vm-1 , Vt/Vm , Vr/Vm

Vx/Vm-1 , Vt/Vm , Vr/Vm

Twenty-Fifth Symposium on Naval Hydrodynamics

0.2
0.1
0
-0.1
-0.2

90
180
270
Angular position of propeller []

360

Vx/Vm-1
Vt/Vm
Vr/Vm
Vx/Vm-1
Vt/Vm
Vr/Vm

0.5
0.4
0.3

LDV
LDV
LDV
CFD
CFD
CFD

0.2
0.1
0
-0.1
-0.2
0

90
180
270
Angular positionof propeller []

360

G.%3 EKJ
3)$(*
( .3*<2 *&
H  E)C.HD2$%\  H+C
2$H+EAa;3)/# '%&. 3)3%  '3

G%34
3$(*
( 3)*)2 .&*.
H  H D 2\%   H'+G $2H'+E Aa;3/'#R % 33)% 3

G.%3HJY3)$(* 3#dH'+E A;3/'#T % 33%


 3w68.% *&6  HD@ 6;.#$3E.2:1A

G%3
F V3
(* '3#H+EA;3)/'#T % 33)%
'3V68% .&*.  . HD@6;#3E2=?A

G.%3EY3)$(* 3#dH'+E A;3/'#T % 33%


 3w68.% *&6  IHD@ 6;.#$3E.2:1A

G%3IV3
(* '3#H+EA;3)/'#T % 33)%
'3V68% .&*.  I.HD@6;#3E2=?A

 G.3EH
Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

St. Johns, Newfoundland and Labrador, CANADA, 8-13 August 2004

Comparison of Detailed Simulations of a Turbulent


Ship Wake on a Straight and Circular Track
Ibrahim Yavuz, Zeynep N. Cehreli, Ismail B. Celik
(West Virginia University, USA)
ABSTRACT
The goal of this study is to numerically
simulate the wake behind a ship, namely the DDG51,
and investigate the flow physics, in general, and in
particular, the development of turbulence while
cruising on a straight and circular track. Detailed
simulations have been performed using serial and
parallel versions of a large eddy simulation (LES)
code. The ship model data used for these simulations
are those of DTMB 5512 and DTMB 5415. The ship
hull and the near wake are excluded from the
calculations due to computational expenses and the
computations were started from a plane behind the
ship. This is accompanied by implementing a Random
Flow Generation (RFG) technique originally developed
at West Virginia University, which calls for a time
averaged flow field at the inlet data plane to produce
instantaneous turbulent inlet plane flow fields for LES.
Reynolds
Averaged
Navier-Stokes
(RANS)
calculations are used to provide the RFG procedure
with the information needed for the inflow boundary.
The further development of the wake flow is calculated
using the large eddy simulation technique. LES of high
Reynolds number flows with complicated geometries
are enabled by this combined approach. Therefore, the
effect of the bluff body on the flow field is embedded
in the mean flow prescribed at the inflow plane. The
LES data generated is used to analyze the development
of turbulence within 2-3 ship lengths in the near wake.
INTRODUCTION
Wake flows are characterized by rapid
streamwise
adjustments
and
are
strongly
inhomogeneous along all three directions (Shi, 2000).
In the literature, there are numerous descriptions of a
ships wake; due to its complexity it is not well
understood and ship wake flow dynamics is still an
ongoing research area. William Thompson (Lord
Kelvin) (1887) was the first to explain the wave pattern
generated on the water surface by a moving

disturbance. The hydrodynamic phenomenon which


occurs in a ships wake is a result of the generation of
turbulence by the ship hull and its peripherals. This
wake develops and grows in time and characterizes the
scale of turbulence. The properties of the ship wake
depend on the size, shape, the speed, and the
propulsion system of the ship.
Figure 1 illustrates a hydrodynamic wake at
sufficiently high speeds. It consists of white water
(wake generation region), viscous wake, propeller
wake, and Kelvin wave pattern. The white water (local
wave disturbance region) originates at the region
around and behind the ship up to two or three ship
lengths. It is the region where surface foam
(entrainment of the air at and below the free surface),
bubbles, and strong turbulence is generated and rapidly
decayed. The amplitude of these waves depends
strongly on the shape, speed and the propulsion system
of the ship. This near wake region is also the place
where these features decay rapidly. The viscous
(turbulent) wake extends up to the far field region and
the flow is characterized by the wave drag and the drag
of the hull. It is this region where large-scale vortical
flows occur. A turbulent wake can be several
kilometers long and the exact structure of this smooth
water region depends on the ship. The turbulence
behaves like an additional viscosity that dampens the
surface waves (Milgram et al., 1993). The bubbly
turbulent wake is one of the main features to be used
for detecting ships. In the open literature, there are
some limited numerical experiments and full scale
measurements of the turbulent wake behind a ship.
Swean (1987) solved the steady, parabolic,
incompressible time averaged Navier-Stokes equations
for the high speed combatant wake, normalized by its
beam length and its velocity to obtain the wake profile
by using k- turbulence model. Meadows et al. (1994)
measured the Lagrangian velocity profiles and
compared the experimental and numerical longitudinal
surface wake velocities, normalized by ship length and
velocity.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

The Kelvin wave pattern (V-shaped) consists


of two arms (Kelvin arms). It consists of the
superposition of waves, transverse and diverging
wave crests, which propagate in a direction at 145
degrees to the positive x-axis (Hoekstra & Ligtelijn,
1991). It is a characteristic surface wave pattern
resembling a wedge-shaped region behind the ship
(Melsheimer et al., 1999). It plays an important role in
the development of the wake that is imaged by
synthetic aperture radar (SAR) (Reed et al., 1990).
Many researchers have studied Kelvin wakes both
experimentally and theoretically, however, the
experimental results have mostly been done for
idealized mathematical hull forms, that have few
characteristics of real ships (Reed et al., 1990).
Milgram (1988) also analyzed the Kelvin wave system
hydrodynamically. In the far region, the wake decays
relatively slow and steadily, directly proportional to the
axial distance from the ship. The wave amplitudes also
depend on the ship. Further downstream, the surface
foam and bubbles decay and the surface roughness and
thermal characteristics gradually return to those of the
surrounding ambient surface (Reed et al., 1990). In the
far wake of a ship, the effects of turbulent dissipation
are minimal (Dommermuth et al., 1993).

acting like a simple plane of symmetry. Naudascher


(1965) studied the wake of ship-like bodies. He
observed that the wake width had a power law
behavior. Buller & Tunaley (1989)s measurements
have shown that the power law behaviour was x1/4. This
was also indicated by Milgram et al. (1993) who made
field measurements for ship wakes. They found that the
wake width had a power law behavior of x1/5 where x is
the axial distance from the ship. Hoekstra & Ligtelijn
(1991)s experimental study indicated that the turbulent
kinetic energy had an asymptotic behavior of x-4/5.
Dommermuth et al. (1996) studied free surface flows
using LES. They showed the probability distributions
of the velocity field in the wake. Comparison of their
simulations with the results from the field and
laboratory experiments showed good agreement. They
predicted the kinetic energy, the attenuation of the
mean axial velocity field and the enhanced spreading
of the wake near the free surface.
Hoekstra and Aalbers (1997) have done an
experimental study on the structure of the wake of
unpropelled hulls and found very interesting structures
of these nominal wakes. Without the effect of the
propeller, they observed the presence of longitudinal
vortices, two pairs of counter rotating vortices, namely
bilge vortices (close to the stern in the bilge region),
one on either side of the symmetry plane, close to each
other and they both tend to move slowly downward.
The other pair, called side vortex pair generated near
the free surface is observed to be far apart. These side
vortices tend to move away from the longitudinal
symmetry plane of the ship as the distance gets far
from the stern. As these vortex pairs drift in the
spanwise direction, they get weaker and eventually
disappear. In the longitudinal direction, the pattern of
the velocity contour plots showed an upside-down
rimmed hat shape, two side regions close to the free
surface and one in the center. Hoekstra and Aaelbers
(1997) found that the smallest axial velocity was in the
wakes symmetry plane near the free surface.
Experiments indicate that when strong side vortices are
produced, the minimum value occurs below the free
surface, but near the cores. Benilov et al.s (2000)
experimental study and the numerical modeling have
indicated that they could detect the ship wake
turbulence well, and Kolmogorovs inertial range could
be identified, even in the far wake region. The
production and strength of these vortexes are very
much dependent on the ships shape. Sometimes only
one vortex pair can be observed, either bilge or side
vortex (Reed et al, 1990). Particularly, the nominal
wake is symmetric near the free surface; however,
propeller, rudder and combination of bilge vortices
usually produce strong asymmetries.

Figure 1. Schematic sketch of a ship wake (Hennings et


al., 1999)
The vorticity dynamics of wake flows is a
complex phenomenon. The interaction of an
underwater jet with a free surface is a simpler flow that
has similar characteristics to the wake of a body
interacting with the free surface (Reed et. al., 1990).
The jet is laminar at the beginning, then vortex rings
quickly develop around the jet, because of the
interaction of the jet with the free surface, then
transition to turbulence starts to develop. In addition, in
Madnias (1989) experimental study on a round jet near
a free surface, it is seen that the free surface is not

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

For wake signatures, turbulence activity near


the free surface is of great importance. The convenient
measure of the turbulence is the turbulent kinetic
energy, which is observed to be insensitive to the ship
shape. As stated by Hoekstra and Aalbers (1997), the
maximum values of the root mean square fluctuations
for some locations in units of (m/s) are found to be,

discretization schemes, the fractional step method,


poisson solver, etc., the reader is referred to Cehreli
(2004) and Shi (2001).
Governing Equations and Navier-Stokes Solver
The LES code used was originally developed
by Zang et al. (1994). The spatially filtered flow
conservation equations are

0.106 0.02 at x/L=0.25

0.067 0.01 at x/L=0.60

0.047 0.01 at x/L=1.00


where L is the ship length.
To the best of our knowledge there are neither
detailed simulations nor experiments concerning the
wake of a turning surface ship. This study uses LES
data to analyze the turbulent flow dynamics in the
wake of a ship cruising with constant speed on a
straight track and on a circular track in a comparative
manner.

u j
x j

(1)

=0

ui Fij
+
= Si
t
x j

(2)

Where
Fij = ui u j + p ij

COMPUTATIONAL APPROACH
Numerical Method
The computer code is based on an essentially
non-staggered grid, finite volume method using a
fractional time step approach. Non-orthogonal
curvilinear coordinates are applied with an overall
second order accuracy in both space and time. The
Crank-Nicolson discretization scheme has been applied
for diagonal viscous and diffusion terms and an explicit
Adams-Bashforth scheme is employed for other terms.
The central differencing (CD) scheme (with special
care due to numerical instabilities) is applied to
discretize the convective terms. Detailed information
can be found in Shi (2001).
The current solution method is 2nd order
accurate both in space and time. Theoretically the
higher the order of the numerical scheme, the better the
resolution should be under the same grid spacing (Shi,
2001). However, Rai and Moin (1991) have shown that
higher order of accuracy combined with coarse grid
spacing does not necessarily give better results. Jordan
(1999) showed that the results could be improved by
improving the grid spacing. The power law scheme is
inaccurate under some limitations, in that when
convection is dominant; it reduces to 1st order upwind
scheme (Patankar, 1980). On the other hand, higher
order CD schemes have in addition the problem of
artificial high frequency oscillations that may
contaminate the turbulence field (Rai and Moin, 1991).
In LES, explicit schemes are preferable, but if stability
is an issue, some implicitness can be introduced, i.e.
Crank-Nicolson time splitting. For information on the
LES code, i.e. equations, time advancement and spatial

ui
+ ij
x j

(3)

x
x
S i = 2 S u1 3 u3 1

xi
x
i

(4)

ij = ui u j ui u j

(5)

In the above equations, uj is the filtered


velocity vector, p is the reduced dynamic pressure in
which the total pressure, P, is calculated as
1
P = p + 2 r 2 , where r is the distance to the
o

rotation axis, x1 is the axial coordinate, x2 is the vertical


coordinate, x3 is the transverse coordinate, s is the
angular velocity of the system rotation due to the
turning ship, is the kinematic viscosity and ij is the
Kronecker delta (See Fig. 3 for coordinate system). It
should be noted that the effect of centrifugal force is
included in the total pressure term and the variation of
the density in the source term is neglected. In Eqn. (4),
the right hand side (RHS) represents the Coriolis force
due to the turning of the ship. Here, ij is the subgrid
scale stress tensor and represents the effects of the
subgrid-scale motion. Since the discrete solution
represents the resolved field, ij cant be explicitly
computed and has to be modeled using the resolved
quantities. The model for ij is as follows:

ij ij kk = 2t Sij
3

where

t is

(6)

the turbulent subgrid-eddy viscosity

and Sij is the resolved strain rate tensor,

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

S ij =

1 ui u j
+
2 x j xi

on the local clusters of the CFD-Laboratory or on the


clusters at the Pittsburgh Supercomputing Center.
Currently, around 10 million nodes can be used in these
simulations with a reasonable turn-around time. The
details of parallelization technique will be presented in
a future report (Celik, 2004) to be submitted to Office
of Naval Research (ONR).

(7)

To tackle problems of complex geometries,


the above mentioned equations are transformed from
the physical to the computational space and formulated
for a generalized curvilinear coordinate system.
The SGS-eddy viscosity is calculated using
the standard Smagorinski (1963) model given by

t = Cs ( Sij Sij )
2

1/ 2

APPLICATIONS

The DDG51, which is approximately 154 m


long and 20 m wide has an assumed average speed
around 20 knots. The ship model data used for this
simulation was taken from the data for DTMB 5512
(Longo et al., 1993; Gui et al., 1999) and DTMB 5415
(Toda et al., 1988; Stern et al., 2000), for the wake of
ship on a straight track and circular track, respectively.
As stated in Shi (2001) and Yavuz et al. (2002), the
ship hull and the near wake are excluded from the
calculations due to computational expenses and the
computations were started from a plane behind the ship
(Figure 2(b)).

(8)

where is the volume average box filter used by


Deardoff (1970) usually calculated as the geometric
average of mesh spacings in the Cartesian directions,
defined

as

= ( x1x2 x3 )

1/ 3

and

Cs is

the

Smagorinsky constant; here x1, x2, x3 represent the


sides of a octahedral computational cell. Sij is the large
scale (resolved) strain rate tensor, defined as,
1 u u j . The constant has been found to be
Sij = i +

2 x j xi
a function of the filter width and the integral length
scale of turbulence. Cs = 0.2 does a good job for
isotropic turbulence, and for non-homogeneous flows
this value must be reduced by half or more (Ferziger,
1993).

(a)

Random Flow Generation (RFG) Methodology

The Random Flow Generation (RFG)


approach provides turbulent time dependent initial or
inlet flow conditions for LES. In this technique a
divergence free vector fields from a sample of Fourier
harmonics are synthesized, which allows the generation
of a non-homogenous an-isotropic flow field
representing turbulent velocity fluctuations. The details
of the algorithm can be found in Smirnov et al. (2000,
2001) and Shi et al. (2000).

(b)

Figure 2. a) Geometry of the DTMB 5512, b)


Dimensions of the straight ship calculations in terms of
ship length, L.
LES of Ship Wake Cruising on a Straight Track

Parallel Computations

The ship wake simulations for the straight


track are for the ship model DTMB 5512, a smaller
ship model scaled from DTMB 5415. Model 5415 is a
towing tank model representing a modern naval
combatant.
It
measured
5.72
m
between
perpendiculars. Model 5512 measures 3.04 m between
perpendiculars. The geometry of this ship model is
shown in Fig. 2(a). In order to avoid solving for the
flow around the ship model, the computational domain
starts from the inflow boundary (or initial data plane)
located immediately after the body in the wake, where

To improve turn-around time of the


computations and allow LES in a larger domain,
customized message-passing routines were written and
tested on the basis of MPI to provide a simple
language-independent interprocessor communication.
Both parabolic domain decomposition and elliptic
domain decomposition methodologies were used for
efficient parallelization of the LES code for wake flow
applications. These simulations were performed either

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

ships stern is at (0.5, 0, -3.0) in x-, y-, z- directions


respectively.

there is no flow-reversal. The pseudo-random flow


field generated by the RFG method is added to the
mean flow of RANS simulations to establish the
unsteady boundary conditions at the inlet plane. The
whole ship flow including the ship wake can be
sketched as in Fig. 2(b). The computational domain
starts from x/L = 1.05 (where x starts from the front of
the ship model). At this plane the RFG method is used
in conjunction with the RANS calculations (Stern and
Wilson, 2000).
At the inflow boundary all components of the
velocity are specified as a function of time and space.
At the outflow boundary Neumann (free gradient)
boundary conditions are applied. Symmetry conditions
have been used in y direction and periodic boundary
conditions have been used in the spanwise (z)
direction. At the free surface a slip condition is allowed
in x and z directions, but the velocity component
normal to the free surface is set to zero. As such the
free surface is approximated as a moving flat plane, i.e.
low Froude number approximation is involved.
The computational domain size is 3.0x0.3x0.5
(given in non-dimensional units in ship length) in x, y
and z-directions (axial, vertical and transverse
directions), respectively. The grid size is 108x50x66
per processor multiplied by 10 processors, which sums
up to ~3 million grid nodes. Non-uniform grid spacing,
stretching smaller than 1:03, is used in both z- and ydirections. The length scale and time scale used in RFG
were selected as constant in this case. The length scale
was 0.02 of the ship length, and the time scale was
0.001, non-dimensionalized by free stream velocity and
ship length. Those numbers were selected because the
turbulence length scale is estimated to be about 15% of
the ship width. The Smagorinsky SGS model, and
second order central differencing scheme were used
unless otherwise stated.

(a)

(b)

(c)
Figure 3. Turning ship wake: a) The geometry b)
velocity profile specified at the IDP (top view) c) The
coordinate system and numerical mesh (Only the core
region is shown, distances are non-dimensionalized
with ship length, L)

LES of Ship Wake Cruising on a Circular Track

The ship model studied on a circular track is


DTMB 5415, a 5.72m long model of the Navy DDG51
surface ship (Stern et al., 2000). A computational
domain of 1.75x0.15x1.0 (given in non- dimensional
units in ship lengths and it starts from x/L=1.50, where
x/L=1 is the end of the ship model) and a grid of
130x33x66 vertices in x (axial), y (vertical), z
(transverse) directions has been used to represent the
near wake region. More grid distributions were tested
and reported by Cehreli (2004). It should be noted that
a thorough study was not conducted as to whether the
side boundary is sufficiently far from the wake
centerline. In the IDP (inlet data plane), the RFG
method (Smirnov et al., 2001) is used in conjunction
with the RANS calculations (Hyman, 2001) and the

The core region of the numerical grid and the


geometry of the ship model are illustrated in Figure 3.
The ship turns with a dimensionless angular velocity of
1/3 (Figure 3(b)) with a radius of curvature
corresponding to three ship lengths, which will result in
a dimensionless ship velocity of 1. Only a 30o turn was
investigated. The Reynolds number used, 1.0x107, in
the simulations is based on the model ship length. .
Here, Reynolds number similarity is assumed, which is
attained by changing the laminar viscosity of the fluid
in the simulations. The coordinate system used is with
respect to an observer on the ship. The Smagorinsky
model was used with Cs=0.205 with a central
differencing scheme. The length scale and time scale

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

formations are very much alike for all ship hulls.


Figure 4 indicates that the predicted straight ship wake
physics are in good agreement with the macro wake
measurements at the same location. In addition, axial
turbulence intensities are also in very good agreement
with the measurements (Hoekstra& Ligtelijn, 1991), as
shown in Table 1.

used in RFG are calculated from =k/ and l =


0.09k1.5/, non-dimensionalized by free stream velocity
and ship length. Non-uniform grid distributions are
used in this study in both x and z directions with
expansion ratio not exceeding 1.03. The time step is
0.001 for all grid resolutions.

Table 1 Comparison of the maximum values of the


root mean square fluctuations for some locations in
(m/s) for the wake of a ship on a straight track (serial)

for any ship hull


maximum value of root mean square fluctuations
location
experimental
numerical
(values obtained *Us)
(m/s)
(m/s)
LES calculations start
x/L=0.25
0.1060.02
from x/L=0.5
x/L=0.6
0.069
0.0670.01
x/L=1.0
0.057
0.0470.01

Figure 4. a) Predicted velocity contours at x/L=1.20 for


ship on a straight track (Shi , 2001), b) Macro wake
measurements at x/L=1.20 by Hoekstra& Ligtelijn
(1991) for ship model No. 5452 (Re~1x107)
RESULTS AND DISCUSSIONS
Straight Track Simulations

Figure 5. Temporal history of streamwise velocity


components at x = 0.6, y = -0.012, z = -0.06 (parallel)

The results of the wake simulation of a ship


on a straight track are compared (Fig. 4) with the
straight track wake measurements studied earlier by
Hoekstra & Ligtelijn (1991). Two stable large bilge
type vortices are observed for the straight track wake
simulation using the standard Smagorinsky model
(Re=4.65x106) at x/L=0.65. Additional smaller side
vortex pairs are observed away from the center of the
wake. These vortex pairs drift in the spanwise
direction, get weaker and eventually disappear. Figure
4 shows a comparison of the predicted velocity
contours with the macro wake measurements at
x/L=1.20 by Hoekstra& Ligtelijn (1991) for ship model
No. 5452 (Re~1x107) with the non turning ship wake at
the same location. For both studies, minimum axial
velocity occurs near the free surface of the center of the
wake. Both axial velocity contours look like an upside
down rimmed hat formation, where two side lobes
close to the free surface and a central lobe is observed.
As stated by Hoekstra& Ligtelijn (1991), these

From this point on the parallel straight track


simulations are addressed. The unsteady streamwise
velocity variation is presented in Figure 5, where is
the flow trough time. This point was selected inside the
wake, as seen from the velocity defect. Figures 6 and 7
show the vertical variations of the mean velocities and
the root mean square velocities, respectively, where the
time averaging was made over 3. Figure 6 shows that
the spanwise velocity at the center does not change
much in the axial direction, but same velocity
components off centerline become more uniform as the
wake develops. The vertical velocity at the centerline
has a peak below the free surface, which gets weaker in
the axial direction. Significant changes are observed in
the streamwise velocity profiles in both axial and
vertical direction. The location of peak axial velocity
moves deeper into the wake indicating a plunging
effect as the wake develops in the streamwise direction.
This peak is stronger off the centerline indicating the
6

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

strength of one of the bilge vorticities seen in Fig. 8.


The intensity of the rms velocity fluctuations show
very little decay but significant redistribution and
become more isotropic as the wake develops. Initially
the turbulence is concentrated more near the free
surface. This area plunges deeper with increasing axial
distance.
The development and spreading of the wake is
presented in Figures 8 and 9. The two large bilge
vortices first separate from each other, then further
downstream they gradually merge together as they
loose strength. However, the turbulent kinetic energy
seems to increase in certain regions as seen in Figures
8b and 9b, even towards the end of the domain (see
also Fig. 19). This indicates that the domain should be
longer than the selected 3 ship lengths. Also, one has to
account for the wake spreading and a deeper and wider
cross-section is necessary to accommodate the whole
wake further downstream than 3 Ship lengths.
In addition to these, the wake spreading or
wake width, which is obtained to be roughly w~x1/4 is
consistent with Buller and Tunaley (1989)s

measurements. Milgram et al. (1993) and Hoekstra&


Ligtelijn (1991) found w~x1/5. And the spreading rate
of the ship wake on a straight track is observed to be
significantly higher than that of a ship wake on a
circular track. This is well observed from the parallel
simulation predictions of the time averaged streamwise
velocity magnitude contours in Figure 8 for a straight
track ship wake using 3 million grid nodes on 10
processors (Intel Pentium4) on a domain of 3 ship
lengths. These parallel computations were also
expanded to a 6 million nodes simulation on 6
processors (3rd generation DecAlpha), and more
detailed structures were captured. However, the turn
around time for these simulations has significantly
increased. The predicted extra detail is especially
pronounced in the vorticity contours presented in
Figure 10. The 6 million nodes simulation was still in
the early stages, therefore one should not fall into the
wrong conclusion, that the wake spreading is much
more pronounced. In fact, it will tighten as the
simulations continue.

(a)

(a)

(b)

(b)

(c)

(c)

Figure 6. The root mean square velocity profiles at 3


spanwise locations at a) x/L=0.09, b) x/L=1.60, c)
x/L=2.80 (parallel)

Figure 7. The root mean square velocity profiles at 3


spanwise locations at a) x/L=0.09, b) x/L=1.60, c)
x/L=2.80 (parallel)

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

(a)

(b)
Figure 8 Contours on vertical x-planes of time averaged a) streamwise velocity b) turbulent kinetic energy (parallel
simulations, 108x50x66 nodes per processor)
(a)

(b)

Figure 9 Contours on horizontal y-planes of time averaged a) streamwise velocity b) turbulent kinetic energy
(parallel simulations, 108x50x66 nodes per processor)

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

(a)

(b)

Figure 10 a) Vorticity magnitude contours obtained for the wake behind a ship on a straight track using parallel
computations with 3 Million grid nodes, b) Preliminary vorticity magnitude contours obtained for the wake behind a
ship on a straight track using parallel computations with 6 Million grid nodes (black lines show processor
boundaries, overlap/communication regions)

weaker, however they do not disappear. LES captures


most of the vortical structures as also seen in Figure 13.

(a)

Figure 11. Temporal history of streamwise velocity


components at 10 (x= 1.02 y=-0.001 z=-3.154881)
(b)
Circular Track Simulations

The unsteady velocity fluctuations show an


approximate resolved frequency to be around 24 Hz in
all three directions with standard Smagorinsky model
as seen in Figure 11. In Figure 12 the mean axial
velocity contours are presented, it seems like the
structures diffuse somewhat in the downstream
direction, i.e. big vortices in the IDP become somewhat

Figure 12. The mean axial velocity color contours at a)


IDP (Hyman, 2001) b) an angle of 5o (x/L=0.80)
9

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

The predicted velocity contours for the ship


wake on a circular track show that one large bilge
vortex that moves downwards is observed and is
probably the result of a merger of two vortices under
the action of Coriolis and centrifugal forces. There is a
smaller circulation region (a side vortex) on the outer
rim of the wake. The streamwise flow field causes the
side vortex to weaken and the wake decays in the outer
region of the near wake similar to the straight track
wake simulations. Moreover, the Coriolis force seems
to generate more energetic and less diffusive eddies,
hence it seems to increase the kinetic energy content of
the wake (Yavuz et al., 2002).
Figure 14 shows the mean velocity profiles at
3 vertical locations namely at the IDP (Hyman, 2001),
an angle of 5o (x/L=0.80), and at 15o (x/L=1.30).
Moreover, Figure 15 shows the mean velocity profiles
at 3 spanwise locations selected at IDP (Hyman, 2001),
an angle of 5o (x/L=0.80), and at an angle of 15o
(x/L=1.30).

(a)

(b)

Figure 13. The mean velocity line contours at an angle


of 5o (x/L=0.80, zoomed in) a) axial b) vertical

(a)

(a)

(b)

(b)

(c)

(c)

Figure 14. The mean velocity profiles at 3 vertical


locations at a) IDP (Hyman, 2001) b) an angle of 5o
(x/L=0.80) c) an angle of 15o (x/L=1.30)

Figure 15. The mean velocity profiles at 3 spanwise


locations at a) IDP (Hyman, 2001) b) an angle of 5o
(x/L=0.80) c) an angle of 15o (x/L=1.30)

10

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

And Figure 16 shows the root mean square


velocities along the depth of the wake at angle of 5o
(x/L=0.80) and at angle of 15o (x/L=1.30)
respectively. Both figures show that at the free surface,
for both models, the vertical fluctuations are damped
and the energy is redistributed to spanwise and
streamwise directions, as expected. The vorticity
contours on an horizontal plane are presented in Figure
17. In the far wake region vorticity decay is observed
due to the coarser grid resolution. It should be pointed
out that the turbulence observed is asymmetric across
the wake and there is a sharp edge on inside of the turn.

x/L

Figure 18 Resolved turbulence kinetic energy of the


wake simulation of a ship cruising on a straight and
circular track normalized w.r.t. its inlet value

(a)

(b)
Figure 19 Resolved turbulence kinetic energy of the
wake simulation of a ship cruising on a straight track
normalized w.r.t. its inlet value along a line of center
z = 0.055, y=-0.012 (Parallel computations)
Comparisons

The resolved turbulence kinetic energies of


the wake simulations (on a single processor) of a ship
cruising on a straight and circular track are presented in
Figure 18, where a significant decay of kinetic energy
in the streamwise direction can be observed. The
locations for both studies are taken to be at the highest
kinetic energy value obtained from the IDP (along the
centerline for the non turning ship and at y= -0.1E-02
and z=-3.18 for the turning ship wake simulation).
Comparing to the predictions of a non-turning ship
wake, the turning ship case indicates less kinetic
energy values. However, it should be noted that, this
may be due to the coarser grid resolution in the far
wake. Overall the trends are similar, but there is a
sinusoidal-like distribution of the TKE prediction in the
near wake for the straight ship case. It may be because
of the existing surface wave from the RANS
calculations (Stern and Wilson, 2000). This indicates
that some wave information may be present implicitly
in the inflow boundary. However, this sinusoidal-like
distribution of TKE is not seen in the simulations of a
ship on a circular track, which may imply that there is

Figure 16. The variation of root mean square velocities


with depth at a) an angle of 5o (x/L=0.80), b) an angle
of 15o (x/L=1.30)

Figure 17 The predicted vertical vorticity, y, contours


at y= -0.1E-02 in the ship wake

11

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

not any wave information present in the RANS


simulations (Hyman, 2001).
The same dip seen in Figure 18 is also present
in the 3 million nodes simulation predictions. However,
the kinetic energy presented in Figure 19 also indicates
that the kinetic energy is increasing in the downstream
direction. To have a better perspective, Figures 8b and
9b should be investigated. Here, the bilge vortices
separate and later merge closer as they become larger,
which is probably the cause for the turbulent kinetic
energy rise. It should be mentioned that some of the
initial decay in k could be the out effect of the
application of RFG at IDP. This issue is currently under
investigation.

REFERENCES

Benilov, A. Bang, G., Safray, A. and Tkachenko, I.,


2000, Ship Wake Detectability in the Ocean
Turbulent Environment, 23rd Symposium of Naval
Hydrodynamics, 1-15, Val de Reuil, France.
Buller, E.H. and Tunaley, J.K.E. (1989), The effect of
the ships screws on the ship wake and its implication
for the radar image of the wake, Proc. IGARSS 89,
362-365.
Cehreli, Z.N., 2004, Investigation of Ship Wakes using
LES with Various SGS Models, Ph.D. Dissertation,
West Virginia University.
Celik, I.B., 2004 Parallel computations of two-phase
turbulent flows, Office of Naval Research (ONR)
Report.
Dommermuth, D. and Novikov, E., 1993, Direct
Numerical and Large-Eddy Simulations of Turbulent
Free-Surface Flows, in Proceeding of the Sixth
International Conference on Numerical Ship
Hydrodynamics, 239-270, Iowa City, Iowa.
Dommermuth, D., Gharib, M., Huang, H., Innis, G.,
Maheo, P., Novikov , E., Talcott, J., and Wyatt, D.,
1996, Turbulent Free-Surface Flows: A Comparison
Between Numerical Simulations and Experimental
Measurements, 21st Symposium of Naval
Hydrodynamics, 200-215, Trondheim, Norway.
Ferziger, J. and Peric, M., 1997, Computational
Methods
for
Fluid
Dynamics,
Springer,
Mercedesdruck, Berlin.
Gui, L., Longo, L., and Stern, F., (1999),Towing Tank
PIV Measurements System and Data and Uncertainty
Assessment for DTMB Model 5512, 3rd
International Workshop on PIV, Santa Bar-bara, CA,
16-18 September.
Hennings, I., Alpers, W. and Romeiser, R., 1999,
Radar Imaging of Kelvin arms of ship wakes, Int. J.
Remote Sens., 20 (13), 2519-2543.
Hoekstra, M. and Aalbers, A., 1997, Macro Wake
Measurements for a Range of Ships, 21st
Symposium of Naval Hydrodynamics, 278-290,
Trondheim, Norway.
Hoekstra, M. and Ligtelijn, J. (1991), Macro Wake
Features of a Range of Ships, MARIN Report No.
410461-1-PV, Netherlands.
Hyman, M. (2001) Private Communications
Jordan, S., 1999, A large-eddy simulation
methodology in generalized curvilinear coordinates,
Journal of Computational Physics 148, 322.
Longo, J., Stern, F., and Toda, Y., (1993), "Mean-Flow
Measurements in the Boundary Layer and Wake and
Wave Field of a Series 60 CB=0.6 Ship Model - Part
2: Scale Effects on Near-Field Wave Patterns and
Comparisons with Inviscid Theory," J. Ship Research,
Vol. 37, No. 1, pp. 16-24.

CONCLUSION

Detailed LES calculations have been


performed for the developing wake of a surface ship
cruising on a straight track with 3 million nodes
(108x50x66 nodes per domain, x0.46m; y0.5m;
z1.2m in the wake region).
The results are analyzed and turbulence
statistics have been presented for further comparison
with RANS and/or experiments as they become
available. Qualitative comparison with experimental
observations indicate that the LES results are credible,
but rigorous validation is necessary to make definite
conclusions.
Relatively under-resolved LES of the wake of
a ship turning on a circular track has also been
performed and the results are compared to those of a
ship cruising on straight track.
It is seem that the wake of the turning ship is
much narrower and has more concentrated vorticity
due to the merger of the two bilge vortices. There is
little similarity between these two wakes. The wake of
the turning ship exhibits more dynamic futures. A
thorough comparison of the two cases will only be
possible when the grid resolution of the turning ship
wake is increased by employing parallel runs.
ACKNOWLEDGMENTS

This work has been performed under a DOD


EPSCOR project sponsored by the Office of Naval
Research (ONR). The technical program monitor is Dr.
Patrick Purtel. We would like to thank Dr. M. Hyman
for providing the RANS simulations for the ship wake.
We also acknowledge the partial funding kindly
provided by the WVU Research Cooperation, National
Research Center for Coal and Energy and the College
of Engineering and Mineral Resources.

12

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

On Naval Hydrodynamics, 17-22 Sept 2000, Val de


Reuil, France.
Stern, F. and Wilson, R., 2000b, RANS solution of
Ship Wake, University of IOWA, Private
communication.
Swean, T.F., Jr., 1987, Numerical Simulations of the
Wake Downstream of a Twin-Screw Destroyer
Model, Naval Research Laboratory Memorandum
Report 6131, 41 p.
Toda, Y., Stern, F., Tanaka, I., and Patel, V.C., (1988),
"Mean-Flow Measurements in the Boundary Layer
and Wake of a Series 60 CB=0.6 Model Ship With
and Without Propeller," IIHR, The University of
Iowa, IIHR Report No. 326.
Thompson, W. (Lord Kelvin), 1887, On Ship Waves,
Proc. Inst. Mech. Eng., [Popular lectures, iii.482];
Lamb, H., Hydrodynamics, (Dover, New York,
1985), 434.
Yavuz, I., Cehreli, Z. N., Celik, I. B. (2002) Large
Eddy Simulation of the Wake Behind a Turning
Ship, 2002 Fluids Engineering Division Summer
Meeting, July 14-18, 2002, Montreal, Quebec,
Canada.
Zang, Y., 1993, On the development of tools for the
simulation of geophysical flows, Ph.D. Thesis,
Stanford University.
Zang, Y., Street, R. L. and Koseff, J. R., 1994, A Nonstaggered Grid, Fractional Step Method for TimeDependent Incompressible Navier-Stokes Equations
in Curvilinear Coordinates, J. Comp. Phys. 114, 1833.

Madnia, K., 1989, Interaction of a Round Jet with a


Free Surface, Ph. D. Thesis, University of Michigan.
Meadows, L., Meadows, G., Troesch, A., Cohen, S.,
Beier, K.-P., Root, G., Griffin, O. M., Swean, T.F., Jr.,
1994, "Lagrangian Velocity Profiles in the Wake of a
High Speed Vessel," Ocean Engineering, 21, (2), 221242.
Melsheimer, C., Lim, H. and Shen, C., 1999,
Observation and Analysis of Ship Wakes in ERSSAR and SPOT Images, Proc. 20th. Asian
Conference for Remote Sensing, 22-25 November,
1999, Hong Kong, China.
Milgram, J.H., 1988, Theory of Radar Backscatter
from Short Waves Generated by Ships, with
Application to Radar (SAR) Imagery, J. Ship Res,.
32 (1), 54-69.
Milgram, J.H., Peltzer, R.D. and Griffin, O. M., 1993,
Suppression of Short Sea Waves in Ship Wakes:
Measurements and Observations, J. Geophysics Res.
98, No. C4, 7103-7114.
Naudascher, E., 1965, Flow in the wake of SelfPropelled Bodies and Related Sources of Turbulence,
J. Fluid Mech. 22 (1), 625-656.
Patankar, S.V., 1980, Numerical Heat Transfer and
Fluid Flow, Hemisphere Publishing Corporation, New
York, N.Y.
Rai, M. and Moin, P., 1991, Direct simulation of
turbulent flows using finite-difference schemes,
Journal of computational physics 109 (2), 169.
Reed, A. M., Beck, R. F, Griffin, O. M. and Peltzer, R.
D., 1990, Hydrodynamics of Remotely Sensed
Surface Ship Wakes, SNAME Transactions 98, 319363.
Shi, S., Smirnov, A. and Celik, I., 2000, Large Eddy
Simulations of turbulent wake flows, Twenty-Third
Symposium on Naval Hydrodynamics, 203-209, Val
de Reuil, France.
Shi, S. Towards Large Eddy Simulation of Ship
Wakes, Ph.D. Thesis, West Virginia University,
2001.
Smagorinsky, J., 1963, General circulation
experiments with the primitive equations; I. The basic
equations, Mon. Weather Rev 91, 99-164.
Smirnov, A., Shi, S., and Celik, I. (2000) Random
Flow simulations with a bubble dynamics model,
ASME Fluids Engineering Division Summer Meeting,
No. 11215, FEDSM2000, Boston, MA.
Smirnov, A., Shi, S. and Celik, I., 2001, Random
Flow Generation Technique for Large Eddy
Simulations and Particle Dynamics Modeling JFE
123.
Stern, F., Longo, J., Penna, R., Olivieri, A., Ratcliffe,
T., and Coleman, H., (2000), Interna-tional
Collaboration on Benchmark CFD Validation data for
Surface Combatant DTMB Model 5415, 23rd Symp.

13

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics


St. Johns, Newfoundland and Labrador, CANADA, 8-13 August 2004

Hydrodynamic Wakes of Surface Penetrating Structures


Alan Brandt, Stephen M. Scorpio, Eric A. Ericson and Charles E. Schemm
(The Johns Hopkins University Applied Physics Laboratory)

end, this study has focused on the white-water wake


generated by surface penetrating structures.
Several simultaneous processes contribute to
the white-water wake (WWW) generated by flow past
a surface penetrating structure. The turbulent drag
wake around the body (separation at the bow and stern
and the boundary layer and the near field recirculation
zone) results in air entrainment at and beneath the
surface. At the leading edge water run-up and splashdown lead to surface bubble entrainment, subducted
entrained air that resurfaces behind the object and
breaking surface waves generated by the flow around
the body. Enhanced near-field waves coupled with the
Kelvin wake further result in local breaking. Evidence
of each of these mechanisms has been found, however
it is often difficult to separate them.
Prior studies on the effects of surface piercing
bodies have focused on the flow in the bow region of
ship-like bodies and on the effects of surface piercing
and near surface disturbances on the structure and
vorticity present in the layer just below the surface. In
the former area, studies by Stern et al. (1996), Zhang
and Stern ((1996), Roth et al. (1999), Pogozeliski et al.
(1996) and Jeong et al. (1996) have made considerable
progress toward describing the nature of the bow
vortex and wave-induced separation. The effects, in
the near surface layer, of penetrating bodies have been
investigated, for example, by Sheridan et al. (1998)
and Vlachos et al. (1998), while the effects of
surfactants on the wake structure at low Re have been
studied by Lang and Gharib (1998). In the present
study attention has been directed toward the
downstream near-wake region at high Re, where a
separated, two-phase flow is generated by the surfacepenetrating body. In this regime breaking waveforms
also occur. These turbulent wavefields result in a
visible white-water wake. To date, this regime has
received
little
attention
beyond
qualitative
observations. Our focus is on the WWW generation
region, leaving to subsequent studies descriptions of
the decay of the WWW.

ABSTRACT
Wakes of surface penetrating cylindrical and
streamlined shapes have been measured in a large, salt
water tow tank facility over a range of Froude and
Reynolds numbers of general interest. The white-water
wake (WWW) pattern was observed as a function of
object shape and speed using optical imagery, with the
camera following the model. Using these data the
evolution and decay of the WWW, which is the bubble
pattern generated by the entrained air, was observed in
detail, and related to the relevant non-dimensional
parameters, the Reynolds number (Re) that is
indicative of the degree of turbulence/air entrainment,
and the Froude number (F) that determines the nature
of the wake pattern. Two major source mechanisms
have been observed to contribute to the generation of
the WWW: centerline turbulence (resulting from
separation at the bow and stern as well as from the
boundary layer on the body) and breaking waves in the
body wake. The Re-F regimes where these
mechanisms dominate have been identified. In
addition to the experiments, surface wave patterns
were predicted using a nonlinear potential flow model.
An acceleration threshold criterion was used to predict
regions of likely wave breaking in the wake. In
addition, the concept of relating acceleration threshold
exceedance to breaking intensity was studied.
INTRODUCTION
Submarine operations in the littoral often
require mast penetrations for extended periods. Such
exposures can lead to detection by observation not
only of the mast itself but also the hydrodynamic
wake. In addition, wakes of obstacles in the coastal
ocean or rivers can provide sites for enhanced
biological activity, and the wakes of bridge and pier
supports can affect their structural integrity. While
substantial research has gone into the study of flow
around these structures, less effort has gone into
understanding the processes in the wake. Toward this

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

WHITE WATER EXPERIMENTS


Wakes of cylindrical and streamlined shapes
were measured in the large, salt water1 tow tank at the
Ohmsett National Oil Response Test Facility (cf.
Mullin and Lane 2000), which has dimensions of 203
m long by 20 m wide by 3.4 m deep. A fixture was set
up on the tow carriage (the main bridge) for
mounting the vertical cylindrical and streamlined
shapes. The maximum tow speed was 3.1 m/s (6 kt).
Figure 1 is a photograph of a faired strut towed down
the tank at 2.1 m/s (4 kt).
The WWW pattern was observed as a
function of object shape and speed using optical
imagery. Specifically, a monochrome CCD camera
was mounted on top of the bridge house to image
white-water patterns in the wake. The camera was
fitted with polarizing and red filters to minimize sun
reflection from the surface and bottom of the tank.

Figure 2: Experimental set-up at the Ohmsett facility.


Three mast shapes were tested, a cylinder
with a diameter of 0.168 m (designated as shape C0),
and two faired shapes with blunt leading and trailing
edges, C1 with a chord length, c = 0.305 m and a
thickness, t = 0.152 m, and a larger C2 with c = 0.416
and t = 0.222 m. Both faired shapes have t/c 0.5.
Each mast shape was 1.5 m long, with 1 m
submerged. The effect of submergence depth was not
studied here. At 3.4 m, the tank was sufficiently deep
that the surface waves of interest were effectively deep
water waves.
The speed range investigated, U = {1.5, 3.1}
m/s, yielded a Froude number (F = U/(g t)1/2) and
Reynolds number (Re = U t/) range, F = {1.02, 2.54},
Re = {2.28, 6.88}x105, that encompassed a wide range
of WWW phenomenology. These conditions are
shown in Table 1.

Figure 1: Faired strut (C2) towed at 2.1 m/s.

Table 1: Run conditions.

Camera calibrations were performed to


accurately determine the camera field of view for
subsequent image processing (remapping). Figure 2 is
a schematic of the camera, fixture and main bridge.
The camera position and depression angle were
measured so that images could be remapped to the
ground plane.

C0-1
C0-2
C0-3
C0-4
C1-1
C1-2
C1-3
C1-4
C2-1
C2-2
C2-3
C2-4

U (m/s)
1.5
2.1
2.6
3.1
1.5
2.1
2.6
3.1
1.5
2.1
2.6
3.1

F
1.17
1.64
2.03
2.41
1.23
1.72
2.13
2.54
1.02
1.42
1.76
2.10

Re (x105)
2.52
3.53
4.37
5.21
2.28
3.19
3.95
4.71
3.33
4.66
5.77
6.88

Continuous video images were obtained for


each run condition. An example of an individual raw
image is shown in Figure 3a. The duration of steadystate images, between carriage start-up and

A salt water facility was desirable for this test as there is evidence
that salinity can affect bubble size distribution in a white water wake
(cf. Scott 1975).

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

deceleration, varied with speed, and was at least 15 sec


for all runs, allowing sufficient time to generate
averaged WWW images, further discussed below.
DATA PROCESSING
Figure 3a shows an example single-frame
image from the video camera. To obtain images
remapped to the ground (water surface) plane (i.e. with
true dimensional scales) and to isolate the WWW, the
raw images were processed in the following steps:
1.

2.

3.

4.

5.

Start with an individual raw image (Figure


3a) or compute the temporal mean image
obtained by averaging over a known time
interval, nominally 15 sec.
Re-map the individual or mean images to the
ground plane (Figure 3b) and compute the
physical pixel dimensions, based on the
camera properties and the setup geometry
(Figure 2).
Compute the mean and standard deviation of
the background intensity ( I , ), selected
from a portion of the frame where whitewater does not appear.
Convert the mean, remapped image into
contrast units,
I ( I + n )
C=
I + ( I + n )
where I is the initial intensity of an individual
pixel and n is the number of standard
deviations.
Select a value for n and define the white
water wake as the image region with positive
contrast. For this study n = 3 was used.
Figure 3c shows the result.

Figure 3: Processing sequence for WWW images of


Mast C2, Run C2-2: a) raw image (single video
frame), b) remapped mean image, c) positive contrast
image.

The red and polarizing filters did not


completely eliminate reflections from the bottom of
the tow tank. Bottom reflections from the ambient and
Kelvin wake are visible in figures 3a and 3b. Obvious
reflections were masked out of the final positive
contrast images.
In Figure 3b, which shows the mean image
re-mapped to the water plane, and Figure 4, which
shows individual remapped images, the pixel size is 2
cm x 2 cm. Actual resolution is highest at the bottom
of the remapped images and decreases with increasing
distance from the mast. The remapped and masked
positive contrast images, as illustrated in Figure 3c,
define the WWW area for each run condition.

EXPERIMENTAL RESULTS
Individual images from each of the
experimental runs (see Table 1), remapped to the
water plane, are shown in figure 42. Clearly the WWW
area associated with the large fairing and the cylinder
are significantly larger than that of the smaller fairing.
This is due to the larger flow disturbance of the larger
fairing and the enhanced mixing in the recirculation
zone behind the cylinder, the diameter of which is
comparable to the thickness of a the smaller fairing.
2

The asymmetry of the cylinder wakes at the higher speeds is due


to a tilting of the mast that resulted from the induced oscillation
from vortex shedding coupled with the dynamics of the support
structure.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

water at high Re while the wave breaking mechanism


would dominate at low F. For example, the dominant
wave-breaking pattern apparent in Figure 4b recurs at
the lower speeds for the smaller shapes. Values of F
are similar for the three cases, 4b, 4e and 4i.
Similarly, trends are apparent by comparison of
images 4c, 4f and 4j. Since there were no runs at
intermediate speeds, the points of transition between
the various white-water patterns cannot be definitively
identified.
The angles of the breaking waves, measured
from the temporally-averaged images (not shown; see
Figure 3c for an example), are shown in Figure 5.
Above F > 1.4 the breaking-wave angle decreases with
increasing F (speed), while below this value it tends
toward a constant value of ~ 35. In the classical
Kelvin wave pattern, the angle of the diverging waves
as they approach the cusp line or envelope is 35 16.

C0
C1
C2

Figure 5: Breaking wave angle.


The white-water areas computed from the
temporally averaged images are shown in Figure 6.
The white-water area variations reflect the trends
apparent in Figure 4 as well as general expectations
based on size and bluntness as earlier discussed. The
clear anomaly is the large white-water area for mast
C2 at the 2.1 m/s speed (F = 1.4, see Figure 4b). The
breaking intensity reaches a local maximum at F~1.4
and then decreases with increasing speed and
decreasing wave angle.

Figure 4: Sequence of image stills at increasing


speeds for the three mast shapes. a)-d) C2, the large
faired shape, e)-h) C1, the smaller faired shape, i)-l)
C0, the cylindrical strut.
Also apparent in the image sequences of
Figure 4 are significant changes in the WWW patterns
as the speed changes. This is a result of the
dominance of one or the other of the two mechanisms
responsible for the WWW drag and wave breaking,
as discussed above. Qualitatively, it is expected that
the drag wake would be the dominant cause of white

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

computed white-water areas (from the averaged


images) for those runs where this mechanism
dominates (see Figure 7). Based on these, albeit
limited, data there seems to be a consistent trend with
the white-water area increasing as ~Re3. The higher
absolute level of the cylinder WWW area is expected
due to the enhanced mixing behind its non-streamlined
cross-section.

C0
C1
C2

C0
C1
C2

t/c = 1

Figure 6: Total white-water area for 15 sec averaged


images sequences.

t/c = 0.5

To identify regimes where white-water


generation is dominated by the drag-centerline wake
or the wave-breaking mechanism, a Re-F regime
diagram has been constructed, Figure 7. Each of the
test conditions is shown, and the conditions where
each of the mechanisms dominate have been identified
based on the individual images, Figure 4, and the
temporally-averaged images (not shown). It is clear
that the drag-centerline mechanism dominates in the
higher Re-F regime, while the breaking-wave process
dominates at lower Re-F values3.

Figure 8: White-water area in the drag-centerline


regime.
NUMERICAL MODEL
Simulation of spray and splash next to the
body and breaking waves in the wake is a challenging
hydrodynamics problem. Difficulties arise from the
fact that these physical processes cover a wide range
of scales. The approach taken here is semi-empirical.
Of the various processes that contribute to the total
white-water wake, some may be adequately
represented using relatively simple models while
others are more appropriately represented empirically
or by more sophisticated models. The objective of the
following analysis is to explore the feasibility of
employing an inviscid-flow model to predict regions
of wave breaking and white-water generation in the
surface-wave pattern.

C0
C1
C2

Centerline Dominated
White Water

Breaking Wave
Dominated White Water

Figure 7: Test regime showing regions where the


drag-centerline and the wave breaking mechanisms
dominate the white-water generation process.

The inviscid-flow problem is solved using a


boundary-integral technique.
In this framework,
breakup of the free surface and subsequent formation
of droplets and entrainment of air into the fluid
volume is not explicitly modeled. However, the
model is used here to predict regions in the wake
where breaking waves are likely to occur. The
following is an outline of the numerical model and
solution techniques.

In the drag-centerline regime it can be


expected that the amount of white-water would be
related to the intensity of the viscous mixing, i.e., the
Reynolds number. Shown in Figure 8 are the
3

The dominant mechanism is reasonably apparent in virtually all


images. The exception was the condition marked with the ?, where
it was not clear.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

The flow problem is described using a


velocity potential that is decomposed into a free
stream and perturbation part,
r

free surface domain and utilized a two-dimensional


hierarchical tree structure. The current algorithm
applies multipole acceleration to the entire problem
and, hence, uses a three-dimensional tree structure. In
addition, efficiency improvements reported in Cheng
et al. 1999 have been incorporated. Computation
timing experiments indicate the algorithm improves
computation time by an order of magnitude. Figure 9
shows single time step solution times for a 2.8 GHz
Intel Xeon processor. The multipole algorithm
reduced the solution time from O(N2.2) for the direct
calculation down to O(N1.2); an order of magnitude
reduction.
In addition, the multipole algorithm
eliminates the requirement to store an N2 matrix in
memory, an advantage for large problems. Due to the
overhead associated with setting up the multipole data
structure, the algorithm does not provide improved
solution times until the number of nodes exceeds
~5000 - 6000.

( x , t ) = U (t ) x + ( x , t ) .
The boundary value problem
perturbation potential is as follows,
2 = 0

+
=0
t
z

the

r
x V

P
1
2
+ + g + a = 0
t 2

=0
n

for

r
xS f

r
xS f

r
x Sh
r
x

with the following nomenclature,


V is the fluid volume
S f is the free surface
S h is the body surface,
r

( x , t ) is the perturbation potential


( x, y, t ) is the free surface elevation
g is the acceleration of gravity
r
n is the unit normal pointing out of the fluid

The
far-field
boundary
condition,
r
,
is
not
strictly
satisfied.
This
0, x
boundary condition is left open in the solution method
described below.
Consequences of leaving the
boundary condition unspecified may include wave
reflections off the boundary and mass flux through the
boundary. These effects appear to be minimal for the
steady, forward-speed problem studied here.
The equations are solved using an EulerLagrange time stepping technique. The free surface
boundary conditions are integrated in time using 4th
order Runge-Kutta time stepping. A mixed boundary
value problem is solved at each time step using a
multipole-accelerated desingularized method (c.f. Cao
1991, Beck et al. 1994, Scorpio et al. 1996 and
Scorpio 1997). Multipole acceleration was originally
developed by Greengard (1987) for solving electropotential problems. The multipole algorithm used
here differs somewhat from the one described in
Scorpio et al. 1996 and Scorpio 1997. In the previous
work, multipole acceleration was applied solely to the

Figure 9: Multipole algorithm efficiency.


FREE SURFACE STABILITY
For flow around a vertical strut at Froude
numbers of interest, the near field free surface is
unstable, resulting in spray, splash, breaking waves
and air entrainment. Unstable regions of the near field
wake often appear as white water when the free
surface is sufficiently energetic to entrain air or eject
water droplets. Direct numerical simulation of this
multiphase flow is challenging due to the range of
scales requiring resolution. The goal here is to study
the applicability of using a simplified physics model,
nonlinear potential flow, to predict regions in the nearwake where breaking waves and white water are likely
to exist.
A requirement for this approach is that the
potential flow solution remains stable yet provides
sufficient information to predict regions of likely

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

instability in the wake. In developing numerical


methods to satisfy this requirement, three sources of
numerical instability were addressed: 1) Spray jets at
the body / free surface intersection line, 2) Breaking
waves and 3) Discretization errors.
Numerical
techniques were implemented to address each source.
At high speeds, water run up at the leading
edge turns into a thin liquid sheet and breaks up into a
spray of water droplets. The result of run up and
splash down can be seen near the leading edge of the
strut shown in Figure 1. The nonlinear potential flow
model can simulate steep leading edge water run up.
However, numerical instability results when the liquid
sheet becomes exceedingly steep and thin. Stability is
recovered by applying a slope cut-off along the body /
free surface intersection line. When the slope exceeds
a threshold value, the spray sheet is truncated and the
free surface is locally regridded. No special attempt
has been made at this time to account for fluid
removed from the problem due to free surface
truncation at the intersection line.
The nonlinear potential flow model can
predict steep overturning waves (cf. Longuet-Higgins
and Cokelet 1976 and Xue et al. 2001). However, a
re-entrant breaking wave crest will result in numerical
instability. Free surface stability can be maintained
near steep wave crests by enforcing a slope or
acceleration threshold.
When the threshold is
exceeded the wave elevation and velocity potential are
reduced at the wave crest and the free surface is
locally smoothed. The level of wave elevation and
velocity potential reduction is proportional to the
threshold exceedance. For the calculations reported
herein, the waves did not exceed the numerical
stability threshold and no crest truncation was
required.
Point to point oscillations are known to arise
in nonlinear Euler-Lagrange time stepping techniques.
In addition, some high wave number energy is
introduced into the solution due to the spray truncation
algorithm. A 5-point smoothing function applied in
the cross-track, or y-direction, effectively reduces the
high wave number energy.
The total energy removed from the solution
due to the above free surface stabilization techniques
is computed and recorded. However, results from the
energy analysis are not reported here.
Quasi-steady state solutions are produced by
starting from rest and gradually accelerating the body
up to steady forward speed. Due to high wavenumber
content generated by the free surface stabilization
algorithm and unsteady effects introduced during the
acceleration phase, the solution time history is
unsteady for some time after reaching steady forward
speed. An equivalent steady state is produced by

computing the temporal mean over the constant speed


portion of the solution time history.
WAVE BREAKING CRITERIA
The potential flow code employed above does
not explicitly predict wave breaking. Instead, the
breaking regions were identified using an acceleration
threshold (c.f. Longuet-Higgins, 1985 and Snyder et
al. 1983). All wave crests with downward Lagrangian
accelerations exceeding a fixed threshold, at, were
assumed to become unstable and break. For a Stokes
wave, the theoretical limit on downward acceleration
is 0.39 g. The acceleration threshold for actual waves
is expected to be lower due to nonlinear interactions.

Figure 10: Results from the potential flow model for


geometry C2 at 2.1 m/s. Panel a) Free surface
elevation color image. Panel b) Surface slope
magnitude color image. Panel c) Vertical acceleration
color image in units of g. Panel d) Isometric view of
the wake where regions of threshold exceedance are
colored white.
Figure 10 shows sample potential flow
predictions of the surface elevations, slopes, and
vertical accelerations for the C2 geometry at 2.1 m/s.
Surface regions where the downward acceleration
exceeds 0.15 g are also depicted.
Figure 11 shows calibrated, remapped camera
images for the C2 geometry over the range of
measured speeds. Figure 12 shows the same sequence
of images overlaid with the breaking regions identified
by applying an acceleration threshold of 0.15 g.
Several observations can be made from this
comparision:

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

1.5 m/s

2.6 m/s

Area of Threshold Exceedance (m

The regions where the downward surface


acceleration exceeds the 0.15 g threshold correlate
reasonably well with the observed breaking regions.
The surface area where the acceleration threshold is
exceeded follows the same trend as the white water
area determined from the remapped camera images,
as shown in Figure 13.
For the 1.5 m/s speed, the breaking regions
identified using the acceleration threshold include
the wave crests immediately behind the mast and the
next set of wave crests following downstream.
However, in the corresponding camera images, the
second set of wave crests are not breaking. This
discrepancy results from the lack of breaking wave
dissipation in the potential flow predictions.
Dissipation due to a breaking wave will significantly
reduce the amplitude of the following waves (c.f.
Duncan 1983) making them less likely to break.

2.6 m/s

0.6
0.5
0.4
0.3
0.2
0.1
0.0
1.0

1.2

1.4

1.6

1.8

2.0

2.2

Froude Number

Figure 13: Predicted area of threshold exceedance


based on thickness for the C2 geometry.
BREAKING CHARACTERIZATION
Once a wave breaks, it is valuable to
characterize the resulting surface. For remote sensing
applications, surface disturbances generated by the
turbulent plume formed on the forward face of the
breaking crest may be important.

2.1 m/s

Breaking-generated surface roughness was


investigated in the laboratory experiment reported by
Walker et al., 1996 and Ericson et al., 1999. In that
work, steady breaking waves were created by
submerging a hydrofoil in a re-circulating water
channel and surface profiles of the resulting breaking
crests were captured using a laser sheet imaging
system. Spectral analysis of the surface profiles
suggested that the elevation spectrum of the breakinggenerated surface roughness obeys a saturated powerlaw form given by

3.1 m/s

Figure 11: Calibrated, remapped white water patterns


for the large faired shape.
1.5 m/s

0.7

S ( k ) = 0.004k

2.1 m/s

3.5

3.5 k o 4
exp

4 k

where S is the spectral density in m4 and k is the


wavenumber in rad/m. The low wavenumber rolloff ko
was found to agree reasonably well with 2/Lplume
where Lplume is the turbulent plume length (i.e., the
distance from the breaking crest to the toe of the
breaking plume).

3.1 m/s

For the 71 cm long breaking waves


considered by Walker et al. and Ericson et al., the
plume length was observed to increase with breaking
intensity (i.e., dissipation rate). For now, the plume
length will also be assumed to scale linearly with the
breaking wavelength break, so that

Figure 12: Regions of threshold exceedance predicted


by the model.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

L plume = c break
where the scaling coefficient depends on the breaking
intensity. From the observations of Walker et al. and
Ericson et al., the coefficient has a lower limit of about
0.1 for weak (incipient) breaking waves. An upper
limit of 0.5 restricts the plume from extending beyond
the trough ahead of the breaking crest.
Clearly, the potential flow predictions cannot
be used to explicitly determine the dissipation rates
because breaking is not included in the solution.
However, it may be possible to devise a metric that
relates the potential flow solutions to the breaking
dissipation rate. For example, breaking regions where
the downward acceleration is very close to the
acceleration threshold may lose relatively little energy
due to breaking while those where the downward
acceleration greatly exceeds the threshold may have
much higher dissipation rates. This metric gives

Lplume

break

Figure 15: Analysis of diverging wave crests.


The turbulent plume lengths were estimated
from the potential flow predictions using the above
model as follows. First, the local wave direction was
determined by finding the vector normal to each point
along the wave crests predicted by the potential flow
code. This is demonstrated in Figure 15, which shows
an x-y map of the free surface wave field. Blue (red)
color corresponds to negative (positive) surface
elevations. The black curve is drawn along the
locations of maximum elevation in the y-direction as a
function of x. The white areas indicate where the
downward acceleration exceeds the threshold.

= C (a L at )

where aL is the downward acceleration at the crest of


the breaking wave and at is the acceleration threshold.

Then, the local breaking wavelength


corresponding to each crest point was set to twice the
distance from the crest to the trough ahead of the crest
along the wave propagation direction. This is
demonstrated in Figure 16, which shows a plot of the
interpolated surface elevations perpendicular to the
wave crest. The wave crest is at distance = 0 and the
wave is propagating from right to left. The breaking
wavelength, is defined as twice the distance from
the wave crest to the forward trough as indicated in the
figure. The red curve is included for reference and
indicates the equivalent linear wave. The linear
wavelength is,

turbulent plume length, Lplume

Figure 14: Determination of turbulent plume length


from the mean wake image. The image shown is the
mean wake for the C2 geometry at 2.1 m/s.

linear = 2U 2 / g sin

The mast wake images provide a convenient


opportunity to evaluate this approach. The turbulent
plume lengths of the breaking crests generated by the
masts were determined from the remapped camera
images, as demonstrated in Figure 14.

where U is the speed and is the angle of the wave


crest.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Figure 17: Turbulent plume length predicted by the


model plotted against along crest coordinate. a, b, c
= 1.5, 2.1, 2.6 m/s.

Zmax

Turbulent plume length predictions are overlaid on


mean wake images in Figure 18 for the C2 geometry at
three speeds, a, b, c = 1.5, 2.1, 2.6 m/s. The blue
circles indicate the location of the diverging wave
crest as predicted by the model. The location and
angle of the breaking wave crests appear to be
reasonably well predicted by the model. The length of
the blue lines denotes predicted turbulent plume length
while the angle of the lines are perpendicular to the
local wave crest.
An alternative to using acceleration threshold
exceedance as an indicator of breaking intensity is to
model the dissipation effects of the breaking wave.
Cointe and Tulin (1994) proposed a theory where the
breaker is modeled as a stagnant eddy riding on the
forward face of the breaking wave. Muscari and Di
Mascio (2003) implemented this theory in the context
of a 2D RANSE (Reynolds Averaged Navier-Stokes
Equation) model. Their results compared well with
measurements reported in Duncan (1983). One of the
parameters of the model is the breaker height, that is
related to the turbulent plume length described here.
While not included here, it would be
interesting to study whether Coint and Tulins theory
could be applied to the current 3D problem and used to
predict the turbulent plume length. This approach
would provide an estimate of the dissipation due to
breaking that is lacking in the current implementation.

Zmin

/2

Figure 16: Diverging wave profile.

0.4 * * (a L -a t)

1.0
0.8
0.6
0.4
0.2
0.0
-0.2
-0.4
-0.6
-0.8
-1.0

0.4 * * (a L -a t)

At the time of this writing, extraction of


turbulent plume lengths from the images was
incomplete. However, the constants C and at have
been initially estimated at 0.4 and 0.11 g respectively.
Figure 17 shows plots of predicted turbulent plume
length versus along-crest coordinate given constants C
and at.

1.0
0.8
0.6
0.4
0.2
0.0
-0.2
-0.4
-0.6
-0.8
-1.0

0.0

1.0

1.5

2.0

2.5

Along - Crest Coordinate (m)

0.0

0.4 * * (aL-at)

0.5

0.5

1.0

1.5

2.0

2.5

Along - Crest Coordinate (m)

1.0
0.8
0.6
0.4
0.2
0.0
-0.2
-0.4
-0.6
-0.8
-1.0

0.0

0.5

1.0

1.5

2.0

2.5

Along - Crest Coordinate (m)

10

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

breaking induced white water. A WWW area local


maximum was noted at F~1.4 for the C2 geometry.
Results from a potential flow model were
analyzed to explore the feasibility of predicting
regions of wave breaking and white water generation
in the surface wave pattern. Application of an
empirically determined acceleration threshold
produced predicted regions of wave breaking that
agreed well with the breaking regions seen on the
images. In addition, the level of acceleration threshold
exceedance was used to estimate the wave breaking
intensity. It was hypothesized that the level of
threshold exceedance is related to the turbulent plume
length or the extent of the turbulent region on the
forward face of the breaking wave. Predicted plume
lengths agreed reasonably well with the data.

ACKNOWLEGEMENTS
This work was sponsored in part by
NAVSEA PMS 435, Mr. Swarn Dulai under contract
N00024-03-D-6606 and by internal research and
development funds at the Johns Hopkins Applied
Physics Laboratory. The help of the engineering staff
at the Ohmsett facility in running the tow tank is
gratefully acknowledged. We would like to thank Dr.
T. Gieseke, Mr. B. Jantzen, Mr. C. Henoch, Mr. J.
Skutnik and Mr. E. Rabe from the Naval Undersea
Warfare Center for their collaboration in planning and
executing the experiments at Ohmsett. Dr. D. L.
Porter, Mrs. S. Daniels and Mr. C. Buckingham from
the Johns Hopkins Applied Physics Laboratory are
acknowledged for there assistance in processing the
images.

REFERENCES
Figure 18: Mean wake images for the C2 geometry.
a, b, c = 1.5, 2.1, 2.6 m/s. Blue circles denote location
of the wave crest predicted by the model. Blue lines
denote plume length predicted by the model.

1.

CONCLUSIONS

2.

In general, significant variations in the total


WWW area and pattern were observed as a function of
mast shape and speed. The WWW area followed a
power law relationship with Reynolds number (Re3)
for the high speed cases where near field effects such
as turbulent air entrainment and spray and splash next
to the body were the primary sources of white water.
White water generated from wave breaking was the
dominate source for the slower speeds. A critical
Froude number has been identified (F~1.4) above
which the breaking wave angle abruptly begins to
decrease along with the breaking intensity and

3.

4.

Beck, R.F., Y. Cao, S.M. Scorpio and W.W.


Schultz, Nonlinear Ship Motion Computations
Using the Desingularized Method, in 20th
Symposium on Naval Hydrodynamics, Santa
Barbara, California, 1994.
Cao, Y., Computations of Nonlinear Gravity
Waves by a Desingularized Boundary Integral
Method, Ph.D. Thesis, Department of Naval
Architecture and Marine Engineering, The
University of Michigan, Ann Arbor, Michigan,
1991.
Cointe, R. and M.P. Tulin, A Theory of Steady
Breakers, Journal of Fluid Mechanics, Vol. 276,
1994, pp. 1-20.
Cheng, H., L. Greengard and V. Rokhlin, A Fast
Adaptive Multipole Algorithm in Three
Dimensions, Journal of Computational Physics,
Vol. 155, 1999, pp. 468-498.

11

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

5.

6.

7.

8.

9.

10.

11.

12.

13.

14.

15.

16.

17.

Duncan, J.H., The Breaking and Non-Breaking


Wave Resistance of a Two-Dimensional
Hydrofoil, Journal of Fluid Mechanics, Vol. 126,
1983, pp. 507-520.
Ericson, E.A., D.R. Lyzenga, and D.T. Walker,
Radar backscatter from stationary breaking
waves, J. Geophys. Res., 104, 29,679-29,695,
1999.
Greengard, L., The Rapid Evaluation of Potential
Fields in Particle Systems, The MIT Press,
Cambridge, Massachusetts, 1987.
Jeong, U.-C., Y. Doi and K.-H. Mori, Numerical
Investigation on the Turbulent and Vortical Flows
Beneath the Free Surface Around Struts,
Twenty-First
Symposium
on
Naval
Hydrodynamics, 1996, pp. 794-809.
Lang, A. W. and M. Gharib, On the Effects of
Surface Contamination in the Wake of a SurfacePiercing Cylinder, ,1998 Conf. on Bluff Body
Wakes and Vortex-Induced Vibration, ASME
Fluids Eng, Div. Annual Summer Meeting,
Washington D.C., Bearman, P. W. and C. H. K.
Williamson eds., 1998, Paper No. 28, 6 pp.
Longuet-Higgins, M.S. and E.D. Cokelet, The
Deformation of Steep Surface Waves on Water: I.
a Numerical Method of Computation, In
Proceedings of the Royal Society of London, Vol.
A350, 1976, pp. 1-26.
Longuet-Higgins, M.S., Accelerations in Steep
Gravity
Waves,
Journal
of
Physical
Oceanography, Vol. 15, 1985, pp. 1570-1579.
Mullin, J.V. and J.S. Lane, R&D Users Guide to
the Ohmsett Oil Spill Response Test Facility,
Spill Science & Technology Bulletin, Vol. 6, No.
1, 2000, pp. 77-87.
Muscari, R. and A. Di Mascio, A Model for the
Simulation of Steady Spilling Breaking Waves,
Journal of Ship Research, Vol. 47, No. 1, March
2003, pp. 13-23.
Pogozelski, E., J. Katz and T. Huang, The
Shoulder Wave and Separation Generated by a
Surface-Piercing Strut, Twenty-First Symposium
on Naval Hydrodynamics, 1996, pp. 346-358.
Roth, G. I., D. T. Mascenik and J. Katz,
Measurements of the flow structure and
turbulence within a ship bow wave, Phys. Fluids,
Vol. 11, 1999, pp. 3512-23.
Scorpio, S.M., R.F. Beck and F.T. Korsmeyer,
Nonlinear Water Wave Computations Using a
Multipole Accelerated, Desingularized Method,
in Proceedings of the 21st Symposium on Naval
Hydrodynamics, Trondheim, Norway, 1996.
Scorpio, S.M., Fully Nonlinear Ship Wave
Computations Using a Multipole Accelerated,
Desingularized Method, Ph.D. Thesis, Department of Naval Architecture and Marine

18.

19.

20.

21.

22.

23.

24.

25.

Engineering, The University of Michigan, Ann


Arbor, Michigan, 1997.
Scott, J.C., The Role of Salt in Whitecap
Persistence, Deep Sea Research, Vol. 22, 1975,
pp. 653-657.
Sheridan, J., J.-C. Lin and D. Rockwell, The
Interaction of a Cylinder Wake and a FreeSurface,1998 Conf. on Bluff Body Wakes and
Vortex-Induced Vibration, ASME Fluids Eng,
Div. Annual Summer Meeting, Washington D.C.,
Bearman, P. W. and C. H. K. Williamson eds.,
1998, Paper No. 26, 6 pp.
Snyder, R.L., and R.M. Kennedy, On the
formation of whitecaps by a threshold mechanism.
Part I: Basic formalism, J. Phys. Ocean., 13,
1482-1492, 1983.
Stern, F., J. Longo, Z. J. Zhang and A. K.
Subramani, Detailed Bow-Flow Data and CFD
for a Series 60 CB =0.6 Ship Model for Froude
Number 0.316, J. Ship Res., Vol. 40, 1996, pp.
193-199.
Vlachos, P., M. J. Donnelly and D. P. Telionis,
On The Wake of a Circular Cylinder Piercing a
Water Free Surface, ,1998 Conf. on Bluff Body
Wakes and Vortex-Induced Vibration, ASME
Fluids Eng, Div. Annual Summer Meeting,
Washington D.C., Bearman, P. W. and C. H. K.
Williamson eds., 1998, Paper No. 27, 6 pp.
Walker, D.T., D.R. Lyzenga, E.A. Ericson, and
D.E. Lund, Radar backscatter and surface
roughness measurements for stationary breaking
waves, Proc. R. Soc. Lond. A, 452, 1953-1984,
1996.
Xue, M., H. Xu, Y. Liu and D.K.P. Yue,
Computations of Fully Nonlinear Three
Dimensional Wave Wave and Wave Body
Interactions. Part I. Dynamics of Steep Three
Dimensional Waves, Journal of Fluid Mechanics,
Vol. 438, 2001, pp. 11-39.
Zhang, Z. J. and F. Stern, Free-Surface WaveInduced Separation, Trans. ASME, J. Fluids
Eng., Vol. 118, 1996, pp. 546-554.

12

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION

AUTHORS REPLY

Dane M. Hendrix
Naval Surface Warfare Center, Carderock Division,
USA

An alternative approach would be to use the


perturbation velocity vector to define the propagation
direction. The breaking wavelength can then be
estimated using the linear dispersion relation such

I would like to thank the authors for a very


interesting paper addressing a difficult area of freesurface hydrodynamics. Breaking waves are a
characteristic of many flows where steady waves are
important; however, our ability to model them has
been limited by their complexity.
I have two questions to ask the authors. The
first concerns the separation of white watergeneration regimes into centerline dominated and
breaking wave dominated. In Figure 7, the Reynolds
number/Froude number space is divided by a dashed
line indicating a separation into these two regimes. It
appears to me that from this limited data set
presented in this paper, the separation could just as
well depend only on velocity. Do the authors have
any comment on why they believe that Reynolds
number and Froude number are the correct
dimensionless ratios to characterize this flow or how
they would go about verifying this dependency?

that linear = 2U 2 / g sin as shown by the red


curve in figure 16.

AUTHORS REPLY
Dimensional analysis leads to the conclusion
that the Reynolds and Froude numbers are the
governing parameters.
Verification of the
dependencies of the wake characteristics on these
parameters would entail a larger data set than
currently available, i.e. one where each of the
parameters could be varied independent of the other.
DISCUSSION
Dane M. Hendrix
Naval Surface Warfare Center, Carderock Division,
USA
The second question concerns the
determination of local breaking wavelength. I am
concerned that the method of determining a crest line,
propagation direction, and corresponding breaking
wavelength described in the paper might not be
robust enough for general application. For example,
the first crest at the leading edge of the strut shown in
Figure 15 exceeds the breaking criteria, but does not
have a smooth crest line from which to determine a
propagation direction. Would the authors like to
comment on any ideas for handling such situations?

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

St. John's, Newfoundland and Labrador, CANADA, 8-13 August 2004

A pseudo-spectral method for non-linear wave


hydrodynamics
Wooyoung Choi and Christopher Kent
(University of Michigan, USA)
ABSTRACT

optimization.
Though the potential flow assumption is often violated
in many flow situations, the computation of potential flows
still serves as a powerful tool for practical applications to
predict hydrodynamic wave forces on ships and offshore
structures. These outer potential flow solutions also play
an important role of providing the boundary condition for
inner viscous problems around the body.
The mixed Eulerian-Lagrangian method based on a
boundary integral formulation, originally developed by
Longuet-Higgins & Cokelet (1976) for free surface waves,
has been the most widely-used computational method for
inviscid free surface problems. See, for example, the review by Beck & Reed (2001) for its application to wavebody interaction problems. A major difficulty of this
method lies in the free surface representation using a distribution of a large number of singularities whose location
and strength must be computed at every time step. Erroneous results are sometimes produced by numerical errors introduced when approximating the singular integrals
and redistributing singularities on the free surface, and care
must be taken when carrying out these steps.
An alternative numerical approach to solve inviscid free
surface problems has been proposed by Fenton & Rienecker (1982) using the Fourier-series expansion. For free
wave problems, Dommermuth & Yue (1987) have further
improved the method by expanding the nonlinear free surface boundary conditions about the mean free surface and
by solving the resulting boundary value problems for each
order using a pseudo-spectral method based on Fast Fourier
Transform (FFT). Liu et al. (1992) then modified this
approach for wave-body interaction problems. Although
these methods have been adopted for practical applications
(Lin & Kuang 2004), it is still a nontrivial task to study
unsteady wave-body interactions since they are computationally expensive and numerical implementation is rather
complex.

We present a new hybrid asymptotic-numerical method


to study nonlinear wave-body interaction in threedimensional water of arbitrary depth. After solving a simplified body problem numerically using a distribution of
singularities along the body surface, we reduce the nonlinear free surface problem to a closed system of two nonlinear evolution equations, using a systematic asymptotic
expansion, for the free surface elevation and the velocity potential at the free surface. The system, correct up
to third order in wave steepness, is then solved using a
pseudo-spectral method based on Fast Fourier Transform.
We study the evolution of unstable Stokes waves and the
generation of nonlinear surface waves by translating twoand three-dimensional dipoles. In order to validate our numerical method, a translating circular cylinder is also considered and our numerical solution is compared with the
fully nonlinear numerical solution.
1 INTRODUCTION
The nonlinear wave-body interaction problem is one of the
most exciting and challenging problems in ship hydrodynamics. With the advent of faster computers, there has
been remarkable progress made over the last decade in
computational hydrodynamics. Recently various numerical
methods including Smooth Particle Hydrodynamics (Colagrossi et al. 2000) and the Level Set Method (Iafrati et
al. 2001) as well as finite difference/volume methods have
been widely used but, due to the uncertainty of turbulence
modeling for free surface flows and the high computational
cost, it is still problematic to solve the fully nonlinear,
three-dimensional, time-dependent viscous hydrodynamic
equations in the presence of a body near or on the free surface. These numerical tools are therefore rarely used in
industry for preliminary ship/offshore structure design and
1

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

At the free surface, the velocity potential and the surface


elevation satisfy the kinematic and dynamic free surface
boundary conditions:

Here we present a relatively simple third-order nonlinear


formulation which can be solved quickly and accurately by
using a surface singularity distribution method for the body
problem and a pseudo-spectral method for the free surface
problem.
We first decompose the total velocity potential into the
velocity potential for the body problem and that for the free
surface problem. The body velocity potential satisfying the
body boundary condition and the zero boundary condition
at the mean free surface can be found, for example, in terms
of a distribution of singularities. Using an asymptotic expansion similar to that used by Choi (1995) for free waves,
the free surface problem is reduced to two nonlinear evolution equations for two physical variables defined on the free
surface: the free surface elevation and the velocity potential. The resulting evolution equations are then solved by
using a pseudo-spectral method based on the FFT. Since
we only distribute singularities along the body surface and
we solve a closed set of the evolution equations in the horizontal plane, our numerical method is substantially faster
than other fully nonlinear numerical methods.
This paper is organized in the following order. After introducing the governing equations in 2, the detailed formulation is described in 3 and 4. With the numerical
method described in 5, some numerical solutions for infinitely deep water are presented in 6.

t + u = w
p=0

at

at z = (x, t).

t + 12 ()2 + 12 z 2 + gz + p/ = 0,

h z (x, t),




+ = 1 + ||2 W ,
t
1
+ 2 ()2 + g =
t

(1)

at z = h.


1 + ||2 W 2 ,

(x, t) = (x, , t),

(11)

(12)

and W is the vertical velocity evaluated at the free surface


defined as


W =
.
(13)
z 
z=

If we can find the expression for W in terms of and ,


equations (10)(11) will be a closed system for and . In
order to write W in terms of and , we first expand
and W in Taylor series

(3)

where V is the body velocity, n is the normal vector directed into the body, and SB (t) represents the instantaneous body position. The bottom boundary condition to
be imposed at z = h is given by

=0
z

1
2

(10)

where (x, t) is defined as

The body boundary condition can be written as


on SB (t),

(7)

the free surface boundary conditions can be written in terms


of the surface elevation and the free surface velocity potential as

where (x, t) is the free surface elevation, x = (x, y), and


the horizontal gradient is defined by



,
.
(2)
=
x y

= Vn
n

(6)

where g is the gravitational acceleration and is the fluid


density.
By substituting z = into (5) and (6), and using the
following chain rule for differentiation:

 




,
(8)
=

t z=
t
z z= t







,
(9)
=


xj z=
xj
z z= xj

For an ideal fluid we can introduce the velocity potential


(x, z, t) satisfying the Laplace equation:
for

(5)

The pressure p can be found by using the Bernoulli equation:

2 GOVERNING EQUATIONS

2
2 + 2 = 0
z

z = (x, t),

(x, t) = 0 +

(1)n

n=0


n=1

(4)

2n+1
n w0
(2n + 1)!
(1)n

2n n
 0 ,
(2n)!
(14)

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

W (x, t) = w0 +

(1)n

n=1

where k 2 = k1 2 + k2 2 . Then the Fourier Transform of the


total vertical velocity at the mean position w0 (x, t) can be
written as


F 
B 
w0 (k, t) =
+


z 
z 

2n1
n 0
(2n 1)!
(1)n

n=1

2n n
 w0 ,
(2n)!
(15)

z=0

= k tanh(kh) 0 +

where we have used 2 /z 2 = ,  = 2 and


0 = (x, 0, t),

w0 = w(x, 0, t) .

(23)

where f B represents the vertical velocity at the mean free


surface induced by the body:

B 
B
,
(24)
f (x, t) =
z z=0

(16)

With the relationship between 0 and w0 to be found under


the assumption of small wave slope, we can write W , from
(14)(15), in terms of and successively up to any order
of wave slope but in this paper we will consider the thirdorder expression for W .

which can be computed after solving the body velocity potential described below. After taking the inverse Fourier
Transform, we have the relationship between w0 and 0 as

3 FORMULATION

w0 (x, t) = L[0 ] + f B (x, t).

In this section we will describe how to find the relationship


between 0 and w0 defined in (16). We first decompose the
velocity potential into
= F + B ,

z=0

fB,

(25)

Since we are using a pseudo-spectral numerical method,


the linear integral operator L (Choi 1995) acting on a
Fourier component is of interest and is defined as


L eikx = k tanh(kh) eikx .
(26)

(17)

where F and B are the solutions of the free surface and


body problems, respectively, as shown in figure 1.

3.2 BODY VELOCITY POTENTIAL


3.1 FREE SURFACE VELOCITY POTENTIAL

The body effect on the free surface denoted by f B (x, t) can


be found by solving the following boundary value problem:


2
2
for h z 0, (27)
+ 2 B = 0
z

The relationship between two physical variables at the


mean free surface, 0 and w0 , can be found by formally
solving the following linear Dirichlet-Neumann boundary
value problem for F :


2
2 + 2 F = 0 for h < z < 0,
(18)
z
F

= 0 (x, t)

at

z = 0,

F
=0
at z = h.
z
By using the Fourier transform defined by

F (k, z, t)
F (x, z, t) eikx dx,

B = 0
B
=0
z

(19)

at z = h,

(29)

on SB (t),

(30)

where F is given by (22) for known 0 (x, t) (see also


equation (46)). Compared with classical linear formulations for wave-body interaction problems, the boundary
condition at the free surface is very simple and there is no
need to introduce complicated free surface Greens functions satisfying the physical linear free surface boundary
condition, in particular, for infinitely deep water. Notice
that the body boundary condition in (30) is imposed at the
instantaneous body position. The solution of (27)(30) can

(21)

where k = (k1 , k2 ), the solution of (18)(20) in the transformed plane can be found as
cosh[k(z + h)]
,
cosh(kh)

(28)

B
F
= VB n
n
n

(20)

F = 0 (k, t)

at z = 0,

(22)

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

(x,t), (x,t)

SF
SB

(a)

h
= F+ B

= 0 (x,t)

=0

n = V.n n

(b)

(c)

Fn = 0

n= 0

Figure 1: Decomposition of the original problem into the free surface and body problems: (a) original problem, (b) free
surface problem, (c) body problem.
where the first term represents the velocity potential for a
source located at (x , z  ) for z  < 0 and the second term
is for the image potential for the zero free surface boundary condition. For finite water depth, Greens function is
more complicated due to the additional bottom boundary
condition and is given (Wehausen & Laitone 1960, 13) by

be found by various methods including a distribution of singularities, the multipole expansion method, Greens identity, etc. In this paper, B is represented by a distribution
of singularities as

B (x, z, t) =
(x , t) G(x, x , z, z  ) dS  ,
(31)
SB

G=

where is determined by imposing the body boundary condition (30) and Greens function G(x , x) satisfies


2
2
+ 2 G = (x x , z z  )
z
for h z 0,
(32)

|x x |2 + (z z  )2

1
4 

1
|x x |2 + (z z  )2

1
4 

|x

x |2

+ (z +

z  )2

1/2 ,

1
4 

1/2

 kh
2e
cosh[k(z  + h)] cosh[k(z + h)]
+
cosh(kh)
0
J0 (k|x x |) dk, (36)

where represents a principal value integral and J0 is the


zeroth-order Bessel function of the first kind. In our numerical computations, we need to use doubly periodic Greens
functions, as described in Xue et al. (2001).
After solving the body problem for B and evaluating
B
B
f (x, t) =
, we have the relationship between 0
z 
z=0
and w0 from (25) and then, with using (14)(15), we can
close the system of evolution equations given by (10)(11).

1/2 ,

1/2 ,

|x x |2 + (z + z  + 2h)2

G=0
at z = 0,
(33)
G
=0
at z = h.
(34)
z
The solution of (32)(34) can be easily found, for infinitely
deep water, as
G=

1
4 

(35)

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Then, from (10)(11), the evolution equations correct to


third order in wave steepness can be written as

4 EVOLUTION EQUATIONS
4.1 LINEAR APPROXIMATION

+ + () + L []
t
 



+2 12 2 + L L + 12 2 2 = 0, (43)

For the leading-order approximation for / =   1 and


/(c) = O(), from (14), (15) and (25), W can be found
as
W  w0 = L[0 ] + f B  L[] + f B ,

+ g + 12 12 2
t


2 + L[] = 0. (44)

(37)

and, after dropping nonlinear terms, (10)(11) can be reduced to

+ L[] = f B ,
t

+ g = 0,
t

Notice that, using the formal expression for the vertical velocity in terms of and , the system of equations (43)
(44) is closed. This is the unique feature of our formulation
different from other formulations for nonlinear free surface
problems. In the absence of the body, the system can be
reduced to that derived by Choi (1995).

(38)

which can be combined into a single equation for :


2
g L[] = f B .
t2

(39)

For free waves (f B = 0), by substituting into (38)


= a ei(kxt) ,

5 NUMERICAL METHODS

= b ei(kxt) ,

To solve the evolution equations (43)(44), we adopt a


pseudo-spectral method. First we approximate the free surface elevation and the free surface velocity potential
using a truncated Fourier series:

and, using (26), we have


i a k tanh(kh) b = 0,

i b + g a = 0 ,

which gives the dispersion relation for surface gravity


waves in water of finite depth:
2 = gk tanh(kh),

(x, t) =

(40)

Nx

2


n= N2x

N
m= 2y

Nx

where k = |k|.

(x, t) =

Ny

2


bnm (t) eink1 x+imk2 y ,

where Nx and Ny are the numbers of Fourier modes in


the x- and y-directions, respectively, and k1 = 2/L1 and
k2 = 2/L2 with L1 and L2 being the computational domain lengths in the x- and y-directions, respectively. The
Fourier coefficients anm (t) and bnm (t) can be computed
by the double Fast Fourier Transform (FFT). All the linear operations are evaluated in the Fourier space, while the
product between two functions for the nonlinear terms in
the evolution equations are computed in the physical space.
For example, the two linear operators ( and L[ ]) are evaluated in the Fourier space as


anm (t) eiKx = i K anm (t) eiKx ,


L anm (t) eiKx =

For the third-order approximation, equations (14) and (15)


can be approximated by
 0 + w0 12 2 2 0 ,
W  w0 2 0 12 2 2 w0 ,
whose solutions can be found, by succession, as
(41)

W = L[] 2 12 2 ( 2 )


L L[] + 12 2 2 + ||2 ,
where we have used (25) and is defined by
(x, t) = L[] f B (x, t).

2


anm (t) eink1 x+imk2 y ,

n= N2x m= N2y

4.2 THIRD-ORDER APPROXIMATION

0 = + + L[] + 12 2 2 ,

Ny

2


i |K| tanh(|K|h) anm (t) eiKx ,

(42)

Copyright National Academy of Sciences. All rights reserved.

(45)

Twenty-Fifth Symposium on Naval Hydrodynamics

where K = (nk1 , mk2 ) and |K| = (n2 k1 2 + m2 k2 2 )1/2 .


In order to integrate the evolution equations in time, we
use the explicit third-order Adams-Bashforth method. We
also use a pressure beach to absorb waves propagating toward the computational boundaries similar to that used by
Cao et al. (1998). The details of the numerical method to
solve the evolution equations including the pressure beach
and a numerical filter to eliminate instability due to aliasing
error can be found in Kent & Choi (2004a).
Now we summarize our numerical method. For given
(x, tn ) and (x, tn ), we

has been used to rewrite the time derivative of the velocity


potential, which is evaluated using a backward finite difference scheme, while other terms are calculated at the present
time step.
6 EXAMPLES FOR INFINITELY DEEP WATER
For numerical solutions shown here, we consider the case
of infinitely deep water (h ), for which the linear
operator L defined in the Fourier space by (45) becomes


L anm (t) eiKx = i |K|anm (t) eiKx .

1. compute 0 (x, tn ) from (41) and find its Fourier coefficients using FFT,
2. compute
(46)),

F
n

6.1 STOKES GRAVITY WAVES

on the body surface (SB ) (see equation

In order to validate our computer code, we first study the


evolution of free gravity waves in the absence of a body
(f B 0). It is well-known that Stokes waves are unstable to both two-dimensional (Benjamin & Feir 1967)
and three-dimensional perturbations (McLean 1982). To
simulate three-dimensional instability of Stokes waves, we
adopt the following initial conditions for and :

3. solve the linear boundary value problem for


B (x, z, tn ) given by (27)(30),
4. solve the evolution equations (43)(44) for
(x, tn+1 ) and (x, tn+1 ) after computing
from (42).
F

In order to compute
n in step 2, we need to know
the gradient of F (x, z, t) which can be found from the
Fourier series of F given by
F =


n

Anm (t)

cosh[|K|(z + h)] iKx


e
,
cosh(|K|h)

(x, 0) = s (x)
+a0 cos(k0 (1 + p)x) cos(k0 qy),
(x, 0) = s (x)
+ca0 sin(k0 (1 + p)x) cos(k0 qy),

(46)

(47)
(48)

where s and s are the surface elevation and the velocity potential, respectively, for one-dimensional Stokes
waves of wave slope k0 a0 = 0.314 traveling in the positive
xdirection. For small two-dimensional perturbations, we
have chosen = 0.16, p = 0.5 and q = 1.22, for which the
linear growth rate is close to its maximum value (McLean
1982). As shown in figure 2, our numerical solutions of
the evolution equations (43)(44) show the development of
crescent wave patterns first observed experimentally by Su
(1982) and simulated by Xue et al. (2001) using a boundary integral method to solve the fully nonlinear Euler equations. In our simulation, the total energy is conserved typically to 0.01% or less, which demonstrates that our numerical method described in 5 introduces no artificial energy
source or sink.

where Anm is the Fourier coefficients of 0 found in step


1.
Initial conditions at t = 0 for the case of a body or singularity in motion are (x, 0) = 0 and (x, 0) = 0, since its
motion starts from rest, while those for the free wave case
are Stokes wave solutions.
The hydrodynamic force on the body is calculated by
numerically integrating, using the trapezoidal rule, the dynamic pressure on the body SB (t) given by:


1
2
p =
+ 2 |3 |
t



|SB
B
=
V 3 + 12 |3 |2 ,
t

6.2 TRANSLATING SINGULARITIES

where |SB is the total velocity potential evaluated at the


instantaneous body position,
V B is the velocity of the


body, 3 = , z . The chain rule for differentiation

Next we consider a steadily translating three-dimensional


dipole located at z = D. For infinitely deep water, the

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Free surface Comparison Along Centerline


0.04
st

(a)

1 Order
rd
3 Order

0.03
0.02

0.01

0.2
0.1

/l

0
-0.01

0
-0.1

-0.02

0
1

y/

-0.03
-4

-3

-2

-1

Free surface Comparison Along Outer Line


0.02

x/

st

1 Order
rd
3 Order

0.01

/l

0
-0.01

(b)

-0.02
-4

0.2
0.1
/
0
-0.1

x/

-1

velocity potential (or Greens function for the body problem) for a three-dimensional dipole of strength is given,
in a frame of reference moving with speed U , by

Figure 2: Development of crescent waves from initial


Stokes waves of wave slope k0 a0 = 0.314 subject to a
perturbation give by (47)(48) with = 0.16, p = 0.5
and q = 1.22. The numbers of Fourier modes for this
simulation are Nx =128 and Ny =128. (a) t/T = 0, (b)
t/T = 4.338, with T being the wave period.
1

B (x, z) =

0.02

-0. 5
0.01

/l

-1
-1. 5

-2
- 0.01

-2.5
-3

- 0.02

-3.5
- 0. 5

0.5

(49)

(50)

Notice that the velocity potential in (49) satisfies the free


surface boundary condition (B = 0 at z = 0) for the body
problem shown in figure 1(c).

Numerical results for D/l = 2.5 and U/ gl = 1 where


l = (/U )1/3 are shown in figures 3 and 4. The right
half of figure 3 shows the third-order nonlinear solution,
while the left half represents the linear solution. It can be
seen that the diverging waves are more pronounced in the
nonlinear solution.
Similarly a translating
two-dimensional dipole of

strength with U/ gl = 1 and D/l = 5, where l =


(/U )1/2 , is considered and the result is shown in figure
5. Our solution of the linear evolution equations shows
good agreement with the linear analytic solution of Havelock (1926). For this two-dimensional case, the difference
between nonlinear (both second- and third-order) and linear solutions is greater than that for three-dimensional case
with a smaller submergence depth.

0.03

x
x
3,
r+ 3
r

where r is defined as
1/2

.
r = |x|2 + (z D)2

0.5

-1

-2

Figure 4: Comparison of the free surface elevation along


the straight lines at two different transverse locations
shown in figure 3: first-order (- - -) and third-order ()
solutions.

y/

- 1. 5

-3

x/l

y/l

1. 5

x/l

Figure 3: Comparison of wave pattern between firstorder (left half) and third-order (right half)
solutions for
a translating three-dimensional dipole of U/ gl = 1 and
D/l=2.5, where l = (/U )1/3 .

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

0. 4

0. 2
linear analytic
1st Order
2nd Order
3rd Order

0.15

0.3

0.1

0. 2

0.05
0. 1

/l

g/U

-0.05
-0.1

-0.1
-0.2

-0.15
-0. 2
-30

-0. 3
-15

-25

-20

-15

-10

-5

10

-10

6.3 TWO-DIMENSIONAL BODY


For a two-dimensional body in infinitely deep water,
Greens function satisfying (32)(34) is given by


1
log (x x )2 + (z z  )2
4


1
log (x x )2 + (z + z  )2 .

As explained in 5, we have used a periodic array of this


Greens function in our computation.
To approximate the velocity potential B in (31) numerically, we use a desingularized method (Beck 1999): Point
sources are distributed at discrete points (xj , zj ) displaced
slightly into the body in the normal direction from collocation points on the body surface, so that
j G(x, xj , z, zj ),

Fourier modes Nx = 15/wavelength to solve the evolution equations (43)(44) and the number of sources for the
body problem NB = 40. See Kent & Choi (2004b) and
Kent & Choi (2004c) for more numerical results for twodimensional translating and oscillating bodies.

G(x, x , z, z  ) =

NB


Figure 6: Free surface elevation of a steadily


translating
circular cylinder of radius D/R = 5 with U/ gR = 1.79.
The third-order solution () is compared with the firsorder solution and the fully nonlinear numerical solution
of Scullen & Tuck (1995) using a boundary integral formulation.

Figure 5: Free surface elevation () for a two-dimensional

dipole located at z = 5l in uniform stream of U/ gl =


1, where l = (/U )1/2 .

B (x, z) =

-5

gx/U 2

x/l

7 DISCUSSION
We propose a new third-order nonlinear formulation to
solve unsteady wave-body interaction problems in water of
arbitrary depth.
We first find numerically the body velocity potential
(B ) satisfying the exact body boundary condition and the
simplified free surface boundary condition, which can be
easily solved using, for example, a singularity distribution
method. Since the boundary condition at the mean free surface is simple (B = 0 at z = 0), it is no longer necessary
to evaluate complicated free surface Greens functions containing multiple integrals which appear in classical linear
free surface formulations.
After solving the body problem at the instantaneous
body position, it is required to solve the system of coupled
nonlinear evolution equations to update the free surface elevation () and the velocity potential at the free surface ().
It has been shown that the system can be effectively solved
by using the pseudo-spectral method described in 5.
Advantages of our formulation include: (1) compared

(51)

j=1

where NB is the number of sources to represent the body.


Imposing the body boundary condition (30) at the collocation points, the source strength j is found by solving a
system of linear algebraic equations. Figure 6 shows the
free surface elevation due to a circular cylinder of radius
R moving with speed U/(gR)1/2 = 1.79. It is interesting to notice that our numerical solution shows excellent
agreement with the earlier fully nonlinear numerical solution of Scullen & Tuck (1995) using a boundary integral
method. For this simulation we have used the number of

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

with other spectral/pseudo-spectral methods, since our system of evolution equations is closed, no intermediate steps
to close the system are required and therefore our numerical method is more effective; (2) as in the boundary integral formulation, we solve only two-dimensional equations for three-dimensional problems; (3) compared with
the mixed Eulerian-Lagrangian method, no distribution of
singularities along the free surface is necessary. On the
other hand, the limitations of the method are that (1) being
perturbation-based, the method does not allow wave breaking to occur, and (2) the method is valid up to third order
in wave slope, though a test with traveling wave solutions
has shown promising results even at relatively high wave
amplitudes.
Here we present numerical solutions of a translating
two- and three-dimensional dipole and a translating twodimensional submerged body in this paper, more general
three-dimensional submerged or floating body problems
are under investigation.

solitary-wave interactions. J. Fluid Mech. Vol. 118, 1982


pp. 411443.
Havelock, T., The method of images in some problems
of surface waves, Proc. Roy. Soc. A. Vol. 115, 1926, pp.
265277.
Iafrati, A., Mascio, A. D., and Campana, E. F., A level
set technique applied to unsteady free surface flows,
Int. J. of Num. Methods in Fluids, Vol. 35, 2001, pp. 281
297.
Kent, C. and Choi, W., A fully nonlinear theory for surface wave hydrodynamics. Part 1. Pressure forcing in deep
water, Submitted for publication, 2004a.
Kent, C. and Choi, W., A fully nonlinear theory for surface
wave hydrodynamics. Part 2. Interaction with a submerged
body, Submitted for publication, 2004b.
Kent, C. and Choi, W., A new numerical method to
compute nonlinear hydrodynamic forces on a submerged
cylinder,
Proceedings of 19th International Workshop
on Water Waves and Floating Bodies, Cortona, Italy,
2004c.
Lin, R.-Q. and Kuang, W.,
Nonlinear waves of
a steadily moving ship in environmental waves,
J. Mar. Sci. Technol. Vol. 8, 2004, pp. 109-116.
Liu, Y., Dommermuth, and D. G. Yue, D. K. P., A higherorder spectral method for nonlinear wave-body interactions, J. Fluid Mech. Vol. 245, 1992, pp. 115136.
Longuet-Higgins, M. S. and Cokelet, E. D., The deformation of steep surface waves on water. I. A numerical method of computation, Proc. Roy. Soc. Lond. A Vol.
350, 1976, pp. 126.
McLean, J. W., Instabilities of finite amplitude water
waves, J. of Fluid Mech. Vol. 114, 1982, pp. 315330.
Scullen, D. and Tuck, E. O., Nonlinear free-surface flow
computations for submerged cylinders, J. of Ship Res.
Vol. 39, 1995, pp. 185193.
M.-Y. Su Three-dimensional deep-water waves. Part 1.
Experimental measurement of skew and symmetric wave
patterns, J. of Fluid Mech. Vol. 124, 1982, pp. 73108.
Wehausen, J. V. & Laitone, E. V., Surface waves,
Encyclopedia of physics Vol. IX, 1960, pp. 446-778.
Xue, M., Xu, H. B., Liu, Y. and Yue, D. K. P., Computations of fully nonlinear three-dimensional wave-wave and
wave-body interaction. Part 1. Three-dimensional steep
waves, J. of Fluid Mech. Vol. 438, 2001, pp. 11-39.

REFERENCES
Beck, R.F., A fully nonlinear water wave computations
using a desingularized Euler-Lagrange time-domain approach, Nonlinear Water Wave Interaction, WIT Press,
1999, pp. 1-58.
Beck, R.F. and Reed, A. M., Modern computational methods for ships in a seaway, Transaction SNAME, Vol. 109,
2001, pp. 152.
Benjamin, T.B. and Feir, J.E., The disintegration of wavetrains on deep water. Part 1. Theory, J. of Fluid Mech.
Vol. 27, 1967, pp. 417430.
Cao, Y., Beck, R.F. and Schultz, W., An Absorbing
Beach for Numerical Simulations of Nonlinear Waves in a
Wave Tank, Proceedings of 13th International Workshop
on Water Waves and Floating Bodies, Delft, Holland,
March 1998.
Choi, W., Nonlinear evolution equations for twodimensional surface waves in a fluid of finite depth,
J. of Fluid Mech., Vol. 295, 1995, pp. 381394.
Colagrossi, A., Landrini, M. and Tulin, M. P., Near shore
bore propagation and splashing processes: gridless simulations, Proceedings of 6th International Workshop on
Wave Hindcasting and Forecasting, Monterey, 2000.
Dommermuth, D. G. and Yue, D. K. P., A higher-order
spectral method for the study of nonlinear gravity waves,
J. of Fluid Mech. Vol. 184, 1987, pp. 267288.
Fenton ,J. D. & Rienecker, M. M. A Fourier method
for solving nonlinear water-wave problems: application to

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Alan Brandt
Johns Hopkins University, USA
To what extent will your method be able to
simulate a surface piercing flow considering the
importance of wave breaking and entrainment in such
flows?
AUTHORS REPLY
The method presented here is unable to
simulate flows where the free surface is re-entering
the fluid domain and makes no attempt to model such
flows. In fact our formulation is based on the
asymptotic expansion and it is no longer applicable
when the free surface is double valued. The purpose
of this research is to provide a fast and robust method
for simulation of non-overturning free and forced
waves.
It should be noted that other inviscid
methods including the Mixed Eulerian-Legrangian
method are not able to simulate re-entering flows and
either artificial viscosity or pressure patches are often
used to prevent such flows. In our method, a filter in
Fourier space is used to remove the steep waves
leading to wave breaking and is found effective in
many flow situations.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION

AUTHORS REPLY

Lawrence J. Doctors
The University of New South Wales, Australia

For
computation
period with
fourth-order
steps/period.

This most impressive paper provides a


theory and a set of very encouraging results for the
predictions of non-linear wave effects in an inviscid
fluid. For example, Figure 2 illustrates the interesting
development of the lateral instability from initial
Stokes waves.
I would like to ask the authors to comment
on the following:
(a) Why did they specifically choose a third
order approximation to the free-surface boundary
condition? Presumably there is insufficient modeling
of the non-linear effects using only a second-order
approximation.
AUTHORS REPLY
We truncate the expansion at third order
because the nonlinear resonant interaction of gravity
waves occurs at this order. Also we notice that the
second-order solution of the translating circular
cylinder problem is not as accurate as the third-order
solution when compared with the fully nonlinear
numerical solution using a boundary integral method.

the simulation shown in figure 2, the


time is about 25 seconds per wave
a 2.4GHz Pentium PC. We use the
Runge-Kutta method with 91 time

DISCUSSION
Lawrence J. Doctors
The University of New South Wales, Australia
(d) Can the method be easily extended in
order to incorporate a surface-piercing body, such as
a ship hull?
AUTHORS REPLY
The formulation presented in this paper is
still valid for a surface-piercing body, although the
numerical implementation is more complex. The
work is underway and will be reported in the near
future.
DISCUSSION
Lawrence J. Doctors
The University of New South Wales, Australia

DISCUSSION
Lawrence J. Doctors
The University of New South Wales, Australia

I would like to thank the authors again for


their original contributions to the understanding and
prediction of non-linear waves.

(b) How difficult would it be to implement a


higher-order approximation and what would the
implications be, both from the point of view of the
complexity of the algebra and the required numerical
computations?
AUTHORS REPLY
A higher-order approximation can be easily
found by the recursion formula, as shown in Kent and
Choi (2004b). Due to the recursive nature of the
formulation, many terms appear repeatedly in the
evolution equations and therefore the increase in
computational cost is not so significant when
evaluating the higher-order terms.
DISCUSSION
Lawrence J. Doctors
The University of New South Wales, Australia
(c) What are typical computation times?

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Arthur M. Reed
Naval Surface Warfare Center, Carderock Division,
USA

At present our numerical method for


surface-piercing bodies is still under development
and we will report results in the near future.

The authors have formulated the problem


with a body present, as an idealized flow about the
body, with a simplified free surface boundary
condition (double body flow or flow with an image
body) and a problem satisfying the prescribed
nonlinear free surface boundary condition that
corrects the free surface from the idealized flow
about the body. Could the authors comment on how
they would formulate the idealized body flow for a
3D body piercing the free surface, in incident waves?
How would the authors deal with the motions of that
body? How would they compute the forces on the
body moving in waves?
AUTHORS REPLY
We have found that the requirement for
image potentials for the body problem can be relaxed
and is no longer used when solving submerged body
problems. Although the formulation for a submerged
body is equally applied to a surface piercing body,
there are several numerical issues for which more
careful considerations have to be taken: for example,
how to treat the intersection points and how to solve
the evolution equations in a horizontal plane with a
small domain inside the body. Currently a surface
piercing body in motion is being considered and
validation of numerical results is still underway.
Our current approach for a surface piercing
body can be summarized as follows. In the
(idealized) body problem, we impose the body
boundary condition on the instantaneous body
surface immersed under the undisturbed free surface.
The effect of incoming waves appears in the
modified body boundary condition. After we find the
body potential, we solve the evolution equations for
the free surface variables, which include those of
incoming waves. Preliminary two-dimensional
calculations for a circular cylinder undergoing
prescribed small vertical oscillations show
encouraging results that agree well with linear theory.
For larger amplitude motions, the nonlinear result has
yet to be validated. The hydrodynamic forces are
found by numerically integrating the pressure on the
body. The pressure is found using the full form of the
Bernoulli equation and the time differential of the
velocity potential is found using backward
differencing, which can be improved.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

! !"! #$%& ! '

(()

*+," - .
0
+

/
+

'
0

$
'

'
'
' '

+
1

1
-

'
3
3

'
$

!
1

'
1

1 '
+

'
+

'
*

.
+

'

1 '

'

'

'
'

+
'
'

$
'
'

/
+

+
/

+
'
+

/
'
0

'

'
0

+
/

2
$

/
4
'

'

/
'
+

'

'

'

!
-

+
'

'
'
.

*
$

6
4

/
-

'

4
4
-

'
. -

Copyright National Academy of Sciences. All rights reserved.

'

Twenty-Fifth Symposium on Naval Hydrodynamics

'

'
+
+

'

4
1

'

' 4

* .

'

'

'

/'
'

*
.

9
(

'

+
*

'

+
'

0
0

'

+
;

'
&$

* 9 .

/'

(
*

+
6

:
' 4

'

* .

*% .

'

% @% =*

'
'

'

< *

;
* .
* .

+
- * .

*
'

.
*% .

4
0

'

.
'

4
'

'
4

+
*% .

. <

'
'

'

4
4

* .

+
+
+
*% .

1
'

'

' 4

*%.

+
*

>

'
'
+

*
. " #!
$
%
&
$
#
!
&' "
&( "
&"
&
& )
)
&
$
& ! " ( *& ) ( " *
&!
&% * " !
&
#%
( "*
)
+) "
( "
*$ (
$
,&
!! !
# "
,
"
&
%
*"
(&
) %
*
-&
) ) *!
%
*
"
! "%
%*
,.-&
)(

%
#
( "

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

"%

%
%

-&
%

*
*

$ (
$ (

(,

'

'

"

1
'

$ (.

'

'
*

$
'

A
$

1
$

*% .

*% .

'

'

*%BC&.;

(
*

'
+

*& .

4
=

*& .

. <

;
'

*& .
+

5
/

'

;
*& .
:

'
*

+
(

'

'

.
;
*& .

+
4

+
'

'

4
;

%= % *

*& .

<

* .

'
9

+
;

'

'

(.

*&'.

(* .

'

'
>
;
4

'

4
>

.'

,0-

>

'

*& .
'

*& .

'
'

&

Copyright National Academy of Sciences. All rights reserved.

*%BB&.

Twenty-Fifth Symposium on Naval Hydrodynamics

$ (.

>
0

* .

*) .
A

'

*% .;
#

(
%

*) .

$.

'

*) .

+
*

;
*) .

+
;
*) .

;
(

*.
%*

*) .

*)'.

(* .

'
(

;
*E .
+

*) .

%*

0 '

*E .

. <

(* .

*E .

">

(.

*E .

' 4

,1-

)
4

*E .

% =%*

<

*E .

'
4

'
;

'

(.

*E'.

(* .

+
7

* (((. +

'

;
*

%*

*E .

1'

,2-

*
.

'

'

' 4

*) .

*) .
*

*) .;
%

+
;

* .

Copyright National Academy of Sciences. All rights reserved.

*C .

Twenty-Fifth Symposium on Naval Hydrodynamics

'

*C .
%*

'

>

*# .
*

;
%* %

*C .

;
*C'.

(* .

*E .
/

'

;
(

"

%*

*C .

(.

*# .

*) . -

*C .
;
*C .

. <

%*

*) .

*C .

*E .

*E . +

'

*E . >

*C .
/

,4-

!
$
+

>

;
%

*# .

'
*

4
'
4

$.

*# .

'

/
'

*# . +

' *% .

*%
*

'

'

'
*% .
;

'

2
/

'

*C . +
'

>

+
'

*C .

' ;
'

>
;

4
*

!
4

*# .

$ .

'
$

!
*# .

'
*

*) ..

4
4
*C .
%

'
+

Copyright National Academy of Sciences. All rights reserved.

1
;

Twenty-Fifth Symposium on Naval Hydrodynamics

#%

2
"*

%*#
!

"

%
!

5
/

#%

"*

/
;

4
!
" #! /
+
/
6

) )
#%

)
(

#! =

"*

"

.<

*B .

+
0
'

'

'

'

6
'
6 )

;
! !
) (
"*

!
! !
$ ) (
#%
( "*
( "*
" #!
&)
) ( "*
/
+
'
*

(
"*

)
+ )

)
)
! !

"*

#
( "*

%*
) )
! !
)
/

8 4

0
+
'

'

'

'

'
-

'
'
+

'

'

'
'
';

!
*B .
*B .

';
7
)
) "
!
" #! #
*&
( "*
"
) !
-

! !
#!
)

"*
) #%

&)
! (
" #! #

"
$ ) ! !
* #%
7
;
*B .

'

&

%* ) ! !
) ( " *&!
&
% *
"
)
!
&
"
!
$ *& ) ) ( " *&!
&
% *
"
) #%
&
) " ##
# &
%*
$ ) #%
( "*
) #%
) #%
( "*
)
/ )
$
&"
)
! !
&
)
) ! !
) !
) ) " #! #
* #%
/
+
;

"*

+
#! *

" #!

'

!=

&

.<

*B .

5
;
8
!
" #! #
* #%
! !
( "*

" #!
&
)

"

'

#%

Copyright National Academy of Sciences. All rights reserved.

'
'

Twenty-Fifth Symposium on Naval Hydrodynamics

*B .

$ (.

'

*B .

*%% .
.

! !

*%( .
.

"

'
9 ## .
%
)
! !
" #! #
*
(

'
;

!*

*%( .

$ (.

%
'

*%( .

%
%*
) " #! #
*
/
.
9 ##

;
#! *

'

;
*%( .
*%( .
*%( .

)
9 ##

>

(
.

;
" %

%*

)
" #! #

! !
+ )

* #%

/
1
+

'
1

'
;
+

'
'

0
'
'

'

4
*
9
' 4

''
' 4

'

36

.
6

'
/

1+
*B.
*%(.
'
%
!
) " #! #

4
9 ##
! !

"
(

>

'

'
'

'
!
'

'

)
*

'

'

/
+

'

*3
30 3
3 FF
'

'

.
3
3 6

3
3 +
6

3
3

3
3

'

'
"

*
*

$ (.

4
!*

$ (.

'
-

'
'

*%% .
;
*%% .

;
.

' '
'
'

*%% .
*%% .
#! *

'

>

'
6

;
6

*%% .
*%% .
4
*%% .

*%% .
'

:
*

..
$

+
'

*%( .

Copyright National Academy of Sciences. All rights reserved.

*C .

Twenty-Fifth Symposium on Naval Hydrodynamics

'

'
+

'
!

'
'

'
*

'

'
'

* 9 .

'

* 9 .

/
+
+

'

/
'

$ (.

/'

'
+

'

'

!
* 9 .

'
0

'

0
+

/
'

'

+
/

+
+
*

4
'

'
,

'

#! =

'

$ ( .<

!(

$ (.<

!(

;
#! =

'

'

*%

'

'

'

+
'

!=

!(

.<

1
, .-

0 '
$

*% .

!(

'

/
/

!(

'

"

/
+
*%
!

*% .

$ (.

;
) )

*% .
*% .

"
4

4
'
'

0
+

'
+

'

'
H

'
'
'

/ *%BB . H

*%BB).

*
.

+
0
'
'

G / $ $
'
*%B#%. A

>

/$

Copyright National Academy of Sciences. All rights reserved.

I
$ $
*%BC%.

Twenty-Fifth Symposium on Naval Hydrodynamics

/ *%BB .

0
'

;
*%

% *(

*%&6.

+
1
1

% *(

;
, 0-

"

'

'

.
1

+ / '

'
'
'

*% .
*% .

;
*%) .
*%) .
*%) .

4
*%

*% .

:
*

!(

+
1

;
(
*

.
*

*% . *%&.
6

*%& .

*%).

'
7

6
-

/'

$1
0

*%) .

*%& .
%

*% . 1
'

'

'

'
+
'
!

+
(

'
-

*%& .
;
*%& .
*%& .

. <

4
*%& .

*%
'

'

'

* . 5
/
;
(

*% .

*%& .
'

">

'

/'
+

'
1

;
% *(

% *(

*%&'.

%* % !

;
9 ## 0
) )
"
! "* $
%
! "* $ ) "
!
$
4
'
'
/
'
!
+
/

;
*%& .

+
;
*

(.

*%& .

(* .

(* .

Copyright National Academy of Sciences. All rights reserved.

+
"
(
%

&)
)
"
/

5
5
/

'

'

9 ##

Twenty-Fifth Symposium on Naval Hydrodynamics

0 -

'

/
;
0

'

+
!

'
'

* !

! .

#
*
#

#
#

'

!.

*%E .
* .

'

! (

'
:

0
! !*

*% .

'

*%& .

1 +

(
%

* !

"

%".

*%

*%C
*%C
*%C
*%C

*%& .

*% .

'

;
(

'
*E . -

*E . *E .

*% .
*% .
*% .

.
.
.
.

*%C .

*%E .

!.

"

'

0
%

#<

'

#
=
*
#

*%E .

".

*%

'
0

! (

'

3
;
% *(

%* (

+
!

*% '.

;
*%#
*%#
*%#
*%#

*% .

;
!*

*% .

!( * .

(.

'

!( * .

'
!
!(

!(

!(

(
% *(

*% 6
.

% *(

!(

'
#! =

#! =

3'
%

"

'
*C .

*C . *C .
A '

.
.
.
.

, 3-

!=

*%

!
!

(
!

*
*

.<

*%B .

.<

*%B .

! (
! (

.
(

.<

%(

Copyright National Academy of Sciences. All rights reserved.

*%B .

Twenty-Fifth Symposium on Naval Hydrodynamics

#! =

! (

#! =

! (

! (

!=

'

*%B .

.<
.<

*%B .

! (

.<

'

'

*%B .

! (

'

'

(
(
+
4

*%B.

/
1

' 1

1
2
/

,
*%B.
1
9 ## 1
%
& )
#
%
)
#
)
) ) ##
% * "
&
!
) $
"
"
)
/
7
9 ## 1
0

';

'

#%
, 7% "

4
5
/
'
+

#! =

! (

$ (.<

#! =

! (

* ( .

!=

+
'

$ (. <
*

* ( .

!
!

! (
!

! (

4
.<
!

' > 1'

* ( .

H J*%BBE.

$ (.

*% .
*%B.
+

* (.

'

0
'

+
'
0

9 ## 1
'

9 ## 3
)
"
9 ## 1&
)
"
"
" #! %
)) $
" &
)
% "
"
!
! !
$
"
*! /
9 ##
3
'
'
1 '
1 '
/
+
;
'
'
9
4
1 '
'
1 '
'

)
"
%

'

:
4

*%&.

* 9 .

;
* % .

* .

'

* % .

( (

0
+
9

*%
>

2
!
9 ##

'
1

'
/
9 ##

2
1

%%

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

* % .

>

'
' 4

4
!

'
'
*% .

*E .
*%& . *%C.

*E .
*%B .

'

;
*

* ) .

/'

;
!

'

*
*

.
.

%= % *!

<

* ).

5
5

* ).

;
*

* ) .

;
*

* )'.

*E .

*E .

+
*% '.

*% 6.

'

* %. -

*E . /
* %. %*

*% '.

%*

;
*

* ) .
* ) . * ) .

* ) . -

. -

'
*

* ) .

.
4

*%

'
-

"

"

* ) .
'

%
*

"

* & .

'
4

;
%

1+
'

H
4

* ) .;

&

&

* & .

+
4
>
;

4
*

'

*%

*% .

;
!

&
'

.
%

=*

<

* ) .

+
*E .
1
+

4
*%C.

&

*E .

'
+
>

'

;
* ) .
* ) .

!
!

/'

*C .

6
3
*%#.

* ) .

'
'

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

'

.
%

'

* E .

+
'

* E.

* C.
4

'

;
8

* E .
* E .

%* % !

'
* 4

* . * ).
* 4

* E .

* E..
* &. * .

H
* C..

* E .
>

;
*
*

* ).

. *

* E .

*C .

'

*C .

'
/

* E.
* E .

* E . A

'

* E.
* .

'

* &. !

*
H

* %. -

*C . /

* 4
'

* C..
>

'

* ) .

'
+

4
4

'H
H

'
'

'
+

3 !'
4

!
%

* C .

4
.

;
!

&

&

=*

<

+
*C .
3

*
4

* C .

'

'

/
'
*%

* %. +

*% . +

4
' '
'

4
-

*%

. *%E .

'

/
4

'

4
'

/'
*% . +

'
; 4

*%E .

* &.

* ) .
4

%&

Copyright National Academy of Sciences. All rights reserved.

* E .. +
4
/'

Twenty-Fifth Symposium on Naval Hydrodynamics

/
*

*
* ).

'

. -

/
4
* )'.

'

'
+

'

* ) .

+
/

/
/
* 4

' +
/

* ) .

.. +
/'

'

- '
4

* E .

'

''

'
+
+

"

'
'

'

>

/
''

4
'

* E . +

' /
+

>

'

/
'

'
* E . +
/
4
* E .

0
'

'

/
4

/'

!
' 4

'
2
/

* ) . +
4

/'

* E .

4
+

'
+

/'
N

1
-

'
-

+
+
*
4

.
"
* ) .

K 1 L M / *%BC%.
5

$
'
H

* ) . !
4

*%BBE.

* ).

'!
/

A
2

5
0

'

'
'
5

2
+

%)

Copyright National Academy of Sciences. All rights reserved.

'

Twenty-Fifth Symposium on Naval Hydrodynamics

'

'

'
+

$
0

$
+

'

0
$

'

7
'

4
$

!
'

'

'
'

/
1 '

'
'

1 '
*

'

$
$

'

'.

'
'

'

>

/
+
4

'

'

'

/
'

7
+

N K O! ' A
H

'

>

"
'

P
(((

%(#$%%)
/
*

> 1'
>
%BBE

'
4

7"
"
(% %B#%

&B

OI
>

K 1 L M / G OH
"
5
P I
% H: $H
H
: 1
%BC%

4
4
4
'

G
'

*%

7"

. +

>

'
80
&) %BC%

:
>

"

/
PH

'

Q /

:
*I5>. G
H

'
4

H JG O
P
' $I

"

>
7

84

'+

O!

'
!

4
:
!

'

"

/
>

G
'7

%BB

7
, 8 O,
' :
+
P 5
:

Copyright National Academy of Sciences. All rights reserved.

&&$ )
!
8'

'

Twenty-Fifth Symposium on Naval Hydrodynamics

!
G
%E $ ((

%BB)
H

>

D
I

'

'
G
,

PG
G

%BBE

7 /

$E&)

O
$
'7

%BB&
G O

/
$

>
Q / %BC&

84

P 7 /R

G
%( $%(#

PN

%E

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Emile Baddour
National Research Council-Canada, Institute for
Ocean Technology, NL, Canada
The present reviewer would like to
congratulate the author for his paper and his new
perspective of a rational formulation to a large class
of fluid flow problems.
In this paper the author has researched and proposed
a general approach to fluid flow analysis and
solution. A rational framework is presented that will
allow the conceptual design of a solution by a
decomposition method using component flows
defined by the author. Two structures with their
algebras are introduced and used simultaneously for
that purpose.
The author in his paper defines the following
basic flows:
1) Generic Structure and flow types:

Entire flow

Irrotational Flow

Diffusive Flow

Turbulent flow.
The above types form a hierarchy of flows that must
satisfy boundary and initial conditions as well as
compatibility conditions hence suggesting a
procedure for flow analysis over the whole fluid
domain. The author proves that an entire flow would
be in general decomposed into irrotational, diffusive
and turbulent flows each defined over the whole fluid
domain. The author presents relations and operations
(structure) on these flows using in parallel a process
of operations defined with the use of:
2) Heuristic Structure and its corresponding flow
types, namely:

Ambient flow

Disturbance flow

Disturbed flow
In the latter structure the Ambient flow is known
while the Disturbed and Disturbance flows are
unknown. The heuristic analysis defines the
disturbance flow as the final sought-for flow.
Together with appropriate boundary, initial and
compatibility conditions, the author uses relevant
Generic and Heuristic information to develop a
procedure and a sequence of operations between
component flows (of generic and heuristic types) for
the complete fluid flow solution.
As expressed in the paper, one important
feature of the proposed framework is that induced
forces on boundaries are additive for the component
flows making up the disturbance flow. Solving the

component flows will hence be the objective of any


rational approach to a complex fluid flow problem.
The author also presents a definition of limit
problems of the disturbance flow hence presenting
the application of the proposed framework to the
solution of the maneuvring of ships at slow speeds
showing the practical importance of his method.
As presented the framework shows
dependencies at a higher structural level rather than
at particular methods of solutions of component
flows. The structure is independent of specific
numerical/analytical algorithms used to solve for the
component flows. Hence it is possible to use the
framework to precisely elaborate what are the needs
for a complete fluid flow solution using its
component flows. It does not offer a guarantee of the
practical solution of a component flow, which
depends on the analytical or numerical methods
chosen to solve the corresponding boundary and
initial value problem.
Could the author comment on the problems
of uniqueness of the decomposition of a particular
flow into the component flows?
Would he comment on the causality
condition and how the proposed decomposition will
take care of it?
Would the author comment on the source of
turbulence in his turbulent flow? Would his turbulent
flow (or in that matter, the irrotational and diffusive
flows as defined in the paper) be amenable to current
available solvers?
Also, would he elaborate further on the
respective roles of the generic and heuristic flow
structures in finding the solution to a fluid dynamics
problem?

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

AUTHORS' RESPONSE

5
8 9

! "#
$ % &'()*+(,-./0.
12
3
4
5

6
6

6
6

!
7

1
59

6
6

63

6
5

5
%
59

8
8

%
6

:
59

5 8

6
6

6
9

9
%
59

9
'& +

5 8
6

8
:

6 8

5
5

'
'5

6
6

$
6
<

6
$

9
5

'
5

9
5

'9

%
6

'9
+
'9
+
5 :

5
8

'

9
5
5

'
+
#

5
$

5
9

6
6

5
:

$
8
5
8

5
'

%
6

9
9

"

8
6
:
6

9
5
9

5
#
6

5#

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics


St. Johns, Newfoundland and Labrador, CANADA, 8-13 August 2004

Air entrainment induced by the impact of a planar


translating jet on a flat free surface
Alessandro Iafrati1 , Emilio F. Campana1 , Ramon Gomez Ledesma2 , Kenneth T.
Kiger2 and James H. Duncan2
(1 Italian Ship Model Basin, Rome , 2 University of Maryland, USA)
ABSTRACT
The free surface flow generated by the impact of a twodimensional jet onto a still water surface is numerically
and experimentally investigated. The plunging jet is assumed to be inclined and translating with respect to the
still water. The study is aimed at exploring the mechanisms leading to the air entrainment processes taking
place during the jet entry. Different flow conditions are
investigated by varying the inclination angle and the translating velocity of the impacting jet, although only a single condition is presented in the current work. Experimentally, the free surface dynamics are observed with a
high-speed digital camera. The same problem is numerically simulated using a two-fluid approach that allows
for the continuous description of both phases, while accounting for the change in interface topology resulting
from the air entrainment process. Both experimental and
numerical results are presented, and comparisons among
the two are established for a case with zero translation
velocity. Although the numerical tool still requires further improvements, the comparison with experimental
data in terms of free surface dynamics and the time and
depth for pinch-off of the entrained air bubble are in general qualitative and quantitative agreement, presenting
encouraging results for further development.

INTRODUCTION
The flow generated by the impact of a planar water jet
onto a still water surface is investigated. The study, carried out both experimentally and numerically, is aimed
at exploring the mechanisms leading to air entrainment
during the jet entry.
Air entrainment has important practical implications
in ocean processes and naval hydrodynamics. In the
open ocean, wind generated breaking waves induce entrainment of air bubbles into the water and water droplets

are ejected into the atmosphere as a result of the plunging jet. These bubbles and droplets greatly enhance the
gas and heat transfer between air and water, with important biological consequences.
Large plunging breaking waves are also generated at
the bow of fast ships with a pronounced flare. The impact of the plunging jet onto the water surface entrains
air bubbles that travel along the hull and eventually flow
into the ship wake. These bubbles have a twofold undesirable effect on military ships. First, the entrainment
process, the subsequent fragmentation of the bubbles
into smaller volumes and the resonant pulsing of these
bubbles generate intense hydrodynamic noise that affects the signal received by the sonar located in the fore
part of the hull. Second, the finest bubbles remain in the
ship wake for long periods of time and scavenge surfactants during their slow rise to the surface. Once the bubble do surface, they deposit the surfactants on the surface
with profound effects on the propagation and damping
of capillary waves in that region, leaving a clearly detectible signal observable by high-altitude radar.
Experiments on a stationary plunging axisymmetric
jet have been carried out aimed at measuring both the
minimum impact velocity required to cause air entrainment, see Sheridan (1966) and Lara (1979), as well as
the void fraction, bubble size distribution and radiated
noise, see Kolaini & Crum (1994). For the same problem, the role played by a horizontal translation velocity
in a vertically oriented axisymmetric jet has been studied more recently by Chirichella et al. (2002), who measured the critical translation velocity for air entrainment
for a wide range of jet impact conditions.
In order to get flow conditions closer to those generated by breaking bow waves, the transient impact of
a translating planar jet is studied in the current work.
Experimentally, this is achieved by ejecting a water jet
from an high-aspect ratio rectangular nozzle placed parallel to the recieving pool. The resulting flow is observed through a high-speed digital camera taking 1000

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

frames per second, and for several conditions, particle


image velocimetry is used to measure the flow around
the leading edge of the impact crater. The nozzle, the
jet cutter and the camera system are mounted on an instrumented carriage that translates along the tank at a
constant speed, up to 0.6 m/s.
The same flow is numerically studied through a twodimensional Navier-Stokes solver which describes the
two-fluid flow of air and water as that of a single, incompressible, fluid with variable density and viscosity
(Iafrati & Campana, 2003). The interface location is
captured by a level-set technique that allows for simplified handling of the topology changes induced by the
impact and by the entrapment of air bubbles. Initial and
boundary conditions are properly chosen to approach as
close as possible the experimental conditions.
Experiments are carried out by varying the impacting angle of the jet, its speed and thickness at the still
water level, in addition to the translation velocity of the
carriage. The movie recorded by the camera shows the
development of the cavity induced by the jet impact. In a
later stage, under the gravity effects, the upper part of the
cavity collapses and touches the jet surface. This pinchoff event entraps a packet of air which is successively
fragmented into many small bubbles. At the present
stage of the study, numerical simulation is performed
only for a case without translation velocity. Several of
the experimental findings can be seen from the numerical results. A more stringent comparison is established
by overlapping the free surface profiles extracted from
the numerical results on the experimental frames and a
rather good agreement is found in terms of the cavity
shape and of the size of the entrapped air pocket.

EXPERIMENTAL SETUP
The experiments were conducted in a 0.76 1.22
7.31 m (width height length) steel-reinforced glasswalled tank, with a planar jet nozzle mounted on a computer-controlled carriage. A skimming filtration system
was used in conjuction with a diatamacous earth filter
and 3 ppm chlorine to maintain nominally clean and
repeatable free surface conditions. The nozzle was supplied with a continuous stream of water from the tank by
a 3-hp centrifugal pump, whose flowrate was controlled
by a variable speed AC inverter. The maximum exit velocity for the 0.6 m wide nozzle depends on the nozzle
opening width, but was limited to less than 4 m/s for the
smaller nozzles widths to prevent excessive backpressure from damaging the flexible supply lines running to
the carriage. The flow was then passed through a turbulence management section composed of a 4 cm thick
layer of 2 mm open-cell foam, followed by a fine screen

Nozzle

Nozzle Inlets

Jet
Open cell
foam

Jet cutting
device

Screen

Tank wall

Troughs
Receiving Pool
Jet confinement
side walls
Camera

Figure 1: Detailed schematic of jet nozzle and cutting


system used to produce a uniform two-dimensional jet
impact on the free surface. Carriage system used to
translate the jet nozzle and cutting device is not shown
for clarity.
mesh. The nozzle then underwent a planar 10:1 contraction designed to prevent separation within the nozzle.
The nozzle was also designed to have a variable width
opening, 0.003 < h < 0.006 m, to allow for different
jet thicknesses at the impact location. A schematic of
the nozzle and jet forming system is shown in Fig. 1.
In order to form a two-dimensional transient impact,
the steady output from the nozzle needed to be conditioned using a pneumatic deflector and cutting device.
The device consisted of a deflector plate that was initially positioned in the path of the descending free jet,
and angled to deflect the jet downstream away from the
intended impact site (see Fig. 2). To trigger a transient
impact, the deflector was rapidly pivoted by a pneumatic
actuator in such a manner that it sliced cleanly through
the falling jet, while at the same time removing itself
from blocking the jets trajectory. This provided a reasonably clean and repeatable interface for the impact
with the recieving pool. Due to the interaction of the
cutting device with the jet, a slight bolus of fluid would
typically form at the jet tip with a diameter that was approximately two to three times the jet thickness.
As the jet descended towards the receiving pool, surface tension acting on the free edges of the jet would
cause the sides to contract, acting to decrease the width
of the jet. The contracted fluid accumulated into rounded
streams along the edges of the orginal planar sheet, giving the cross-section a distinct dumb-bell type shape.
In order to prevent these end effects from influencing
the intended two-dimensional impact, cutting troughs
and transparent confinement sidewalls were placed inside the tank (see Fig. 1). The walls and troughs acted
to cut the three-dimensional edges from the falling jet

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Jet impact velocity, Vj (m/s)


Jet impact angle, j (degrees)
Translation velocity, Vt (m/s)
Jet thickness, Tj (m)

Experiment
3.320.2
72
0
0.0030.0005

Simulation
3.16
7
0
0.003

Table 1: Initial conditions for experimental and numerical case.

Incoming Jet
Cutting/Deflecting
Plate

Impact site

The two-fluid flow of air and water is approximated as


that of a single incompressible fluid whose density and
viscosity varies smoothly through the interface. With
these assumptions, the continuity and momentum equations in generalized coordinates reads:
Um
=0 ,
m

Deflected Jet

Cut and
Undeflected Jet

Figure 2: Schematic of deflector/cutting system used to


generate the transient jet impact.
without disturbing the visualization of the impact while
further minimizing the formation of additional end effects.
A high-speed CMOS camera (Visible Solutions Phantom v4.0) with a resolution of 512 x 512 pixels and
framing rate of 1000 frames per second was used to
record the transient impact events. Image processing of
the movies was performed using Adobe Photoshop and
Matlab to extract the velocity and shape of the jet and
impact crater.
Although experiments have been conducted for a variety of parametric conditions that span a range of impact velocities, translation speeds, jet impact angles and
jet impact thicknesses (Gomez Ledesma, 2004), only a
single case is reported for the current work in which to
compare to the numerical simulations. The conditions
for the case and the associated uncertainties with the
given values are shown in Table 1.

NUMERICAL MODEL
Two-fluids Navier-Stokes solver

(1)



1

1
1 m
(J ui ) +
(Um ui ) =
J
p
t
m
% m
xi



i2
1 m
J
H
(d)
J 1 2
T
Fr
% W e2 m
xi


1

mlji uj
ml ui
+
+ B
G
,
(2)
% Re m
l
l
respectively. In the above equations ui is the ith Cartesian velocity component, ij is the Kronecker delta,
Um = J 1

m
uj
xj

(3)

is the volume flux normal to the m = const. surface


and J 1 is the inverse of the Jacobian. Non-dimensional
ratios are defined as
r
Ur
%w Lr
Ur Lr %w
Fr =
, Re =
, W e = Ur
w

gLr
for Froude, Reynolds and Weber numbers, respectively.
Here, Ur and Lr are reference values for velocity and
length, is the surface tension coefficient while %w , w
are the values of the density and dynamic viscosity of
water, which are used as reference values. By following
the suggestion proposed in Brackbill et al. (1992), the
contribution of surface tension effects to the momentum
equation are modeled as a continuum force distributed
in a thin region close to the air/water interface, and expressed functionally as the gradient of the smoothed Heaviside function
 

d
1 1

+ sin
for |d|

2
2
2

H (d) =
0
for d < (4)

1
for d > +

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

with d denoting the signed distance from the interface,


taken with the convention of a positive displacement being toward the water and negative displacement toward
the air. In the surface tension contribution, the parameter = T represents half of the thickness of the region
across which surface tension forces are spread. Finally,
in (2)
Gml = J 1

m l
xj xj

B mlji = J 1

Step 3

 u
3i u
2i
=
J 3 tDI
t
i
h
u2i ) + Ti (d2 ) +
3 C(
u2i ) + DE (
h
i
2 C(
u1i ) + DE (
u1i ) + Ti (d1 ) +


i2
23 J 1 2 + DI (
u2i ) ,
Fr
1

m l
xj xi

are metric quantities and is the local curvature.


The system of the Navier-Stokes equations is numerically solved through a numerical scheme similar to
that suggested in Zang et al. (1994): Cartesian velocities and pressure are defined at the cell centers, whereas
volume fluxes are defined at the mid-point of the cell
faces. The advancement in time is achieved through a
fractional step approach in which the pressure contribution is neglected when integrating the momentum equation in time (Predictor step) and then is reintroduced in
the next stage (Corrector step) when the continuity of
the velocity field is enforced. To reduce the constraints
of the related stability limit, the diagonal part of the
first viscous contribution in (2) is computed implicitly
with a Crank-Nicholson scheme, whereas all other terms
are computed explicitly with a three-steps low storage
Runge-Kutta scheme. The grid being fixed in time, the
discretized form of the momentum equation at the step
n is

un+1
u
3i = 3
i

The coefficients i , i , i are reported in Rai & Moin


(1991) and in literature cited therein. In the above equations, for the sake of clarity, a compact notation is used
to represent the convective, diffusive and surface tension
contributions:
C(ui )
DI (ui )
DE (ui )

Step 1
J 1 1 tDI


 u
1i uni
=
t

1 [C(uni ) + DE (uni ) + Ti (dn )] +




1 i2
n
21 J
+ DI (ui ) ,
F r2

u
1i u
1i = 1

Ri (1 )
%1 J 1

Step 2

 u
2i u
1i
J 2 tDI
=
t
h
i
2 C(
u1i ) + DE (
u1i ) + Ti (d1 ) +
1

1 [C(uni ) + DE (uni ) + Ti (dn )] +




1 i2
1
22 J
+ DI (
ui ) ,
F r2

u
2i

u
2i

Ri (2 )
Ri (1 )
= 2 2 1 + 1 1 1
% J
% J

Ri (2 )
Ri (n+1 )
+

.
2
%n+1 J 1
%2 J 1

Ti (d)
while

(Um ui ) ,
m


1

ui
=
Gml
m=l ,
% Re m
l


1
uj
ui
=
Gml
+ B mkji
% Re l
l
k
m 6= l ,



1 m
=
H
(d)
,
J

T
%W e2 m
xi

Ri (f ) =
m

1 m

xi

(5)

is the gradient operator in generalized coordinates.


The pressure corrector, , is found by enforcing the
continuity of the velocity field at the end of the substep
(Chorin, 1967; Kim & Moin, 1985). Once the auxiliary
velocity field is found, say u
li , the fluxes at the mid-point
l
m
of the cell faces associated to this velocity field (U
)
are computed by (3). To this purpose, Cartesian velocity components at the cell faces are evaluated through a
quadratic upwind interpolation. In terms of fluxes, the
corrector step can be written as
!
mj
l
G

l
l
m U
m = l t
U

%l j
l1 t

Gmj l1
%l1 j

Copyright National Academy of Sciences. All rights reserved.

(6)

Twenty-Fifth Symposium on Naval Hydrodynamics

l
m
so that, by applying the continuity (1) to U
, the following Poisson equation for the pressure corrector is obtained:
!
l
m

Gmj l
1 U
=

m
%l j
t m

l1
l m

Gmj l1
%l1 j

(7)

When the velocity field is assigned throughout the boundary of the computational domain, (6) provides Neumann
boundary conditions for the solution of the Poisson equation (7). The pressure field can be derived from the corrector term through the equation:
!
 Ri (l )
l
l 1
Ri (
p ) = % J l tDI
.
(8)
%l J 1
As the solution of this equation is not straightforward,
usually an approximate pressure field is obtained as (Rosenfeld et al. 1991):
Ri (
pl ) ' Ri (l ) pl = l + O(t) .

(9)

The maximum Courant


number allowed by the threesteps Runge-Kutta is 3. An additional constraints to
the time step result from the stability conditions required
for surface tension effects (Brackbill et al. 1992)
r
(1 + %a )
t < W e
x3 .
4
Moreover, additional limitations to the time step may
be needed, since not all of the viscous contributions are
treated implicitly.
A multigrid technique is adopted for the solution
of the Poisson equation of the pressure corrector term,
which is the most expensive part of the computational
procedure. A corrector scheme is used for restriction
and prolongation, (Brandt 1992) and a LSOR method
is employed as a high-frequency smoother. Since metric quantities and fluid properties appear into the coefficients of the Poisson equation, a simple average of the
metric and of the distance function is used in the restriction phase. Additional details about the numerical
solution of the Navier-Stokes equations are reported in
Iafrati & Campana (2003).

Interface capturing technique


The interface between air and water is captured as the
zero level-set of a signed normal distance from the interface d(x, t) which, at t = 0, is initialized by assuming

d > 0 in water, d < 0 in air and d = 0 at the interface (Sussman et al. 1994). Physical fluid properties are
assumed to be related to d by the equation:
f (d) = fa + (fw fa )HP (d)

(10)

where HP is a smoothed Heaviside function as it follows from (4). The parameter = P is set so that the
density and viscosity jumps are spread across five grid
cells, at least (Iafrati et al., 2001).
During the time advancement, the distance function
is advected with the flow as a non-diffusive scalar using
d
+ u d = 0
t

(11)

Equation 11 is integrated in time, with the new interface configuration recovered as the d = 0 contour. The
integration of (11) is carried out with the same threestep Runge-Kutta and discretization scheme employed
for the convective terms, that is
dl = dl1 + l tC(dl1 ) + l1 tC(dl2 )

(12)

with d0 dn , d3 dn+1 and


C(dl ) =

 l l 
Um d
.
m

To keep the width of the jump region separating the


two fluids constant through time, the distance function is
reinitialized by computing the minimum distance from
the reconstructed interface to the cell centers at each
time step.

Initial and boundary conditions


The numerical calculations are carried out on a computational domain which, in dimensional variables, extends from 0.3 < x1 < 0.3 m in the horizontal direction and from 0.36 < x2 < 0.08 m in the vertical
direction. Hereinafter, lengths are expressed in meters,
time in seconds and velocity in meter per second. When
the calculation starts, the still water level is located at
x2 = 0 and a jet of water is initialized at the top of the
domain. In order to keep the position of the water jet
fixed during the calculation, the distance function is not
updated in the three cell rows closest to the top boundary.
Initial and boundary conditions are assigned to approach the experimental values at the time of the jet impact with the free surface. Among all the experimental tests, the numerical simulation is performed for the
case with a jet impact velocity of 3.32, impact angle of
7 degrees from the vertical and jet thickness of 0.003
(see Table 1). In the numerical calculation, a water jet

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

0.04

0.03

x2

0.02

0.01

-0.01
0

0.02

0.04

0.02

0.04

x1

0.04

0.03

x2

0.02

0.01

and horizontal direction to give a very fine resolution in


the vicinity of the jet impact site. In the horizontal direction, a uniform spacing x1 = 0.0002 is used in the region x1 (0.068, 0.068), with an increasing geometric ratio of 1.06 up to |x1 | = 0.18 and is subsequently
kept constant once x1 ' 0.0133. In the vertical direction, a uniform grid spacing x2 = 0.0002 is used
in the central region x2 (0.165, 0.01). In the lower
region, the spacing is gradually increased with a growth
factor of 1.06 over the region 0.24 < x2 < 0.165
and is assigned a constant value x2 ' 0.0104 for
x2 < 0.24. The grid spacing is also gradually increased in the upper part of the domain, but is limited
to a growth factor of 1.01 in order preserve the minimal resolution required to ensure proper dynamics of the
falling jet. Above x2 = 0.05 the vertical grid spacing is
kept constant to the value x2 ' 0.001.
In the numerical simulation, the ratio of the gas/liquid
density and viscosity are assigned the same values as
that for air and water, although the viscosity magnitudes
for air and water are assumed ten times larger than the
actual ones.

-0.01
0

x1

Figure 3: Shape of the jet tip before (a) and after (b)
the cut. For the present calculation the cut is made at
t = 0.021 s from the beginning of the calculation. The
velocity field, shown at every other cell center, remains
unchanged.
with initial thickness 0.003 is initialized about the top of
the domain with its mid-line intersecting the top boundary at x1 = 0.015. In the calculation presented below,
P = 0.0003 is adopted, giving the thickness of the transition region to be about one fifth of the jet thickness.
In order to account for the gravitational acceleration,
the two velocity components at the top boundary are set
as u1 = 0.405 and u2 = 3.051, thus resulting in an inclination angle of about 7.56 degrees. To assure a zero
velocity field in air, the two velocity components are
multiplied by HP (d), which implies that the velocity
jump is smoothly spread across the same transition region adopted for physical variables. On the left and right
boundaries of the computational domain, u1 = u2 = 0
are assumed. In order to satisfy the continuity condition, a uniform vertical outgoing flow is assigned along
the bottom boundary to balance the total flux incoming
from the top.
The computational domain is discretized with 768
1024 grid cells that are suitably clustered in the vertical

RESULTS
As the simulation starts, the jet is forced to move through
the quiescent air, and due to the initial shape and velocity
field, an odd tip shape develops prior to jet impact (see
Fig. 3a). Numerical tests have shown that some details
of the impact may be influenced by this shape. Hence,
in a way very similar to what is done experimentally, the
tip of the jet is cut before the impact to keep it as sharp
as possible, and the simulation is re-started from the new
configuration, as shown in Fig. 3b. After the restart, the
water jet falls down until it touches the free surface at
the impact time tI ' 0.0325. The jet speed at impact
is about Vj = 3.16, which is slightly smaller than the
value Vje = 3.30 estimated by energy considerations.
It is not immediately verifiable what the exact cause of
this deceleration is, though the use of elevated viscosity
values may be a significant factor.
In the early stage after the impact, the jet penetrates
into the water and produces a crater that grows in time,
as shown in Fig. 4a for the numerical simulation, and
Fig. 5 for the experimental results. The water originally lying beneath the still water level is pushed upwards, giving rise to two vertical sprays. This behavior
has strong similarities with what occurs during the early
entry stage of a flat plate impacting a quiescent pool, as
discussed in Iafrati and Korobkin (2004). Therein, it is
shown that in the early stage of the impact, gravity effects are significantly smaller than inertial ones which
allows for an approximately self-similar scaling factor

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

of (ttI )2/3 used to collapse the cavity shape. Specifically, the scaling is performed by applying the dimensional transformation
and x02 = x2 (ttI )2/3 (13)

0.05

0
x2

In order to verify if the self-similarity also applies in the


current case, the same free surface profiles are rescaled
using (13) and plotted in Fig. 4b. From this figure it
can be seen that the portions of the five profiles above
the (original) quiescent water level are very well collapsed in the horizontal direction. As to the vertical
direction, the agreement is still satisfactory, but some
effects related to the grid resolution are evident in the
first and fifth free surface profiles, wherein the elevation
of the splash tips are markedly below the others. This
can be explained by considering that, at the beginning,
the upward jets are very thin and the grid is not sufficiently resolved to correctly capture them. For a similar reason, since vertical grid spacing is growing above
x2 = 0.01, the spray portion above this region is numerically eroded, which can be clearly observed in the last
profile on Fig. 4a.
Beneath the free surface, the cavity exhibits a continually widening and deepening crater, as depicted in
both Figs 4a and 5. The rate of penetration and widening of the cavity are qualitatively similar in the lower
depths (a quantitative assessment is provided later in this
section), but several notable differences can be observed
near the original free surface of the pool, largely due to
nominal differences in the jet condition prior to impact.
First, the experimental profiles are noticeably wider at
the surface from the outset of the impact. This is due
to a rounded bolus of fluid at the leading edge of the
jet, approximately twice the diameter of the jet thickness, that is caused by the cutting device used to generate the transient impact. This provided the jet with a
greater initial impulse during the first instant of impact,
and consequently a greater expansion of the near-surface
cavity. For the numerical simulations, the jet could be
truncated in a much cleaner fashion, leaving a jet with a
near-ideal uniform thickness throughout its impact history. The second difference is that the location of the
experimental jet drifts an amount x1 = 0.003 (approximately 1 jet thickness) during the first 50 ms of the
impact. This is caused by a slight vibration transmitted
to the nozzle support during the actuation of the pneumatic jet deflector. The primary effect of this drift is to
slightly alter the appearent trajectory of the cavity penetration, as will also be discussed later.
In a next stage, the dynamics of the cavity shape
changes, and from the sequences shown in Fig. 6, two
main features can be recognized. The first one concerns
the presence of a portion of the cavity where negative
curvature occurs, and an approximate fixed point of the

-0.05

-0.1

-0.15
-0.1

-0.05

0.05

x1

a
0.4

0.2

x2

x01 = x1 (ttI )2/3

-0.2

-0.4
-0.4

-0.2

0
x1

0.2

0.4

Figure 4: a) Interface profiles at times t tI = 0.0075,


0.0175, . . . , 0.0475 are shown. b) the same profiles are
rescaled by the dimensional factor x0i = xi (t tI )2/3 ,
thus recovering a self-similar behavior that seems to
characterize the initial stage of the impact.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

x2

-0.05

0.05

0
x2

-0.1

-0.05

0
x1

-0.05

0.05

-0.1

Figure 5: Interface profiles for experimental conditions


at early times, t tI = 0.01, 0.02, . . . , 0.06 are shown.

-0.15
-0.1

-0.05

0
x1

0.05

x2

cavity. From both the numerical and experimental results presented in Fig. 6, the occurrence of two inflection points located about x2 ' 0.025 is evident that
divides the cavity into two regions: an upper one where
the crater is still enlarging and a lower one where the
cavity is collapsing due to the gravity, eventually leading to the entrapment of an air pocket.
Another peculiarity of this second stage concerns the
existence of a portion of the cavity which rigidly translate with the jet tip. The sequence of the free surface
profiles shown in Fig. 6, reveals that as the jet penetrates the bulk of the fluid, the lowest portion of the cavity appears to translate while keeping its shape roughly
constant. In order to show this fact with greater clarity,
the free surface profiles for both the numerical and experimental tests are plotted on a shifted coordinate system that is fixed to the tip of the penetrating cavity, and
shown in Fig. 7. From this figure it can be easily seen
that the cavity portion below x2 = 0.03 (the overscript
denotes the shifted coordinates) is substantially frozen,
while the upper part of the cavity collapses under the
effect of the gravity.
The occurrence of a second stage during which the
lowest part of the cavity simply translates keeping its
shape can be also recognized by looking at the contours
of the velocity components taken at three different times,
as shown in Fig. 8. Both for u1 and u2 , the contours
nearby the jet tip are essentially frozen, whereas in the
upper part, the effect of the gravity and subsequent cavity collapse is evident.
With the aim of evaluating the propagation velocity

-0.05

-0.1

-0.05

0
x1

0.05

Figure 6: a) Interface profiles at times t tI =


0.0575, 0.0675, . . . , 0.0975 are shown. b) Similar profiles from the experimental conditions for t tI = 0.07,
0.08, . . . , 0.13.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

-0.05

-0.05

x2

0.05

x2

x*2

0.1

-0.1

-0.1

-0.15

-0.15

-0.1

-0.05

0.05

0.1

-0.1

-0.05

x1

0.05

0.1

0.05

0.1

0.05

0.1

x1

0
0
x 1*

-0.05

-0.05

x2

0.05

x2

-0.05

0.1
-0.1

-0.1

-0.15

-0.15

-0.1

-0.05

0.05

0.1

-0.1

-0.05

0.05

-0.05

-0.05

x2

-0.1

-0.1

-0.15

-0.15

-0.1

-0.05

0.05

0.1

x1

-0.05

x1

x2

x2

x1

0
x1

-0.1

-0.05

x1

0.05

Figure 7: a) Interface profiles at times t tI = 0.0575,


0.0675, . . . , 0.0975 are plotted rescaled to place the
leading edge of the crater at position (x1 = 0, x2 = 0).
b) Similar profiles from the experimental conditions for
t tI = 0.03, 0.4, . . . , 0.08.

Figure 8: Contours of the horizontal (left) and vertical


(right) velocity components at t = 0.09, 0.11, 0.13. For
the sake of the clarity, contours are shown only in water
and the red lines represents the density contour at % =
(%a + %w )/2.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

0.01

-0.025
0.005

p1

yp1

-0.05

-0.075

-0.1

-0.005

-0.125
0

0.02

0.04

0.06

0.08

0.1

-0.01
0

0.02

0.04

t-ti

0.06

0.08

0.1

t-ti

Figure 9: Time history of the vertical position of the


lowest interface point. Numerical result is given by the
red line, and the experimental location by the blue symbols. The least squares fit to the later portion of the profile is given by the dashed black line, giving a value of
-1.455 for the simulations and -1.537 for the experimental results.

Figure 10: Time history of the horizontal position of


the lowest interface point. The red line represents the
numerical simulation results, the blue symbols the experimental results, and the dash line, derived by least
squares, shows that the horizontal velocity component is
about 0.1308 for the numerical simulation and 0.227
for the experimental case.

Vp of the jet/pool contact point, the coordinates of the


lowest point along the interface as a function of time is
extracted from the profiles and plotted in Fig. 9. As
can be seen, the tip undergoes an acceleration during the
first 0.015 s before relaxing to a relatively steady velocity corresponding to the constant slope on the x2 vs t diagram. The magnitude of these two velocities is notably
different, with the experiment converging to a value of
up2 = -1.537 and the numerical simulations resulting in
up2 = -1.455. This can most likely be explained by the
difference in their initial jet impact velocity, as the ratio
of up2 /Vj = 0.46 for both cases.
The evaluation of the horizontal component of the
propagation velocity is more difficult. In fact, the horizontal displacement between two successive configurations is smaller than the grid size in the numerical simulations, which is responsible for the saw tooth behavior
of its time history displayed in Fig. 10. There is an even
greater uncertainty associated with the experimental determination of this point, owing to the typically rough
and disturbed surface of the interface. In spite of some
oscillations, the trend appears to be rather well approximated by the interpolating line, which has a slope equal
to 0.1308, with a noticably larger value for the experiment, equal to 0.227. This discrepancy is likely due to
the addition of an apperant horizontal velocity resulting
from the drift of the jet stream as noted above. Con-

sidering the relatively small magnitude of this velocity


component, this is not likely to represent any significant
difference in the overall cavity dynamics.
On the basis of what discussed above, after an initial stage which lasts up to t tI ' 0.02, the jet tip
is shown to follow a rather rectilinear path in the simulation, with an inclination angle of about 5.14 degrees
from the vertical and the modulus of the propagation velocity is about 1.461. The corresponding values for the
experiments are approximately 8.4 degrees and 1.554.
The most interesting aspect of these results is not in
their disagreement (which can be reasonably explained
by the differences in the initial jet shape and vibration),
but rather in their agreement: the ratio of Vp /Vj ' 0.46
for both cases. From simple energy arguments of an inviscid flow (Birkhoff & Zarantonello, 1957; Oguz, et al.,
1992), one would expect that the velocity of penetration
is exactly equal to one half of the impact velocity, or
Vp /Vj = 0.5. For the case of the simulations, one may
consider that the increased viscous effects are responsible for the decrease, but it seems an unusual conincidence that the experimental results give the same value.
Clearly additional measurements and simulations under
varied conditions are needed to resolve this issue.
In order to observe the steady flow pattern generated about the moving jet tip, the two components of
the propagation velocity are subtracted from the veloc-

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

time of t tI = 0.117 0.004, the values are remarkably close given that a few additional time-steps are required in the simulation to complete the pinch-off process. The dimension of the cavity and of the air pocket
that would be entrapped, which are of primary interest
for the present work, are correctly predicted as well.

-0.125
-0.13
-0.135

x2

-0.14

CONCLUDING REMARKS

-0.145
-0.15
-0.155
-0.16
-0.03

-0.02

-0.01

0.01

x1

Figure 11: Streamtraces about the jet tip. In order to


get the velocity field in a frame of reference translating
with the cavity, the penetration velocity of the jet Vp =
(0.1308, 1.455) has been subtracted.
ity field and the related streamtraces are plotted in Fig.
11. According to what was found experimentally, a stagnation point exists near the contact region. Furthermore,
the streamtraces coming from inside the jet impinge on
the stagnation region and divide, thus forming a thin
layer which remains attached to the cavity.
Results presented so far are in a relatively good agreement with the experiments, both qualitatively and quantitatively, when one accounts for the slight difference
in initial conditions. With the aim of getting a more
complete physical comparison of the interface behavior,
the free surface profiles obtained numerically are superimposed onto the corresponding frames of the recorded
movie from the experiments (see Fig. 12).
The comparison is started from the time of initial
impact and the scale factor is assigned in a preliminary
stage, by using a shapshot of a submerged ruler to fix the
dimensions of the numerical window. The time step between successive plots is 0.01 but for the last one which
is 0.006. The reason is that at t tI = 0.106, that is just
before pinch-off, the numerical calculation stops due to
difficulties to accurately describe the dynamics of the
flow within the thin region lying between the sprays and
the jet surface at this stage. In spite of this limitation, the
numerical jet is rather well overlapped to the experimental one showing a rather good agreement in both space
and time. The only major exception is the decreased
depth of penetration, which again, primarily stems from
the difference in intial impact velocity magnitude. In
comparison to the experiments, which have a pinch-off

With the aim of understanding the process of air entrainment in breaking waves, the plunging of a planar
water jet onto a still water surface has been studied, carrying out experiments and simulations. The development of the cavity induced by the jet impact has been
observed by high-speed digital camera and the corresponding velocity field measured by particle image velocimetry techniques. The flow has been also numerically studied via a Navier-Stokes solver coupled with a
two-fluids level-set approach for the interface capturing.
Two distinct phases are clearly recognizable.
The penetration of the jet into the quiescent water
during the early stage after the impact is responsible for
the formation of a crater that grows in time and for the
generation of two vertical sprays, in a way much similar
to the water entry of a flat plate. For that problem, the
solution is found approximately self-similar with a scale
factor of t2/3 and the scaling of the free surface profiles
agrees with this finding.
During the second phase, the cavity deepens in the
bulk of water, translating under the continuous action
of the impinging jet. However, the shape of the cavity
remains roughly constant (frozen), as observed by properly shifting the free surface profiles with the position
of the jet tip. Furthermore, by superimposing these profiles in the fixed frame of reference, two inflexion points
are easily identified, and clearly separate an upper region, close to the free surface, where the crater is continuously widening, and a lower region where the cavity
is closing, collapsing under the action of the gravity. In
this later stage an air pocket is entrapped, which successively breaks into a cloud of tiny bubbles.
At the present stage, the numerical model is not able
to reach the pinch-off. Although comparisons with the
experiments looks rather encouraging, a additional effort has to be pursued to overcome the pinch-off stage
and, also, include the translation velocity. This would
allow a larger amount of data, which is necessary to establish a parametric study.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Figure 12: Time sequence of the jet impact without translation velocity, for a angle of inclination of 7 degree from the
vertical, jet speed 3.32 and jet thickness at impact of 0.0029. The time step between successive plots is 0.01 but for
the last one which is 0.006. The red line represents the mean density contour % = (%a + %w )/2.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

REFERENCES
Birkhoff, G., and Zarantonello, E.H., Jets, Wakes, and
Cavities, Academic Press, Inc., New York, 1957.
Brackbill, J.U., Kothe, D.B., and Zemach, C., A continuum method for modeling surface tension, Journal
Computational Physics, Vol. 100, 1992, p. 335.
Brandt, A., Guide to multigrid development, Multigrid
Methods, Hackbush W., Trottenberg U. Eds., SpringerVerlag, Berlin, Germany, 1992.
Chorin, A.J., A numerical method for solving incompressible viscous flow problems, Journal Computational
Physics, Vol. 2, 1967, p. 12.
Chirichella D., Gomez Ledesma R., Kiger K.T., Duncan
J.H., Incipient air entrainment in a translating axisymmetric plunging laminar jet, Physics of Fluids, Vol. 14,
2002, p. 781.
Gomez Ledesma, R., An experimental investigation on
the air entrainment of translating plunging jets, Ph.D.
Thesis, University of Maryland, 2004.
Iafrati, A., Di Mascio, A., and Campana, E.F., A levelset technique applied to unsteady free surface flows,
International Journal for Numerical Methods in Fluids,
Vol. 35, 2001, p. 281.
Iafrati, A., and Campana, E.F., A domain decomposition approach to compute wave breaking, International
Journal for Numerical Methods in Fluids, Vol. 41, 2003,
p. 419.
Iafrati, A., Korobkin, A.A., Initial stage of flat plate
impact onto liquid free surface, Physics of Fluids, in
press, 2004.
Kim, J., and Moin, P., Application of a fractional-step
method to incompressible Navier-Stokes equations, Journal of Computational Physics, Vol. 59, 1985, p. 308.
Kolaini, A.R., Crum A.A., Observation of underwater
sound from laboratory breaking waves and the implications concerning ambient noise in the ocean, Journal
Acoustical Society of America, Vol. 96, 1994, p. 1755.
Lara, P., Onset of air entrainment for a water jet impinging vertically on a water surface, Chemical Engineering Science, Vol. 34, 1979, p. 1164.
Oguz, H.N., Lezzi, A.M., and Prosperetti, A., Examples of air entraining flows, Phys. Fluids, Vol. 4, 1992,
p. 649.
Rai, M.M., and Moin, P., Direct simulations of turbulent flow using finite-difference schemes, Journal of
Computational Physics, Vol. 96, 1991, p. 15.
Rosenfeld, M., Kwak, D., and Vinokur, M., A fractional step solution method for the unsteady Incompressible Navier-Stokes equations in generalized coordinate
system, Journal of Computational Physics, Vol. 94, 1991,
p. 102.
Sheridan, A.T., Surface entrainment of air by a water
jet, Nature, Vol. 209, 1966, p. 799.

Sussman, M., Smereka, P., and Osher, S., A level set


approach for computing solutions to incompressible twophase flow, Journal of Computational Physics, Vol. 114,
1994, p. 146.
Zang, Y., Street, R.L., and Koseff, J.R., A non-staggered
grid, fractional step method for time-dependent incompressible Navier-Stokes equations in curvilinear coordinates, Journal of Computational Physics, Vol. 114,
1994, p. 18.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Experimental Investigation of a Cavitating Propeller in


Non-Uniform Inflow
Francisco Pereira,
Fabio Di Felice, Massimo Soave
INSEAN - Propulsion and Cavitation Group
CEIMM - Italian Navy Cavitation Tunnel
Rome, Italy
ABSTRACT

Upstream axial inflow velocity

Pressure and noise measurements are performed around a


cavitating propeller in non-uniform flowfield. High speed
visualizations are used to quantify the cavitation extension
across the wake. Harmonic analysis is performed that allows
the identification of the different pressure and noise sources
and their respective contribution to the induced fluctuations.
An attempt is made to develop a simple model to predict the
pressure fluctuations due to the occurence of cavitation, based
on the cavity volume acceleration.

Vc

Cavity volume

BPF

Blade passage frequency

Fluid density

Cavitation number referred to U , p0

NOMENCLATURE
t

Sampling frequency

Angular position

An , n Amplitude and phase of the Fourier transform


Cn , Sn Coefficients of the Fourier transform
D

Propeller diameter, 2R

E0

Blade face area for r R

Ec

Cavity extension

f0

Fundamental frequency (=BPF)

Advance coefficient, U nD

KT

Thrust coefficient, T n2D4

lc

Cavity characteristic length

Propeller angular velocity (rps), normal

p0

Pressure at propeller axis

pv

Vapor pressure

Propeller radius

Thrust

03


pv


1
2
2 U

INTRODUCTION
Cavitation occurrence on marine propellers is the source of
undesirable effects: radiated noise, structural vibrations and,
though of a lesser relevance in the field of marine propulsion,
erosion and loss of efficiency. In order to minimize these effects, the actual effort is put into the development of analysis
tools capable of predicting, in particular at the design stage,
the cavitation pattern on a propeller geometry. However, the
definition of these tools is dependent upon the experimental validation of current theoretical models for the analysis
of cavitating flows. The present research paper is a direct
follow-up of the work presented at the previous ONR symposium (Pereira et al., 2002), where the cavitation pattern on
a propeller in uniform inflow was analysed experimentally
using a novel approach based on the crosscorrelation of highresolution cavitation images. These measurements were purposely performed to validate a cavitating flow model based
on an inviscid flow boundary element method for the analysis
of blade partial sheet cavitation and supercavitation.
Following the successful outcome of this initial work, the
aim of the work described here is to present the results of
an experimental investigation of a cavitating marine propeller
under non-uniform inflow conditions. The primary objective
is to provide a more comprehensive insight into the correlations between the cavitation pattern and the radiated pressure
fluctuations, as experienced in a typical hull-propeller configuration. Another important intent is to complete the current
experimental database and provide guidelines to validate new
modeling features. In a first place, we seek to validate the

Copyright National Academy of1Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

current theoretical and numerical approach in non uniform inflow conditions, accounting for effects such as viscosity and
tip vorticity. Also, efforts are ongoing to address the hydroacoustics of this type of flows, and our experimental work is
designed to help understand and develop this important aspect.
Marine screw propellers operate in a flow perturbed by
the presence of the hull, which generally causes the upstream
flow to decelerate in a limited region of the hull wake where
the levels of turbulence are also significantly increased. The
non-uniformity of the incoming flow into the propeller disc
causes a periodic variation of the blade loading and of the hydrostatic pressure. Apart from the varying thrust developed
by the propeller, this situation creates the necessary set of
conditions for the occurrence of unsteady cavitation on the
propeller blades. Unsteady cavitation generates strong pressure fluctuations radiating in the far field or to the submerged
part of the hull structure through the water medium. The
broadband frequency content of the noise generated by cavitation makes it the major source of noise and vibrations on
a ship, hence contributing to crew discomfort on a passenger vessel, interfering with scientific instrumentation on a research ship or promoting the detectability of a naval warship.
Our purpose is to investigate the spatial and temporal correlations between the cavitation pattern and the radiated noise
of a cavitating propeller in a non-uniform flowfield. A rich
litterature is available on the subject. Ito (1966) investigated
the effect of the variation of the angle of attack of the flow
on the unsteady cavitation, using an oscillating blade and extrapolating the results to the propeller behind a hull. Increase
in pressure fluctuations due to propeller cavitation was probably first reported by Takahashi and Ueda (1969). Bark and
Berlekom (1978) have looked into the relations between cavity dynamics, type of cavitation and cavitation noise. The
study was performed in a cavitation tunnel with an oscillating hydrofoil. In this work, the periodic growth of the cavity at the leading edge was found to increase the noise levels in the low frequency range, in particular at multiples of
the oscillating blade frequency, whereas tip vortex cavitation
and bubble cavitation generate high frequency noise. Matusiak (1992) proposed a method to evaluate the noise contributions of the individual types of cavitation, respectively
low-frequency pressure fluctuations from fixed blade cavitation and high-frequency broadband noise due to collapse of
vapor bubbles. Several works present attempts to correlate
the noise and cavitation results obtained in a cavitation tunnel to those at full scale (Johnsson et al., 1976; Chiba et al.,
1980). Breslin et al. (1982) presented an important theoretical and experimental research work directed at the prediction of vibratory pressures, forces and moments induced by
intermittently cavitating propellers. Scale effects have been
addressed for instance by Friesch et al. (1992), who also considered pressure fluctuations at orders higher than twice or
thrice the blade frequency (Friesch, 1998). More recently and
with the help of high-speed video imaging, Johannsen (1998)
tried to derive distinct scaling laws based on fluid mechanical connections between cavity dynamics, hull pressure time

series and inflow velocity field.


In the present work, detailed temporal and spatial correlations are established between quantitative measurements of
the cavitation pattern and the pressure and noise data measured at different locations in the vicinity of a propeller. The
propeller used in this work, the E779A, has been extensively
used in previous research works, the results of which are compiled in a public database. The E779A propeller provides a
complete and challenging test case for the validation of numerical codes. The current experiments are performed in a
cavitation tunnel for a wide range of non-cavitating and cavitating flow conditions. The effect of the hull on the wake
is simulated with an array of plates, which generate a nonuniform inflow to the propeller. The cavitation pattern measurements are established on the basis of high-speed visualizations analysed using the novel correlation-based algorithms presented in the first part of the project. The pressure and noise measurements are performed simultaneously
at very high frequency to provide an accurate time history
for comparison with the time-resolved cavitation pattern data.
Noise signals are analysed both in the time and spatial domains.
This work completes the existing database to unsteady
cavitating operation of propellers.
EXPERIMENTAL SETUP

Facility and Propeller


The experiments are carried out at the Italian Navy cavitation
tunnel facility (C.E.I.M.M.). The tunnel is a closed water circuit with a 0 6 m 0 6 m 2 6 m square test section. Optical
access to the section is possible through large Perspex windows. The nozzle contraction ratio is 5 96 : 1 and the maximum water speed is 12 m s 1 . The maximum free stream
turbulence intensity in the test section is 2%. The flow uniformity is within 1% for the axial component and 3% for the
vertical one.
The propeller used in the present tests is a modified Wageningen type model propeller, referred to as the INSEAN
E779A propeller, used in previous experiments (Pereira et al.,
2002). This skewed four-blade model propeller has a radius
R D 2 0 117 m, a pitch-to-diameter ratio of 1.1 and a forward rake angle of 43. The blockage ratio in the test section
is about 10%.
The measurement configuration is pictured in Fig. 1. The
propeller shaft is equipped with a rotary encoder that supplies
an electrical trigger signal, used to pilot a pulse delay generator, which in turn drives the image and signal acquisition.
This rotation impulse is also recorded to time mark the pressure and noise signals, in particular for the purpose of analysis
during the propeller rotation period. This time marker determines the start time for the time signals, the image high-speed
sequence and defines the 0 for the propeller angular position
.
In the following sections, J and KT are the advance and the
thrust coefficients, respectively. 0 is the cavitation number


Copyright National Academy of2Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Synchronization

 
 
 
 
 P2
 
 
 

pulse
















 
 
 
 
 
 
 















H1
 
 
 
 
 
 
 
  Flow



H2




















 

 
H3
 
 























Pressure &
H4
 
 




















+
P1
























P3
noise signals
P2
 
 
 
 
 
 
 
 
P1
P3
















P4
 
 
 
 
 
 
 
 








 H1




 
 
 
 
 
 
 
  H2


Images
H3
















H4




































P4



               
Light








        








Wake generator

Rotation

Trigger

Signal conditioning
& A/D conversion

180

Highspeed

(Trigger)

camera

270

Water
prism

H : hydrophone

P : pressure transducer










Control & acquisition

Figure 1: Experimental setup.


related to the reference pressure p0 measured at the propeller
axis.

Wake Generator
The non-uniformity of the wake downstream of a hull is
here simulated using an array of plates, following closely
the design described in the recommendations notes of the
International Towing Tank Conference on measurements of
hull pressure fluctuations (Huse, 1996). The wake simulator,
shown in Fig. 2, is composed of five plates spaced 20 mm
apart from each other and assembled together to form an array 90 mm thick and 300 mm long. The rake is fixed to the
ceiling of the test section, upstream of the propeller, as shown
in the photograph.

High-speed Visualizations
A high-speed camera is used to record the cavitation pattern
at a frame rate of 2000 frames per second with a resolution
of 1024 1024 pixels. Hence and with the rotation speed
used in our experiments ( 30 rps), the angular resolution is
about 5 5. The camera is oriented at an angle with respect
to the test section window, as depicted in Fig. 1. Looking at
an angle through the thick window introduces strong aberrations, which are almost fully cancelled using a water filled
glass tank in the form of a wedge, and placed so that the camera optical axis is normal to the wedge face. The illumination
consists in a set of high power flood lights to accomodate the
short exposure time used to record the cavitation on the rotating blades. A typical shutter time of 1 10000th to 1 20000th
of a frame is used to avoid image blurring due to motion. The
method described by Pereira et al. (2002), based on the cross-

Figure 2: Non-uniform wake generator.

Copyright National Academy of3Sciences. All rights reserved.

To highspeed camera

90

Twenty-Fifth Symposium on Naval Hydrodynamics

correlation operation between a non-cavitating and a cavitating image and on the non-linear dewarping of those, is used
to provide an accurate estimate of the mean and fluctuating
attached cavity extension.

The pressure measurements are performed with four Endevco 8510C-15 type piezo-resistive pressure transducers, flush
mounted to the test section walls and on the propeller plane,
as indicated on Fig. 1. These pressure transducers are designed by P1, P2, P3 and P4, and are respectively placed at
0, 90, 180 and 270. Hence, P1 is 90 behind the wake
simulator and P2 is above the propeller.
Four Brel&Kjaer 8103 type hydrophones fixed to a
streamlined strut are used for the noise measurements in the
fluid, and are referred to as H1 , H2 , H3 and H4 , see Fig. 1.
The sensors are located in a radial plane at a distance of
about one radius downstream of the propeller plane, and are
at 80 mm 0 7R (H1 ), 100 mm R (H2 ), 120 mm (H3 ) and
200 mm (H4 ) from the propeller axis, as illustrated in Fig. 3.
Hence H1 and H3 are respectively inside and outside the slipstream, while H2 is roughly located at the tip vortex position,
and H4 is far from the flow perturbation created by the propeller. H4 is thus essentially used for background noise considerations.

The experiments are performed for a propeller angular speed


n 30. The advance coefficient is adjusted to meet the thrust
identity between the isolated propeller value and that behind
the wake simulator. The thrust coefficient is set to 0 175,
which corresponds to J 0 89 for a flow upstream velocity of
6 22 m s 1 . The values of the cavitation coefficient considered here are: 7.5, 6.5, 5.5, 4.5, 4.0, 3.5, 3.0, 2.5, 2.0 and 1.5.
The cavity extension has been established for 0 equal to 7.5,
6.5, 5.5, 4.5, 3.5 and 2.5. The oxygen content is monitored
and is maintained below 5 ppm.

Pressure and Noise Measurements

Flow conditions

HIGH SPEED VISUALIZATION


Figure 4 illustrates the cavitation pattern for a value of the
cavitation coefficient equal to 3 5, as the blade passes through
the turbulent wake generated by the hull simulator. Such sequences are used to help interpret the pressure and noise signals discussed hereafter. They are also used to carry out quantitative information about the cavity pattern.
The measurement of the cavity extension is performed on
every frame. The recording sequence lasts about 3 seconds,
during which the propeller performs approximately 90 revolutions. Because the camera is triggered once on one single
rotation impulse, the mean and fluctuating values of the cavity
extension are calculated within slots centered at discrete angles. In our experiment, the angular step is set to 5 starting
at 60, hence the statistics are calculated inside angular
slots of 2 5 around the being considered.


Generator
Flow

FREQUENCY DOMAIN ANALYSIS OF FLUCTUATING PRESSURE AND NOISE

H1

Pressure and noise signals are sampled N times per propeller


revolution with a time-spacing t corresponding to the sampling frequency (10 5 s). A waveform is then constructed by
ensemble averaging with the following relation:

H2
H3

H4
P4

Pn


1 M
Pm n
M m1

(1)

Figure 3: Hydrophone strut for noise measurement.


The pressure transducers have been previously calibrated
statically in pressure and tension. The discrepancy between
the measured and the actual value is less than 0 5% across
the full pressure range. A dynamic calibration has also been
performed for the hydrophones to verify the manufacturers
calibration sheet. These hydrophones provide a measurement
uncertainty of 1 5 dB re Pa in the range 10 Hz 80 kHz
and a frequency response flat in the range 0 1 Hz 20 kHz
with a tolerance of 1 5 to 1 dB re Pa.
The pressure and noise signals are sampled at a rate of
100 kHz per channel and recorded for 30 seconds, during
which the propeller performs more than 900 revolutions.
Hence, we have a sample of all eight signals approximately
every 0 11 of propeller rotation.


where m and n are the corresponding mth revolution and the


nth data point in that revolution.
Due to the ergodicity of the random process involved, the
average pressure-time history obtained from Eq. (1) can be
transposed into the frequency domain using the direct Fourier
transform. Expressed in mathematical terms, Fouriers theorem asserts that, for a periodic function g t of fundamental
frequency f 0 , we can write g as a sum of periodic functions
such that


gt

An cos 2n f0t


n


(2)

n 0

where An and n are the amplitude and phase of the corresponding harmonic function. Equation (2) can be written as

gt


Cn cos 2n f0t

n 0


Copyright National Academy of4Sciences. All rights reserved.

Sn sin 2n f0t

n 0


(3)

Twenty-Fifth Symposium on Naval Hydrodynamics

10

19

11

20

12

21

13

22

14

23

15

24

16

25

17

26

18

27

Figure 4: High speed sequence of the cavitating propeller for 0




3 5. Interframe time= 500 s.




Copyright National Academy of5Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

where the determination of An and n is now reduced to the


determination of the coefficients Cn and Sn , which are calculated using the following formulas:
f0 2


Cm

2 f0


g t cos 2m f0t dt


m


f0 2

0


f0 2


C0

f0


(4)

g t dt


f0 2


f0 2


Sm


2 f0

g t sin 2m f0t dt


f0 2


This transform allows the decomposition of a signal into


its periodic components, i.e. the fundamental frequency or
first harmonic, and its multiples or higher order harmonics.
The fundamental is here represented by the blade passage frequency (BPF), which is equal to the rotation speed times the
number of blades. In our case, the propeller is rotating at
30 rps and the number of blades is 4, hence BPF is equal
to 120 Hz.
Such a spectral analysis of a periodically unsteady flowfield sorts out the phase-locked coherent flow structures from
an otherwise random unsteadiness in the flowfield. Figure 5
and Fig. 6 represent the time history of the pressure signal
measured by the hydrophone H2 and by the pressure transducer P2 , respectively. The average is calculated by using
Eq. (1) over the more than 900 propeller revolutions performed during the 30s of measurement. For the polar representation in these figures, the pressure is normalized to its
minimum and maximum values in order to allow the comparison between the different flow conditions, hence the levels
are not indicated. The true levels are reported in the harmonic plots by the amplitude coefficients An . The signal is
represented in polar form, with the 0 representing the time
start, as indicated in Fig. 1 and 90 being the mean location
of the flow perturbation created by the wake generator. Each
polar graph is accompanied by a plot of the amplitude An versus the order number, where An is equal to the norm of the
vector Cn Sn corresponding to the nth harmonic.
The two transducers display two very distinct behaviors.
H2 is located in the fluid, one radius from the propeller plane.
P2 is located above the propeller, close to the blade tip. H2
is particularly sensitive to the passage of the tip vortex that
is generated by the blades, for H2 is placed at
0 7R at the
location swept by the vortex. The first and second harmonics
are preponderant, although higher harmonics are activated as
the vortex is progressively transformed into vapor as the cavitation coefficient 0 decreases.
The pressure measured at the P2 location provides a
greater insight into the flow phenomenology, for the transducer is placed close to the blade tip. The polar plots show a
pressure pattern that is strongly modulated by the progressive
excitation of different sets of orders, as the cavitation coefficient 0 is decreased. Although the pressure pattern remains
almost identical from 0 7 5 to 0 5 5, it is noticeable
that higher orders become progressively larger in amplitude
than the first and second harmonics. This is in particular visible for 0 comprised between 3 5 and 2 0. The pressure pat-

tern displays sharp peaks that are clearly phase-locked to the


growth of the sheet cavity over the blades, as illustrated by
the frames 10 to 15 of Fig. 4, to the collapse of this cavity
(frames 16 to 19), to the rebound and collapse of the vapor
structures released from the first implosion and to the growth
and collapse of the tip vortex in the interblade region (frames
20 to 24). The sequence of events described by the pressure
patterns suggests that the pulses from the second series of collapses are stronger than those from the growth and collpase
of the main cavity. Indeed, the high speed sequence seems to
point out to a smooth wipe out of the attached cavity rather
than to a violent collapse, see frames 16 to 18. Instead the
vapor structure that is shed away into the fluid by this process
seems to interact strongly with the tip vortex, feeding it with
vapor matter that contributes to a violent collapse in the fluid.
This phenomenon seems to take place when the approaching
blade is at about 60, as shown by the sharp peak visible on
the polar graph at 0 2 5. Hence, this peak is the result of
the vortex collapse issued from the preceding blade.
As the cavitation coefficient is further decrease, this collapse seems to take place with a greater delay, thus explaining
the progressive phase change observed on the polar plots from
0 3 0 to 0 2 0. This phenomenon might be explained
by an ever smoother pressure gradient between the cavity inside pressure and the surrounding pressure, resulting in a cavity volume with a dampened response to that gradient. During
this particular working range, the orders 3 to 5 are the most
important. At 0 1 5, supercavitation is preponderant and
the main pressure pulse is that produced by the cavity growth
and collapse. The collapse of the vortex from the preceding
blade is not taking place anymore, this explaining the zerophase locking, i.e.the pressure pulse occurs exactly when the
blade is vertically aligned with the turbulent wake.
The occurrence of sheet cavitation, of tip vortex and of
hub vortex induces significant pressure fluctuations at blade
harmonics up to the 12th order. To provide a synthetic view
of the pressure and noise pattern, we define the term Kn as
the ratio between the Fourier coefficient An in a given cavitating situation and the value of this coefficient for 0 7 5,
where the cavitation pattern is minimal. Hence we have
Kn An 0 An 0 7 5 .
This ratio is represented in Fig. 7 for every sensor and all
conditions. For clarity, we consider only the first six BPF
harmonics.
The graphs clearly outline the dependence of the harmonic levels upon the cavitation coefficient 0 . The curves
with the plain symbols are those that present a distinct behavior relative to the others, for a given cavitation condition. It is
important to note that the fourth harmonic is always activated
by the cavitation, except for the transducers H1 and H2 . As
a matter of fact, those two hydrophones are inside the slipstream created by the propeller and witness in first place the
blade wake passage. This wake has it source at the blade trailing edge, where the interaction between the boundary layers
from the blade faces generates a turbulent wake which propagates inside the propeller slipstream. The pressure fluctuations induced by the blade wakes contribute mainly through


Copyright National Academy of6Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

400
800

400
1600 0
800

1600 0

0 = 1.5

800

An [Pa]

800

6
8
4
10
Blade harmonic

12

400

400

1600 0

800

0 = 4.0

90

1200

An [Pa]

90

1600 0

400

400

0 = 2.0

An [Pa]

800

0 = 4.5

90

1200

An [Pa]

90

1600 0

400

400

1200

0 = 2.5

An [Pa]

800

0 = 5.5

90

1200

An [Pa]

90

1600 0

400

1200

1600 0

0 = 3.0

1200

800

0 = 6.5

90

An [Pa]

1600 0

400

1200

An [Pa]

90

1200
800

An [Pa]

0 = 3.5

1200

1600

800

0 = 7.5

90

1200

1600

An [Pa]

90

6
8
4
10
Blade harmonic

12

Figure 5: Time history of the noise signal from hydrophone H2 and corresponding harmonic decomposition,
for different values of the cavitation coefficient 0 . The polar graphs are non-dimensionalized.

Copyright National Academy of7Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics


2000
o

2000 0
1000

2000 0
o

2000 0

0 = 2.0

An [Pa]

0 = 1.5

1000

6
8
4
10
Blade harmonic

12

2000 0
1000
0

100

200

500

400

0 = 4.0

90

300

An [Pa]

90

500 0

100

500

200

An [Pa]

0 = 4.5

90

400

500 0

100

500

200

300

An [Pa]

90

1500

0 = 2.5

1000

An [Pa]

0 = 5.5

90

400

500 0

100

500

200

300

An [Pa]

90

1500

0 = 3.0

1500

An [Pa]

0 = 6.5

90

400

500 0

100

500

200

300

An [Pa]

90

1500
1000

An [Pa]

0 = 3.5

1500

300

0 = 7.5

90

400

500

An [Pa]

90

6
8
4
10
Blade harmonic

12

Figure 6: Time history of the pressure signal from pressure transducer P2 and corresponding harmonic decomposition, for different values of the cavitation coefficient 0 . The polar graphs are non-dimensionalized.

Copyright National Academy of8Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

50

50

Kn [-]

1 x BPF
2 x BPF
3 x BPF
4 x BPF
5 x BPF
6 x BPF

H1

40
30

40

20

10

10

0
100

0
200
1 x BPF
2 x BPF
3 x BPF
4 x BPF
5 x BPF
6 x BPF

H3

Kn [-]

60

160

80

20

40

0
100

0
200

Kn [-]

60

160

80

20

40

0
50

0
100
1 x BPF
2 x BPF
3 x BPF
4 x BPF
5 x BPF
6 x BPF

P3

30

80

40

10

20

0 []

P2

1 x BPF
2 x BPF
3 x BPF
4 x BPF
5 x BPF
6 x BPF

P4

1 x BPF
2 x BPF
3 x BPF
4 x BPF
5 x BPF
6 x BPF

60

20

1 x BPF
2 x BPF
3 x BPF
4 x BPF
5 x BPF
6 x BPF

120

40

40

Kn [-]

1 x BPF
2 x BPF
3 x BPF
4 x BPF
5 x BPF
6 x BPF

P1

H4

120

40

80

1 x BPF
2 x BPF
3 x BPF
4 x BPF
5 x BPF
6 x BPF

30

20

80

H2

0 []

Figure 7: Ratio Kn for the first 6 harmonics. H1 , H2 , H3 and H4 are the hydrophones; P1 , P2 , P3 and P4 are the
pressure transducers. The plain symbols indicate the most relevant cases.

Copyright National Academy of9Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

CAVITATION EXTENSION
Figure 8 represents the evolution of the cavity extension Ec as
a function of the blade angular position . The cavity area is
reresented in terms of a non-dimensional value, where the actual area has been non-dimensionalized by the blade face area
E0 for r R 0 3. It is seen that the cavitation inception takes
place before 60. The cavity extension grows almost linearly
till a maximum that is located slightly after the wake region.
This maximum seems to deviate further in the propeller rotation direction as the cavitation coefficient is decreased. This
is consistent with our previous analysis regarding the shifting observed on the pressure pulses in Figs. 5 and 6: with
decreasing 0 , the pressure gradient is smoother across the
simulator wake, the cavity tends to further extend its growth
in time and space and the collapse occurs with larger delay.


lapse of the main cavity immediately followed by that of the


tip vortex. The second maximum is probably related to the
sudden pressure gradient at the wake boundary.

6%

0 = 7.5

4%
2%
0%

6%
Cavity area fluctuations, rms(Ec)/E0

the 3rd and 5th harmonics. Yet the 3rd harmonic seems to be
connected to the tip vortex collapse, as its contribution is not
noticeable for H2 where the major contribution is given by the
tip vortex/blade wake sweeping.
Outside the slipstream, the fourth harmonic is omnipresent and is linked to the attached cavity growth and collapse. The strong presence of the 3rd harmonic on P2 and to
a lesser extent on P1 indicates that this fluctuation mode is related to the tip vortex collapse as we pointed out before. P3 ,
which is located opposite to P1 , witnesses the attached cavity
influence through the fourth harmonic but senses also a strong
first harmonic. The pressure fluctuation at the blade passage
frequency is thought to be due to the fluid recompression induced by the blade exit from the wake toward the region of
fluid where the sensor P2 is located.

0 = 6.5

4%
2%
0%

6%

0 = 5.5

4%
2%
0%

6%

0 = 4.5

4%
2%
0%

6%

0 = 3.5

4%
2%

60%
0 = 7.5
0 = 6.5
0 = 5.5
0 = 4.5
0 = 3.5
0 = 2.5

50%

Ec/E0

40%
30%

0 = 2.5

4%
2%
0%
40

20%

60

80

100

120

140

Angular position [degrees]

10%
0%
40

0%

6%

50

60

70
80
90 100 110
Angular position [degrees]

120

130

140

Figure 9: Fluctuations of the cavity extension as a


function of the propeller angle .

Figure 8: Cavity extension and corresponding fluctuations as a function of the propeller angle .
CORRELATIONS
Figure 9 shows the evolution of the cavity extension fluctuations, which display a maximum after the vertical position
at 105. This maximum moves to 110 when supercavitation takes place. A second maximum is visible at
80. The first maximum is clearly assignable to the col

Figure 10 displays the joint time history comparison between


the cavity mean extension and the pressure mean values. It
is remarkable that the pressure maximum is not, as often reported, at the cavity maximum extension. The pressure peak

Copyright National Academy of10


Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

90

60

Mean

120

120

90

60

90

Mean

120

90

60

60

120

90

0 = 2.5
60

RMS

90

60

120

90

0 = 5.5

60

120

90

60

0 = 4.5

120

0 = 6.5
o

60

60

0 = 7.5
90

90

0 = 3.5

120

RMS

120

120

0 = 4.5
o

90

60

0 = 5.5

0 = 6.5
60

90

0 = 7.5
o

120

120

0 = 3.5

0 = 2.5

Figure 10: Pressure and cavity extension correlations: mean (top two rows); fluctuations (bottom two rows).
Area is represented by the dashed curve.

Copyright National Academy of11


Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

occurs at the collapse time and is related to the cavitation


event on the preceding blade. Also, it is clear from this graph
that the cavity growth and collapse is concentrated in a rather
narrow angular band, from 60 to 120. The location of the
cavity maximum extension is shifting in the direction of the
rotation. This phenomenon is also observed on the pressure
pulses.
With regard to the fluctuations of the cavity extension and
pressure, it is also clear that the strongest levels are measured
when the blade is about 30 away from the vertical position.
With decreasing 0 , this angle tends to decrease.
We propose here to use the model of the pressure
increment generated by an expanding spherical cavity,
see e.g. Breslin et al. (1982). It is demonstrated that the incremental pressure at a given distance from this expanding cavity
is proportional to the acceleration of the cavity volume:
p Vc

with lc


Ec

6 lc

dlc
dt

-10
10
0 = 6.5
0

-10
15
0 = 5.5

2


3 lc2

p [Pa]

-15
40
0 = 4.5
0

-40
60
0 = 3.5
0

(6)
-60
200

We show then that Eq. (5) can be written


d 2Vc
dt 2

0 = 7.5

(5)

where Vc is the cavity volume. This relation considers


a spherically symmetric, irrotational, incompressible, radial
flow emanating from a single point in a boundless fluid. On
propeller blades, the cavities are obvioulsy far from spherical. Therefore we make the hypothesis that their volume is
proportional to a characteristic length lc . Previous works by
Pereira et al. (1998) showed that on bidimensional hydrofoils,
the height of the leading edge cavity was linearly linked to its
length. These authors also showed experimentally that the vapor structures generated by an unsteady sheet cavitation could
be represented by a characteristic length. More recently, experimental observations and calculations of cavitating propeller flows performed by Pereira et al. (2002) brought new
evidence that the cavity thickness on a propeller blade had a
similar behavior. We therefore make the hypothesis that the
cavity volume Vc is proportional to a length lc such that
Vc lc3

10

d 2 lc
dt 2

0 = 2.5

(7)

Using our experimental data, we derive the volume acceleration and compare it with the measured pressure. We
choose the pressure sensor P2 , for it is located the closest
to the place where cavitation occurs. In order to cancel out
effects not directly related to the attached cavitation, we regenerate the pressure time signal using a limited numbers of
orders. Specifically, we use the first twelve harmonics to reconstruct the pressure signal, see Fig. 6. The cavity extension
data is fitted by a polynomial which is then used to calculate the volume acceleration with Eq. (7). The estimated and
the experimental pressure incremental due to the cavity volume growth are finally directly overlaid to produce the graphs
shown in Fig. 11.
A remarkable agreement is found between both in the time
window covered by the cavity extension measurements. The
estimated pressure variation derived from the cavity volume

-200
0.006

0.009

0.012

0.015

0.018

t [s]
Figure 11: Comparison between p measured ( )
and p estimated from volume acceleration ( ),
Eq. (7).

acceleration accurately matches the pattern of the measured


p. Some discrepancy appears for the lower values of the
cavitation coefficient, which may be due to the number of
orders (12) used to reconstruct the pressure signal. A more

Copyright National Academy of12


Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

restricted set of harmonics may improve the match. Yet, most


of the features are determined. In particular, the time step between the pressure peaks are very well reproduced. Hence,
the simple model based on the hypothesis of a linear relationship between the cavity dimensions is found to perform
surprisingly well.
CONCLUSION
An experimental study of a cavitating propeller operating in
a non-uniform flowfield has been presented. A wake simulator was used to create a local flow perturbation upstream
of the propeller. Using four hydrophones placed in the fluid
and four wall-mounted pressure transducers in the propeller
plane, a detailed harmonic analysis has put into evidence the
contributions of flow features such as the leading edge cavity,
the blade tip vortex and the blade wake. The leading edge
cavity seems to contribute mainly to the fourth order, while
the tip vortex collapse acts in particular on the third harmonics. The blade wake plays a role at harmonic orders as high
as the fifth. With the help of spatially resolved measurements
of the cavitation extension on the blade, the mean and fluctuations of the leading edge cavitation have been established.
We formulated the hypothesis that the cavity dimensions
are linearly linked to a characteristic length, and we derived
the cavity acceleration and compared it with the measured
pressure variations, using a limited range of harmonics of interest. A good agreement was found between both, suggesting that a careful separation of the different sources of noise
and pressure, and the use of such a model, can provide an
interesting practical tool to model the propeller-induced pressure fluctuations.
Further work will consist in the determination of the cavity volume in highly unsteady conditions, where the hypothesis on the linearity between dimensions of vapor cavities is
not valid. The current work provides the initial data to validate hydro-acoustics models for cavitating propellers, currently in development, and will be integrated into the E779A
database.
ACKNOWLEDGMENTS
The experiments were performed as part of the internship
of two students from Ecole Navale de Brest (France), Mrs
Daloubeix and Mrs Bidault. This internship was supervised
by Dr. Astolfi and Prof. Billard from the French Naval
Academy. The authors are also grateful to Mr Stefano Franchi
for his contribution to the measurements.
REFERENCES
Pereira, F., Salvatore, F., Di Felice, F., and Elefante, M., 2002,
Experimental and Numerical Investigation of the Cavitation Pattern on a Marine Propeller, Proc. 24th Symp. on
Naval Hydrodynamics, vol. 3, Office of Naval Research,
Fukuoka, Japan, pp. 236251.

Ito, T., 1966, An Experimental Investigation into the Unsteady Cavitation of Marine Propellers, Papers of the Ship
Research Institute, 11, pp. 118.
Takahashi, H., and Ueda, T., 1969, An Experimental Investigation into the Effect of Cavitation on Fluctuating Pressure
around a Marine Propeller, Proc. of the 12th International
Towing Tank Conference, Rome (Italy).
Bark, G., and Berlekom, W. B., 1978, Experimental Investigations of Cavitation Dynamics and Cavitation Noise,
12th Symposium on Naval Hydrodynamics, ONR, Washington, DC (USA), pp. 470493.
Matusiak, J., 1992, Broadband Noise of the Cavitating Marine Propellers: Generation and Collapse of the Free Bubbles Downstream of the Fixed Cavitation, Proc. of the
19th Symposium on Naval Hydrodynamics, ONR, Seoul
(Korea), pp. 701712.
Johnsson, C.-A., Rutgersson, O., Olsson, S., and Bjrheden,
O., 1976, Vibration Excitation Forces From a Cavitating
Propeller. Model and Full Scale Tests on a High Speed
Container Ship, Proc. of the 11th Symposium on Naval
Hydrodynamics, vol. VIII, Office of Naval Research, London (UK), pp. 4374.
Chiba, N., Sasajima, T., and Hoshino, T., 1980, Prediction
of Propeller-Induced Fluctuating Pressures and Correlation
with Full Scale Data, Proc. of the 13th Symposium on
Naval Hydrodynamics, ONR, Tokyo, pp. 89103.
Breslin, J. P., Van Houten, R. J., Kerwin, J. E., and Johnsson, C.-A., 1982, Theoretical and Experimental PropellerInduced Hull Pressures Arising from Intermittent Blade
Cavitation, Loading, and Thickness, SNAME Transactions, 90, pp. 111151.
Friesch, J., Johannsen, C., and Payer, H. G., 1992, Correlation Studies on Propeller Cavitation Making Use of a Large
Cavitation Tunnel, SNAME Transactions, 100, pp. 6592.
Friesch, J., 1998, Correlation Investigations for Higher Order Pressure Fluctuations and Noise for Ship Propellers,
Proc. of the 3rd International Symposium on Cavitation,
Grenoble (France), pp. 259265.
Johannsen, C., 1998, Investigation of Propeller-Induced
Pressure Pulses by Means of High-Speed Video Recording in the Three-Dimensional Wake of a Complete Ship
Model, Proc. of the 22nd Symposium on Naval Hydrodynamics, Office of Naval Research, Washington, D.C.
(USA), pp. 314329.
Huse, E., ed., 1996, Proc. of the 21st International Towing
Tank Conference, vol. I, chap. Measurements of Hull Pressure Fluctuations, Trondheim (Norway), pp. 6572, cavitation Committee.
Pereira, F., Avellan, F., and Dupont, P., 1998, Prediction of
Cavitation Erosion: An Energy Approach, J. Fluids Eng.,
120(4), pp. 719727.

Copyright National Academy of13


Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Jrgen Friesch /Christian Johannsen
Hamburg Ship Model Basin, Germany
The paper is very interesting to us since we
tried similar investigations at HSVA some years ago
(ONR 1998). The basic idea to correlate particular
cavitation occurrences such as the growth and collapse of sheet cavitation or the bursting of vortex
cavitation with the peaks of the pressure time function is very tempting indeed. Nevertheless, our investigations showed that this is difficult because different occurrences may happen simultaneously at different blades. It might be difficult to distinguish for
example whether a peak in the pressure is related to
the growth of sheet cavitation on one blade or to the
bursting of tip vortex cavitation on another blade. To
minimize this problem it is recommended to do those
investigations with a two-bladed propeller instead of
a four-bladed. This reduces the coincidence of different happenings.
The author claims that the collapse of tip
vortex cavitation mainly contributes to the third harmonic of blade frequency. What is the physical explanation for this assumption?
What is the authors opinion about the common assumption that amplitudes of higher order are
often a result of instable sheet cavitation. Is that true,
and if so, what is the physical mechanism?

structures. This phenomenon is driven by a Strouhal


law, and is well documented in particular for 2dimensional hydrofoils. If such a periodicity were to
exist on the cavitating blades of a rotating propeller,
it would certainly modulate the energy spectrum and
create harmonics at higher orders. However, the authors believe that this type of unsteadiness cannot be
considered here. The high speed visualizations do not
show the typical features of such a behavior. Moreover, the leading edge cavity has a lifetime of about 6
ms, which is largely unsufficient for a Strouhal
mechanism to take place and generate the large scale
structures typical of an unsteady situation. However,
we think that the collapse of structures like the cavity
and the secondary collapse of the tip vortex, discussed above, are to be considered as Dirac-type
excitation sources, which are likely to contribute to a
broad spectrum content. The effect would then be the
same as the collapse of vapor structures of an unsteady cavitation case.
Using a two-blade propeller, as suggested by
the discussers, or even a single blade propeller,
would certainly help clarify these two points.

AUTHORS REPLY
The authors thank Dr Friesch and Dr Johannsen for their interest in the work.
Regarding the first question, the authors believe that the 3rd harmonic is here related to the vortex rebound and implosion that follow the main collapse, which occurs simultaneously to the attached
cavity implosion. This is visible in the high speed
sequence shown in figure 4 for 0=3.5. Frames 14 to
20 show the implosion of the main cavity/tip vortex
structure. Frames 21 to 25 show the rebound and
second collapse of the tip vortex. This second implosion is visible on the pressure transducer P2, and to a
lesser degree on P1. It is also detected by the hydrophone H1, which is located inside the slipstream. The
remaining transducers (P3, P4, H2, H3 and H4) do
not witness this second collapse, but are essentially
sensitive to the major flow structures (blade trailing
wake for H2, attached cavity growth and collapse for
H3, H4, P3 and P4).
Unstable cavitation, or cloud cavitation, is
characterized by the periodic shedding of large vapor

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Roger Kinns
RK Acoustics, United Kingdom
The authors are to be congratulated on an
interesting and useful set of experiments, as well as
on their interpretation of results.
The comparison between measured pressure
fluctuation and the prediction from volume
acceleration is particularly interesting, because it
shows good agreement, even when the volume source
extends over a significant proportion of the blade
surface and when the pressure transducer is close to
the propeller itself.
Ideally, the volume source would be small in
relation to its distance from the transducer, so that the
physical situation corresponds closely to the sample
case of a stationary acoustic monopole. Please would
the author clarify the scale of the volume source in
relation to its distance from the transducers in their
experiments? Also, does the agreement improve
further if the transducer is further away from the
propeller, thereby reducing the influence of near-field
effects on the measured pressure?
AUTHORS REPLY
The authors thank Dr Kinns for his interest
in the work and for his comments.
The pressure data is provided by the
pressure transducer P2 (see figure 1), which is placed
in the plane of the propeller and at a distance of
about 140 mm from the blade tip when the blade is in
its vertical position (90 degrees). The cavity extends
up to 66% of the chordlength at 0.7R for the case at
the lowest cavitation coefficient (0=2.5), which
represents 47% of the distance between the tip and
the pressure transducer. The source is therefore of
the same order of magnitude as its distance to the
measurement point. The good agreement found
between the volume acceleration and the pressure
fluctuation indicates that the volume source behaves,
at least for the attached portion of it, as a monopole.
The agreement is expected to improve with distance,
but this has not yet been addressed in our study.
However, the ability to select the adequate orders,
hence limiting the number of exciting sources to
those of interest, is seen as the key feature for a
correct study of the correlations between volume and
pressure fluctuations.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

presented show that there might be an alternate


approach to that problem in particular, and to the
problem of scaling in general. The good agreement
obtained here between the pressure fluctuations and
the acceleration of the cavity volume indicates that
one could establish a number of spatial transfer
functions at the model scale and use these same
functions to derive the noise pattern from the
cavitation extension (or volume) measurement
performed at a different scale. Work is currently
undergoing to verify the validity of this approach.
The dissolved air content was monitored and
was set to 5 ppm before the measurements, through
intense degassing.

DISCUSSION
Ki-Han Kim
Office of Naval Research, USA
I would like to congratulate the authors for
their efforts in careful measurements and analyses of
the complex propeller blade surface and tip vortex
cavitation. I have a couple of questions that need
some clarifications.
In the text, no explanation was given about
how the cavity area and its fluctuations were
measured. How did the authors obtain the rms values
of the cavity area fluctuations as presented in Figure
9?
In the text (6th page of the paper), the
authors indicated that 0 = 7.5 is a non-cavitating
condition. In Figure 9, however, they showed cavity
area fluctuations in rms values for this cavitation
number, 0 = 7.5. Please clarify this.
In the Cavitation Extension section, the
authors explained that for supercavitating condition
in Figure 9, i.e. for 0 = 2.5, the maximum rms
values moves to = 100 deg. Should this be = 110
deg?
How would the authors correct the noise
data measured in the tunnel for the ship propeller in
free field?
What was the dissolved air content for the
experiments?

REFERENCES
Pereira F., Salvatore F., Di Felice F., Elefante M.,
Experimental and Numerical Investigation of the
Cavitation Pattern on a Marine Propeller,
Proceedings of the 24th Symposium on Naval
Hydrodynamics, Fukuoka (Japan), 8-13 July 2002.

AUTHORS REPLY
The authors thank Dr Kim for his comments
and questions.
The first question regards the measurement
of the cavity area and its fluctuations. The technique
was introduced in a previous work presented at the
24th SNH in Fukuoka (July 2002), which reference is
given below. The cavitation area is measured by an
image cross-correlation technique applied between a
cavitating image and a reference cavitation-free
image. This operation is then followed by a nonlinear dewarping transformation to obtain a correct
estimate of the area.
The
Authors
corrected
the
two
typographical errors quoted by the Discusser.
Experiments have covered a range of values for the
advance coefficient J, but the present work considers
only one case (J=0.175), for which cavitation does
occur at 0 = 7.5.
The Reviewer then addresses the important
issue of noise data correction between tunnel and free
field measurements. This has not been addressed in
the current state of the work. However, the results

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Tom van Terwisga
MARIN, Netherlands
Did the authors consider using another
analysis technique for analyzing the measured
pressures than the Fourier analysis? The background
for this question is that at MARIN we have observed
small phase variations in sheet cavitation for various
blade/passages. These phase variations result in
higher harmonics in the Fourier analysis result and
cannot be linked to the physical phenomena
occurring with the sheet cavitation.
AUTHORS REPLY
The authors thank Dr van Terwisga for his
interest in the work and for his comments.
The authors are aware that the Fourier
analysis can be misleading and create biased
interpretations. Our experience has also shown that
there are indeed small phase variations between
blades, due to the non-perfectly repetitive
phenomenology. We have noticed that those small
phase shifts would cause maxima of the energy
spectrum to shift slightly off the blade harmonics. To
overcome this difficulty, we have considered every
component of the Fourier decomposition and tracked
the maxima in the vicinity of the harmonics of the
blade passage frequency.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics


St. Johns, Newfoundland and Labrador, CANADA, 8-13 August 2004

Cavitation Inception in Co-Flow Nozzles


Richard S. Meyer, Frank J. Zajaczkowski, William A. Straka, and Eric G. Paterson
(The Pennsylvania State University, University Park, PA, USA)
Baker, et al. (1975), Lin and Katz (1987), Chahine, et al.
(1987), Ran and Katz (1994), and Gopalan et al. (1999),
development of empirical scaling relationships remains
elusive. Figure 1 shows a compilation of jet cavitation

ABSTRACT
A
complementary
experimental
and
computational research program has been developed to
study cavitation inception in co-flow nozzles, and initial
results are presented. Cavitation inception indices using
visual and acoustic criteria are presented over a range of
tunnel speeds for two nozzle trailing-edge geometries,
contoured and blunt, and for velocity ratios of 0.5, 0.75,
and 1.0. In all cases, inception was characterized by
vortex cavitation found 1-2 jet diameters downstream of
the nozzle exit and in the cores of Kelvin-Helmholtz
vortex rings. The contoured and blunt nozzles had
mean cavitation inception indices of 0.33 and 0.80,
respectively. Detached-eddy simulation is used to
predict the turbulent vortices in the shear-layer
downstream of the nozzle exit. CFD results show
small-scale eddies very close to the nozzle exit, but with
subsequent growth of coherent vortex rings in the
streamwise direction. It is shown, as expected, that low
pressure regions correspond with the vortex cores.
Mean and RMS computational and experimental axialvelocity profiles are compared at several stations in the
jet and show good agreement in both shape and
magnitude, thus giving confidence that the DES model
is capable of accurately resolving low-pressure events
that trigger nuclei growth and cavitation inception.

inception index, defined by =

P Pv , as a function
1
V2
2

of jet diameter for various experiments.


1.8
Ooi

1.6

Lienhard &
Stephenson
Lienhard & Goss

Jet Cavitation Index

1.4
1.2

Jorgenson

Ball

0.8

Kobayashi

0.6

Baker

0.4

Lin & Katz


Lin & Katz

0.2

Gopalan et al.

0
0

20

40

60

80

100

Jet Diameter (mm)

Figure 1. Jet cavitation index as a function of jet


diameter from previous investigations (compiled by Lin
and Katz, 1987)

INTRODUCTION

In quiescent flow, they found the jet cavitation inception


number to scale directly with jet diameter. Their
database, however, contains conflicting trends in
measured cavitation inception that are most likely due to
facility biases (e.g., water quality, inception-calling
procedures) and large variations in flow conditions and
geometries. Moreover, these data do not address the
case of a co-flowing jet.
Nevertheless, these investigations have shown
that cavitation in jet flows is primarily due to vortical
structures, in the shear layer downstream of the nozzle
trailing edge, and the low pressures that result in the
vortex cores. For example, in Gopalan et al. (1999), it
was shown that by altering the nature of the boundary
layer on the inside of the jet (i.e., untripped vs. tripped),
flow structures in the shear layer were significantly

Defining and scaling of cavitation inception in


jets remains a difficult task. Inception occurs due to the
unstable growth of nuclei subjected to the low pressures
typically found at the core of turbulent vortices.
Typically, inception is defined by the occurrence of a
few cavitation events per second, i.e., low-event-rate
cavitation (LERC), which is determined by acoustic
measurements. It is a transient phenomenon influenced
by turbulence, organized unsteadiness such as vortex
shedding, and bubble dynamics, the latter of which is
also a function of nuclei-size distribution.
While there have been numerous experimental
studies on jet cavitation which have led to improved
understanding of the physics, e.g., Kobayahsi (1967),

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

modeled with a polytropic compression law), is the


surface tension, and is the kinematic viscosity. The
bubble position in the flow field is computed using the
motion equation described by Johnson and Hsieh (1966)
which is a simple balance of momentum and the forces
acting on the bubble.

altered, which in turn modified the cavitation indices.


The untripped case had a cavitation index i=2.5, where
p and pv are the far-field and vapor pressure,
respectively, and showed that inception occurred in the
form of inclined cylindrical bubbles in the core of
longitudinal vortices within 0.55 jet diameters of the jet
exit. In contrast, the tripped jet had a i=1.7, with
inception occurring in the form of distorted spherical
bubbles at 2 jet diameters downstream in the core of
Kelvin-Helmholtz vortex rings.

Hsiao and Chahine (2004) have shown that


modification to the basic approach result in improved
prediction of inception. They introduce a slip velocity
term into Equation (1) and modify the definition of the
encounter pressure p(x,t) to be the average of all liquid
pressures at the various parts of the bubble surface, thus
leading to the acronym SAP (surface-averaged
pressure).

Simulations by Cerutti et al. (2000a), although


somewhat limited due to the laminar and axisymmetric
formulation (i.e., they were unable to resolve
longitudinal eddies), have further shown that the
slenderness ratio /D, where is the boundary-layer
thickness on the inside wall of the jet and D is the jet
diameter, is an important parameter in governing
cavitation inception. Since /D governs the strength of
the resulting shear layer, this is not surprising; however,
Cerutti et al (2000a) did show that as /D decreases, i
increases and the location of inception occurs
increasingly closer to the jet exit. They also showed
that for fixed , variation of Reynolds number Re had
very little effect on pressure minima or eddy shedding
frequency, and therefore on cavitation inception.

Some of the challenges in using computational


methods include incorporation of the randomness in
nuclei populations, definition of inception criteria, and
the sensitivity of getting the nuclei close enough in
space and time to the low pressure events so as to
cavitate. A number of studies have looked at the
sensitivity of cavitation to initial nuclei size and seeding
location (e.g., Meyer et al, 1992; Farell, 2000; Cerutti et
al., 2000b, Kim et al., 2003) such that it is known that
large bubbles are more easily captured by vortices than
small bubbles and that initial position can significantly
influence predicted inception indices if pressure minima
are missed. Most simulations incorporate both a size
criteria (e.g., bubbles of 1mm are assumed big enough
to be seen visually) and an acoustic criteria, the latter of
which is based upon radiated acoustic pressure (Blake,
1986).

Computational
modeling
of
cavitation
inception requires fidelity resolution of the flow
physics, i.e., unsteadiness due to turbulence and
organized vortex shedding, and bubble dynamics
(nuclei-size and number distributions, bubble dynamics,
and bubble-flow interactions). Concerning the former,
there are a number of examples in the literature which
are based upon RANS or laminar-flow simulations (e.g.,
open-water propeller [Hsiao and Pauley, 1999], Schiebe
headform [Farrell, 2000], axisymmetric jet [Cerutti et
al., 2000a; Cerutti et al., 2000b], and a ducted marine
propulsor [Kim et al., 2003]); however, large-eddy
simulation methods have not yet been brought to this
problem.

The objective of our current research is to


study the underlying physics of cavitation inception in
co-flow nozzlesan application which has received
much less attention than a jet emptying into a quiescent
tankwhich includes the complication of two merging
boundary layers.
Contoured and blunt nozzle lip
geometries are studied over a range of velocity ratios,
including the Bollard condition of quiescent outer flow.
Our approach is based upon both experimental and
computational fluid dynamics (CFD) methods and our
goal is to advance the state-of-the-art in computational
modeling through application and validation of a hybrid
Reynolds-averaged Navier-Stokes (RANS) / large-eddy
simulation (LES) methods for co-flow nozzles, and
coupling it with the SAP bubble-dynamics model of
Hsiao and Chahine (2004). Specifically, our approach
is based upon detached-eddy simulation (DES) (Strelets,
2001), a model which is capable of resolving the
important flow structures and which can eventually
eliminate the issue of scaling through direct simulation
of full-scale vehicles/devices.
We report progress
towards meeting these objectives and goals.

For the bubble dynamics component of the


computational model, the typical approach (Hsiao and
Pauley, 1999; Farrell, 2000; Cerutti et al., 2000a; Kim et
al., 2003) has been to solve the classical RayleighPlesset equation for the bubble radius R
2

d 2 R 3 dR
1
+
= ( pv + pg p ( x, t )
2
2 dt

dt
2 4 dR

R
R dt

(1)

where p(x,t) is the local, or encounter, pressure


computed from CFD, t is time, is the fluid density, pg
is the gas pressure inside the bubble (which is usually

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

image velocimetry (PIV). In addition, the large viewing


windows allow visual observation of cavitation with the
aid of pulsed strobe lighting. In situ hydrophones
provide simultaneous acoustic data.

It should be noted that the work presented here


is part of a larger research effort on nozzles and jets
aimed at understanding the effect of nozzle shape,
development of scaling relations (tests performed on
geosims in both 48 Garfield Thomas Water Tunnel and
12 water tunnel), study of structural acoustics, and
design integration into overall systems.
In the following sections, the experimental and
computational methods will be briefly described, after
which, the results will be presented and discussed.
Finally, conclusions and future work will be
summarized.
EXPERIMENTAL METHODS
Facilities and Model Geometry

Tests were conducted in the ARL 12-inch


water tunnel in March 2004.
Figure 3. 101.6 mm and 25.4mm blunt and contoured test
nozzles.

A one-inch-diameter jet was mounted normal


to the flow direction with its exit along the tunnel test
section centerline. It was located axially such that it
provided visual access to the flow downstream of the jet
through the test section windows. The jet flow was
supplied through a two-inch-diameter pipe entering the
tunnel from the top upstream of the contraction and then
turning 90 degrees through a pipe elbow until parallel to
the direction of the tunnel flow. An illustration of the
water tunnel test set-up is shown in Figure 2. Two
axisymmetric nozzles were designedone blunt-nosed,
and one contouredboth with a contraction area ratio
of 1.57, shown in Figure 3. For the current 12-inch
diameter water tunnel experiment, only the 25.4 mm
diameter nozzles were used. The 101.6mm nozzles were
used during the 48-inch diameter water tunnel
experiments, results of which are not presented here.

Figure 2. Illustration of the 12-inch water tunnel test showing


relative location of nozzle, tunnel contraction, and tunnel test
section.

Flow Conditions

The test matrix consists of three co-flow


freestream-to-jet velocity ratios (VR) of 0.5, 0.75, and
This facility features a 304.8 mm diameter by 762 mm
long test section in which velocity is continuously
variable up to 24.4 mps with independent control of
pressure from 413.7 to 20.7 kPa. A de-gassing system
allows control of dissolved air content to as low as 2
molar ppm. The tunnel test section provides optical
access for laser Doppler velocimetry (LDV) and particle

1.00, where VR is defined by VR =

V
VJ

, over a range

of jet velocities (6.1 18.3 m/s).


In addition, a flow blockage plate was installed
at the end of the blunt nozzle so that data for VR=0
(Bollard condition) case could be obtained. Data were
collected at this condition so as to provide a basis to

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

studies of propeller cavitation at ARL to be in the range


of 10-15%, based on several inception calls on the same
test blade. This included operator-to-operator variation
as well as trial-to-trial variation for the same operator.

compare the co-flow jet cavitation inception results of


the present study to the existing database of submerged
waterjets into a quiescent tank (Baker, et al., 1975;
Kobayashi, 1967; and Lin and Katz, 1987); in addition,
it is used for CFD validation. More details of this set-up
will be given in the Results section of this paper.

COMPUTATIONAL METHODS
Grid Generation, Computational Parameters, and
Boundary Conditions

Instrumentation

Flow field measurements were obtained near


the jet exit area using Laser Doppler Velocimetry
(LDV)
and
two-dimensional
Particle
Image
Velocimetry (PIV) so as to provide time-averaged, RMS
and instantaneous velocity fields, and Reynolds stresses.
Pipe velocity profile upstream of nozzle was also
measured using LDV so as to provide CFD inlet
boundary conditions. Inception was determined using
both visual and acoustic techniques. Visual cavitation
inception pressure was determined by systematically
lowering the pressure from a sufficiently high value
where it can be determined that no cavitation exists.
The rate at which the pressure was lowered was kept as
slow as was practical from within the time constraints of
the test. Typically, the rates were on the order of 1-2 psi
per minute. Cavitation was first observed in the core of
the secondary vortices which formed in the mixing layer
between the jet and the surrounding co-flow. The
bubblessingle, sporadic, transient eventsappeared
as short vortex ropes on the order of 2.5-5.0mm in
length, usually observed 1-2 nozzle exit diameters
downstream of the exit. Acoustic cavitation inception is
based upon sound pressure level (SPL) measurements
obtained from 10 to 200 kHz using three tunnel
wall/window mounted hydrophones in parabolic
reflectors (and also a downstream array for 48-inch
diameter test only). A sound pressure level increase of
in 2-3 dB due only to a decrease in the free-stream
pressure indicates cavitation inception. However, data
obtained show a large increase on the order of 10-20dB
when cavitation was present.

The computational domain models the astested flow conditions and includes the inside of the
nozzle and the water tunnel test section.
The grid
system extends from -20 < x/D < 100, where the origin
is at the nozzle exit and D is the nozzle diameter.
Overset grids, as shown in Figure 4, were generated
using Gridgen from Pointwise, Inc. Overall geometry
is simple and, except for the nozzle lip, nested box grids
were used to fill the computational domain. Given the
flexibility of the overset approach, all grids remain
unchanged for the blunt- and contoured-nozzle variants,
except for the lip grids. Near-wall spacing was set to
5.0x10-5, which is based upon a Reynolds of
Re = U j D = 5.1x105 and wall coordinate for the first
point to be y + = U y = 1. Overset-refinement grids
were used to resolve turbulent eddies in the shear layers.
These refinement grids were set to have nearly isotropic
cells with a grid size = 0.012. The overall grid system
was composed of 32 blocks and 1.1x106 points.
Algorithm-related parameters to control the
simulation include:
2nd-order upwind convective
scheme for the RANS regions; 2nd-order in time; PETSc
GMRES solver; 50 sub-iterations maximum for
momentum and turbulence equations; 3 PISO steps; 250
sub-iterations maximum for pressure-Poisson equation;
and convergence criteria of 10-4 for momentum (residual
based upon change between iterations) and 10-4 for
pressure (normalized residual based upon error of
discrete equation).
The simulation matrix is based upon the test
matrix. Simulations were conducted for VR = 0, 0.5,
0.75, and 1.0 for both nozzle geometries, which
represents 8 cases. In addition, both URANS and DES
was performed on several cases for comparison
purposes.

Uncertainty Assessment

Cavitation inception measurement is subject to


error from a variety of sources. Measurement of
pressure is subject to a random error on the order of
0.5%. Measurement of temperature, used in the
calculation of vapor pressure, is subject to a similar
error. The largest source of error by far in the case of
visual cavitation inception determination is the
subjectivity of what constitutes visual inception and the
effect of using strobes to illuminate cavitation events.
The randomness of cavitation at low event rates allows
for inconsistency from trial to trial. Estimates of this
operator uncertainty were determined in previous

The grid system contains a mixture of overset


and pointwise-continuous
multiblock boundary
interfaces. For the overset boundaries, PEGASUS 5.1
(Suhs, et al. 2000) was used to compute double-fringe
interpolation coefficients for all outer- and holeboundary fringe points. Level-2 interpolation was used
to minimize overlap and improve interpolation quality.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

turbulence model uses a linear closure and the blended


k-/k- SST 2-equation (Menter, 1994). Efficient
parallel computing is achieved using coarse-grain
parallelism via MPI distributed computing. For timeaccurate unsteady simulations, global solution of the
pressure-Poisson equation is achieved using preconditioned GMRES and the PETSc libraries.
Detached-eddy simulation (Strelets, 2001) is a
three-dimensional unsteady numerical method using a
single turbulence model, which functions as a subgridscale model in regions where the grid density is fine
enough for a LES, and as a RANS model in all other
regions. Implementation of DES in CFDSHIP-IOWA
was accomplished by modifying the turbulence model
and convective-term discretization. The turbulence
model is modified by introducing a DES length scale
%l = min ( l , C )
k
DES

Figure 4. Overset Grid System.

(2)

which compares the turbulence length scale from


RANS l k , to the local grid size , where is based on
the largest dimension of the local grid cell:
Non-overset boundaries were prescribed to be
either no-slip, exit, or prescribed, the latter of which
uses the experimental data to calibrate our inflow
profile. Details of these conditions are provided in the
CFDSHIP-IOWA users manual (Paterson, et al., 2003).
It is noted that all boundary conditions, including the
overset-grid trilinear-interpolation formula, were built
directly into the global coefficient matrix for the
pressure-Poisson solver which resulted in significant
acceleration of the iterative convergence.

l k =

k 1/ 2

(3)

= max ( x , y , z )

(4)

CDES is a model constant with a value between 0.78 and


0.61 weighted by the Menter k-/k blending function
(Menter, 1994). The new length scale %l in the
dissipative term of the k-transport equation

Flow Solver

k
= * k =
DRANS

DES, and hybrid RANS/LES methods in


general, have received much attention recently due to
the promise of delivering unsteady three-dimensional
flow fields with local stochastic content. For the work
presented here, we used a recently developed DES
model and the CFDSHIP-IOWA (Paterson, et al., 2003)
flow solver. Our DES model has recently been
developed and applied to circulation control (Paterson
and Baker, 2004), trailing-edge flows (Paterson and
Peltier, 2004), and hydrofoil singing (Paterson et al.,
2004). In addition, this DES model has been installed in
other in-house codes for simulation of large-scale
cavitation over a hydrofoil (Kunz et al., 2003).

k 3/ 2
l kw

k
DDES
=

k 3/ 2
l%

(5)

The effect of this modification is to increase dissipation


in LES regions such that the turbulence budget shifts
energy from sub-grid, or modeled, scales to resolvable
scales as defined by the filter width CDES.
The second modification aims to reduce
numerical dissipation inherent in the upwind
convective-term discretization scheme.
The
implemented approach is based upon a hybrid
central/upwind approximation of the convective terms.
Finv = (1 ) Fctr + Fupw

CFDSHIP-IOWA is a general-purpose parallel


unsteady incompressible Reynolds-averaged NavierStokes (RANS) CFD code.
The computational
approach is based upon structured overset-grid, higherorder finite-difference, and pressure-implicit splitoperator (PISO) numerical methods.
Production

(6)

where is define as
m
n

U t
CDES
= max tanh

, tanh
l k

Copyright National Academy of Sciences. All rights reserved.

(7)

Twenty-Fifth Symposium on Naval Hydrodynamics

The result is that smoothly transitions between 1.0 in


the RANS regions, resulting in a local upwind scheme,
and near 0.0 in the LES regions, where a centered
scheme is desired. In addition, a Courant-number
constraint of 1.0 has been imposed which requires that
time step be sufficiently small to support turbulent
eddies. The coefficients n,m permit the interface
between RANS and LES regions to be arbitrarily
"sharpened", however, currently we use n=m=1 due to
the fact that higher-order coefficients have resulted in
unstable simulations.
Figure 5 shows color contours of for the
bollard condition simulation. The "white" region is
interpreted by the code to be URANS. DES is active in
the "green/blue" region. Figure 6 shows the breakdown
of the various energy ranges. In the RANS region,
production and dissipation are sub-grid and hence, are
modeled. In the energy-containing range, some
production is resolved; however, dissipation is still subgrid. In the inertial range, all production is resolved,
while dissipation is still sub-grid. In the dissipation
range, both production and dissipation are resolved.

Figure 5. Contours of Convective-Scheme Blending Function.

1 k
1
= U + U k k + U k U k +k
U
k 2
2

Normalized Power Spectral Density

101

In CFDSHIP-IOWA the convective terms are


discretized with the following higher-order upwind
formula

100
10 -1
uu

10 -2
10 -3
10 -4

DES

10 -5

(8)

The discrete upwinding operators k and +k

10 -6
10

LES

DNS

-7

10-2

10-1

100

101

102

10 3

104

105

Energy-Containing Range Inertial Range Dissipation Range


~1
~1 l
~1 L
Geometric Scale
IR Eddy Scale Kolmogorov Scale

are

RANS

calculated using a 5 point stencil

= Wmmi 2 + Wmi 1 + Wni + W pi +1 + W ppi + 2


i

+ = W ppi 2 W pi 1 Wni Wmi +1 Wmmi + 2

Figure 6. Normalized power spectral density and the


various energy ranges of the Discrete Eddy Simulation
technique.

(9)

with the coefficients Wi defined as a weighted sum of


the coefficients wi for the 4th-order central and the 2ndorder upwind schemes, Wi = (1 ) wi4c + wi2u .

RESULTS AND DISCUSSSION

In this section, the results are presented and


discussed in the following order: Cavitation,
Instantaneous Flow Field, and Mean and RMS Flow
Fields.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

inception index, i was significantly lower, due probably


to the contoured nozzle, in contrast to the blunt nozzle
geometry, where a separated region between the jet flow
and the outer flow resulted in increased vorticity in the
mixing layer. In the next section, the dynamics of the
flow in these regions will be discussed in more detail.
Again, the significantly higher variance at the lower
tunnel velocity conditions was attributed to the low
pressures necessary for inception. At these extremely
low pressures, there were significant amounts of free
gas in the tunnel due to small leaks as well as gas being
forced out of solution in the lowest pressure regions of
the water tunnel.

Cavitation

Cavitation measurements were conducted on


the one-inch nozzles in the 12-inch water tunnel
experiment. A typical view of fully developed cavitation
on the blunt nose is shown in Figure 7. The cavitation
inception index i, based upon visual criteria, is shown
in Figures 8 and 10 as a function of water tunnel
velocity. Each data point shown on the plots represents
four independent visual calls.

Blunt Nozzle
1.2
VR=0.5

3.9 ppm

VR=0.75
VR=1

3.6 ppm

Cavitation Index

3.6 ppm
3.6 ppm

3.6 ppm

0.8

3.6 ppm

3.6 ppm

3.6 ppm

3.6 ppm

3.6 ppm
0.6
3.9 ppm
0.4

Figure 7. Developed Cavitation for blunt-trailing-edge


nozzle.

0.2

0
5

10

15

Tunnel Velocity (mps)

(a) blunt nozzle


Round Nozzle

Air content was between 3.6 and 3.9 molar


ppm for all points. The velocity ratio, VR, ranged
between 0.5 and 1.0. In all cases, the inception index, i
ranged between 0.55 and 1.12. The mean index of all
visual inception calls regardless of the tunnel velocity of
velocity ratio was |i |= 0.8. What is interesting, is the
trends with tunnel velocity and velocity ratio. In the
range tested there appears to be very little change in
visual cavitation inception index with tunnel velocity.
The slight rise in index at the lower velocities may be
attributed to the low pressures necessary for inception.
At these low pressures, there were small amounts of free
gas in the tunnel due to leaks as well as gas being forced
out of solution in the lowest pressure regions of the
water tunnel. What is also interesting is that there does
not appear to be a consistent trend of visual cavitation
index with co-flow jet velocity ratio. In all cases,
inception was characterized by vortex cavitation found
1-2 jet diameters downstream of the nozzle exit.

6.4 ppm Air Content throughout

0.9

VR=0.5
0.8

VR=0.75
VR=1

Cavitation Index
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
5

10

15

Tunnel Velocity (mps)

(b) contoured nozzle


Figure 8. Cavitation Inception Index, i as a function of
water tunnel velocity.

Figure 8b shows the inception data for the


contoured nozzle. The mean cavitation index was found
to be 0.33; the standard deviation was calculated to be
0.17, over 50% of the mean value. In this case, the

Results for the Bollard, VR=0, case are shown


in Figure 10. For this case, the blunt nozzle was
mounted in the tunnel, and an acrylic flow blockage
plate was installed flush with the nozzle exit, normal to

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

the direction of flow. A photograph of cavitation in the


bollard condition is shown in Figure 9.

2.9 ppm Air Content throughout

0.8
0.7

Holes were drilled in the plate to allow for a small


amount of flow, necessary to prevent an excess of
bubbles from collecting in the test section. The plate
was secured to the tunnel wall with four aluminum
supports mounted to the back-side of the plate. The
inception data in Figure 10 show an average, |i |= 0.55.
This result appears to be on the low end of those
presented in Lin and Katz (1987) where inception
values for 25.4 mm diameter submerged water jets
ranged from approximately 0.3 to 1.4. For the present
case, we see the trend that the inception index is greater
at the lower jet velocity. As was mentioned in Lin and
Katz (1987), this trend is not universal.
More
importantly the data show a degradation in cavitation
performance between the simple waterjet and the coflow jet. Comparing Figures 8a and 10, one will notice
that the cavitation index range increases from 0.35-0.75
for the simple jet to 0.55-1.12 for the co-flow jet.

Cavitation Index

0.6
0.5
0.4
0.3
0.2
0.1
0
10

15

20

Jet Velocity (mps)

Figure 10. Cavitation Inception Index, i as a function of jet


velocity for the bollard condition in the GTWT 12-inch water
tunnel.

Figure 9. View of Bollard (VR=0) water tunnel test hardware

Acoustic cavitation results for the Bollard


condition show a sharp rise in the one-third octave
levels at a of approximately 0.6. Figure 11 shows the
acoustic levels in 20 kHz and 50 kHz bands. The
cavitation event rate as determined by analysis of the
acoustic signal as a function of is shown on the same
plots. A dramatic rise in the event rate is seen at =0.6.
Prior to this point, low event rate cavitation events are
present but have minimal impact on the acoustic
performance relative to the background level.

Figure 11. Acoustic cavitation results for the Bollard case at


V=57 fps in both 20 and 50 kHz one-third octave bands

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Instantaneous Flow Field

Visualization of the instantaneous flow field was


accomplished using the DES results and a postprocessing method for extracting vortical structures.
Figure 12 shows Kelvin-Helmholtz vortex rings
downstream of the nozzle lip using iso-surfaces of the
intrinsic swirl parameter [Berdahl and Thompson, 1993]
colored by normalized helicity. Frames are shown at
two different times in the simulation. It can be seen that
after some time, the near nozzle vortices experience
breakdown which leads to small turbulence scales very
near the nozzle. However, downstream, the vortices
roll-up into more coherent structures. From our
simulations, the slenderness ratio was estimated to be
/D = 0.005. Based upon Cerutti et al. (2000b), small
/D leads to decreased vortex size and core pressures,
increased shedding frequency, and increased i. The
near-nozzle vortex spacing gives a wavelength of /D =
0.15, and a shedding frequency of f* = fD/Uj = 3.33.
Qualitatively, this trend in comparison to Gopalan et al.
(1999) is correct wherein /D = 0.2 and fD/Uj = 1. In
addition to the intrinsic swirl parameter, Figure 13
shows iso-surfaces of pressure coefficient at a level of
Cp=-0.2. The correlation of low pressure with vortex
cores is clearly shown.

Figure 13. Visualization of DES Flow Field and Correlation


of Pressure with Turbulent Vortices. Iso-Surface of Intrinsic
Swirl Parameter (Gray) and Cp =-0.2 Iso-Surface of Pressure
(Blue).

Figure 14 shows a comparison of the DES instantaneous


computed flowfield with PIV experimental data for the
Bollard, VR=0, case. The PIV image comprises vector
arrows with color corresponding to velocity magnitude.
The DES results show small eddies near the nozzle,
growing into coherent rings downstream with
occasional bursts of longitudinal vortices. The PIV
results show regions of vorticity consistent with the
CFD results, and an overall spreading angle of vorticity
(yellow lines) consistent with that of the DES results.
Figure 15 shows the co-flow case, VR=0.5. The spatial
development in the streamwise direction is smaller due
the convection velocity of the free stream. Thus, the
spreading angle is less than for the quiescent case, as is
observed in the PIV results.

Figure 12. Contours of Intrinsic Swirl Parameter at


t = t1 and t = t1 + 20

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

a) Experimental PIV results

a) Experimental PIV results

Low pressure
due to blunt lip

CP = -0.8
1.8D

CP = -0.21
1.5 D

Blunt nozzle, VR = 0.5


Iso-surfaces of swirl and pressure

Sharp nozzle, VR = 1
Iso-surfaces of swirl and pressure

b) Computational DES results


b) Computational DES results

Figure 15. Comparison of the PIV experimental results


with the DES computational results for the sharp nozzle
in a co-flow with VR=0.5.

Figure 14. Comparison of the PIV experimental results


with the DES computational results for the sharp nozzle
in a quiescent floe (VR=0).

10

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Mean and RMS Flow Fields

Mean and root-mean-square (RMS) statistics


were computed for each dependent flow variable
(U,V,W,k,w,p) at every point in the grid. For the results
presented here, DES statistics were computed over 2500
time steps and experimental statistics were computed
over 200 PIV images. For the former, this corresponds
to an integration period of 25.0 for t = 0.01.

Experiment; x/D = 0.25


Experiment; x/D = 0.5
Experiment; x/D = 0.75
Experiment; x/D = 1.25
CFD; x/D = 0.25
CFD; x/D = 0.5
CFD; x/D = 0.75
CFD; x/D = 1.25

1
0.8
0.6
0.4

r/D

0.2
0

-0.2
-0.4
-0.6
-0.8
-1

-0.5

0.5

U mean/U ref

1.5

Figure 17. Mean axial-velocity profiles at locations


downstream of nozzle.
Figure 16. Contours of Mean Axial Velocity, and locations
for comparison with experimental data.

Figure 18 shows a comparison of the DES result to the


experiment for the RMS axial-velocity component.
Agreement is quite good, with the DES capturing the
overall trends of increased magnitude and shear-layer
width with increasing distance downstream.

Figure 16 shows contours of the mean axialvelocity field for a symmetric jet with all flow features
averaged out. This result indicates that the statistics
integration period is sufficient, at least for the
preliminary results reported herein. This figure also
shows the location for the comparisons between CFD
and experiment.
Figure 17 shows mean axial velocity profiles
downstream of the nozzle at x/D = 0.25, 0.5, 0.75 and
1.25. Over this distance the mean velocity shows very
little change. Satisfactory agreement is shown between
experiment and DES.

Figure 18. RMS axial-velocity profiles at locations


downstream of nozzle.

11

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Mean velocity and vorticity fields are shown in Figure


19 over the same domain and at the same contour levels.
Again, agreement is shown to be qualitatively good.

SUMMARY AND FUTURE WORK

A
complementary
experimental
and
computational research program has been developed to
study cavitation inception in co-flow nozzles and initial
results are presented. Cavitation inception indices using
visual and acoustic criteria are presented over a range of
tunnel speeds for two nozzle trailing-edge geometries,
contoured and blunt, and for velocity ratios of 0.0, 0.5,
0.75, and 1.0. In all cases, inception was characterized
by vortex cavitation found 1-2 jet diameters
downstream of the nozzle exit and in the cores of
Kelvin-Helmholtz vortex rings.
The contoured and
blunt nozzles had

Finally, an analysis of the turbulent kinetic


energy (TKE) is shown in Figure 20. Contributions to
the TKE are from resolved turbulence field, uir uir and
from the unresolved subgrid scales, uis uis
k tot =

1
1
1
ui ui = uir uir + uis uis = uir uir + k SGS (10)
2
2
2

This figure shows that the resolved scales are most


dominant at one diameter downstream and that most of
the turbulence very near the nozzle lip is sub-grid. This
suggests that the grid should be further refined. It is
also clear that upstream flow to the nozzle is a low
turbulence inflow.
While this replicates the
experimental condition, it highlights the need for
synthetic turbulence inflow boundary conditions for
simulation of future configurations where internal
turbulence levels could potentially be higher.

Figure 20. Contours of turbulent kinetic energy: resolved


turbulence, sub-grid turbulence, and total.

mean cavitation inception indices of 0.33 and 0.80,


respectively.
Detached-eddy simulation was used to predict
the turbulent vortices in the shear-layer downstream of
the nozzle exit. CFD results show small-scale eddies
very close to the nozzle exit, but with subsequent
growth of coherent vortex rings in the streamwise
direction. It is shown, as expected, that low pressure
regions correspond with the vortex cores.

Figure 19. Comparison of Mean Velocity and Vorticity


Fields for the quiescent flow, co-flow blunt nozzle, and coflow sharp nozzle.

Mean
and
RMS
experimental
and
computational axial-velocity profiles are compared at

12

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

several stations in the jet and show good agreement in


both shape and magnitude, thus giving confidence that
the DES model is capable of accurately resolving lowpressure events that trigger nuclei growth and cavitation
inception.

6.

Chahine, G.L, Kalumuck, K.M., Frederick, G.S.,


and Watson, R.W., Visualization Study of
Structured Cavitating Jets, ASME Fluid
Engineering Division Cavitation and Multiphase
Flow Forum, FED-Vol. 50, 1987.

Further work will focus on completing the


matrix of simulations so that effect of velocity ratio and
geometry can be more fully explored. Through more
detailed comparisons and grid studies, it is expected that
grid requirements can be established before
incorporating the bubble dynamics model of Hsiao and
Chahine (2004). Given the unsteady nature of the DES,
it is expected that large numbers of bubbles over a range
of initial radii will be required since it is impractical to
store the entire simulation. In parallel, new PIV-based
techniques for non-intrusive measurement of pressure
will be investigated as a means for acquiring data for
both CFD validation and better understanding of
correlation of flow structures and cavitation inception.

7.

Choi, J.K., and Chahine, G.L., Non-Spherical


Bubble Behavior in Vortex Flow Fields,
Computational Mechanics, Vol. 32, pp. 281-290,
2003.

8.

Farrell,
K.J.,
An
Eulerian/Lagrangian
Computational Analysis for the Prediction of
Cavitation Inception, Ph.D. Thesis, Department of
Mechanical Engineering, The Pennsylvania State
University, August, 2000.

9.

Gopalan, S., Katz, J., and Omar, K., Near-Field


Flow Structure and Cavitation Inception in Jets,
Proceedings of AMSE/JSME Symposium on
Cavitation Inception, July 1999.

10. Hsiao, C.-T., Chahine, G.L., and Liu, H.-L.,


Scaling Effect on Prediction of Cavitation
Inception in a Line Vortex Flow, ASME Journal of
Fluids Engineering, Vol. 125, pp. 53-60, Jan. 2003.

ACKNOWLEDGEMENTS

This work has been supported by ONR Grant


Number N00014-04-1-0242, program manager Dr. KiHan Kim. The DOD High Performance Computing
Modernization Office and Army Research LaboratoryMajor Shared Resource Center are acknowledged for
providing computing resources.

11. Hsiao, C.-T. and G.L. Chahine, Prediction of Tip


Vortex Cavitation Inception Using Coupled
Spherical and Nonspherical Bubble Models and
Navier-Stokes Computations, Journal of Marine
Science and Technology, Vol. 8, pp. 99-108, 2004.
12. Hsiao, C.-T., and L.L. Pauley, Study of Tip Vortex
Cavitation
Inception
Using
Navier-Stokes
Computation and Bubble Dynamics Model,
Journal of Fluids Engineering 121(1): 198-204,
March 1999.

REFERENCES

1.

Baker, C.B., Holl, J.W., and Arndt, R.E.A.,


Influence of Gas Content and Polyethylene Oxide
upon Confined Jet Cavitation in Water, Applied
Research Laboratory Technical Memorandum, TM
75-274, Penn State University, November 1975.

2.

Berdahl, C.H., and Thompson, D.S. Eduction of


Swirling Structures Using the Velocity Gradient
Tensor, AIAA Journal, Vol. 31, No. 1, 1993.

3.

Blake, W.K., Mechanics of Flow-Induced Sound


and Vibration, Academic Press, 1986.

4.

Cerutti, S., Knio, O.M., and Katz, J.,


Computational Study of Jet Cavitation Inception
Using Phase Analysis, 2000 ASME Fluids
Engineering Division Summer Meeting, Boston,
MA, June 2000b.

5.

Cerutti, S., Knio, O.M., and Katz, J., Numerical


Study of Cavitation Inception in the Near Field of
an Axisymmetric Jet at High Reynolds Number,
Physics of Fluids, Vol. 12, No. 10, pp. 2444-2460,
Oct. 2000a.

13. Kim, J., Paterson, E.G., and Stern, F., Verification


and Validation and Sub-Visual Cavitation and
Acoustic Modeling for Ducted Marine Propulsor,
Eight International Conference on Numerical Ship
Hydrodynamics, Busan, Korea, Sept. 2003.
14. Kobayashi, R., ASME Journal of Basic
Engineering, Vol. 89, No. 3, pp. 677-685, 1967.
15. Kunz, R., Lindau, J.W., Kaday, T.A., and Peltier,
L.J., Unsteady RANS and Detached Eddy
Simulations of Cavitating Flow over a Hydrofoil,
Fifth International Symposium on Cavitation
(CAV2003), Osaka, Japan, Nov. 2003.
16. Lin, H. J. and Katz, J., The Effect of Background
Noise on Cavitation Phenomena in Water Jets,
ASME Fluid Engineering Division Cavitation and
Multiphase Flow Forum, FED-Vol. 50, 1987.

13

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

17. Menter,
F.,
Two-equation
eddy-viscosity
turbulence models for engineering applications,
AIAA Journal, Vol. 32, No. 8, 1994.
18. Meyer, R. S., Billet, M.L., and Holl, J.W., Journal
of Fluids Engineering, Vol. 114, pp. 672-779, 1992.
19. Paterson, E.G., Wilson, R.V., and Stern, F.,
General-Purpose Parallel Unsteady RANS Ship
Hydrodynamics Code: CFDSHIP-IOWA, IIHR
Report #432, IIHR Hydroscience and Engineering,
The University of Iowa, October 2003.
20. Paterson, E.G., and Peltier, L.J., Detached-Eddy
Simulation of High-Reynolds Number BeveledTrailing-Edge Flows and Wakes, Symposium on
LES Advancements and Applications, ASME FED
Summer Meeting, Charlotte, NC, July 2004.
21. Paterson, E.G., Poremba, J.E., Peltier, L.J., and
Hambric, S.A., A Physics-Based Simulation
Methodology for Predicting Hydrofoil Singing,
25th Symposium on Naval Hydrodynamics, St.
Johns, Newfoundland and Labrador, Canada,
August 2004.
22. Paterson, E.G., and Baker, W.J., RANS and
Detached-Eddy Simulation of the NCCR Airfoil,
2004 NASA-ONR Circulation Control Workshop,
Hampton, VA, March 16-17, 2004.
23. Ran, B., and Katz, J., Pressure Fluctuations and
Their Effect on Cavitation Inception Within Water
Jets, Journal of Fluid Mechanics, Vol. 262, pp.
223-263, 1994.
24. Strelets, M., Detached Eddy Simulation of
Massively Separated Flows, AIAA 2001-0879,
39th AIAA Aerospace Sciences Meeting and
Exhibit, Reno, NV, Jan 2001.
25. Suhs, N.E., Dietz, W.E., Rogers, W.E., Nash, S.M.,
and Onufer, T., PEGASUS Users Guide, Version
5.1e, Technical Report, NASA, 2000.

14

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics


St. Johns, Newfoundland and Labrador, CANADA, 8-13 August 2004

Numerical Study of Cavitation Inception due to


Vortex/Vortex Interaction in a Ducted Propulsor
C.-T. Hsiao and G. L. Chahine (DYNAFLOW, INC., USA)

ABSTRACT
Cavitation inception in a ducted propulsor
was numerically studied using Navier-Stokes
computations and bubble dynamics models. Previous
experimental observations and RANS computations
indicated the presence of two interacting vortices in
the region where cavitation inception occurred. A
direct numerical simulation with initial and boundary
conditions provided from the RANS solution of a full
ducted propulsor flow was conducted in a reduced
computational domain in the wake region in order to
improve the numerical solution of the liquid flow.
Bubbles were then released in this flow field, and
bubble dynamics models including spherical and nonspherical models were applied to study cavitation
inception. The numerical results were compared to
experimental measurements and observations. Good
agreement, far superior to that obtained by RANS
alone, was found in terms of cavitation inception
number and inception location as well as the
characteristics of acoustic signals and bubble shapes
during a cavitation event.

1. Introduction
Prediction of vortex cavitation inception on
marine propulsors is of great interest to the Navy and
has been the subject of many studies in recent years in
order to derive scaling laws for the prediction of
cavitation inception. However, these scaling laws, are
typically formulated based on data from open
propellers and may not be applicable to a ducted
propulsor. Unlike most open propellers, which
generally have an elliptical shape and form a single
trailing vortex, a ducted propulsor typically forms two
well-defined vortices in the tip region. In addition to a
trailing vortex formed near the tip trailing edge, a
much stronger tip-leakage vortex is generated in the
gap region between the shroud wall and the blade tip.
These two unequal co-rotating vortices introduce
small-scale unsteady motions during vortex merging

that are in addition to upstream turbulent fluctuations


and vortex wandering (Chen et al. 1999, Devenport et
al. 1999).
Recent
experimental
observations
of
cavitation inception on a ducted propulsor (Chesnakas
and Jessup 2003) have indicated that the interaction
between the tip-leakage vortex and the trailing-edge
vortex may cause cavitation inception to occur in the
region where the two vortices merge. However,
predictions of cavitation inception using the pressure
field either inferred from experimental measurements
(Oweis et al. 2003) or obtained by Reynolds-Averaged
Navier-Stokes (RANS) computations (Brewer et al.
2003, Yang 2003) are in poor agreement with the
experimental observations in terms of cavitation
inception number and inception location.
A preliminary controversial conclusion made
by Chesnakas and Jessup was that cavitation inception
does not occur near the minimum pressure region.
This conclusion, however, was drawn based on the
inferred pressure field obtained from the measured
average tangential velocity field using a Rankine
vortex assumption. This method not only neglected the
axial velocity effect on the pressure field, but also
relied on a time-averaged tangential velocity which
could significantly smear out fluctuations due to the
vortex wandering, especially at downstream locations.
Furthermore, the inferred pressure in the vortex core
cannot explain the shape and extent of the fully
developed cavitation vortex in the vortex core
observed at lower cavitation numbers (see Figure 6).
RANS computations with inadequate
turbulence models and grid resolution are also known
to cause over diffusion and dissipation in the vortex
flow (Dacles-Mariani et al. 1995, Hsiao and Pauley
1998). This usually leads to a significant
underprediction of the velocities in the vortex core at
downstream locations. In a combined numerical and
experimental study of a tip vortex flow, DaclesMariani et al. (1995) used the measured flow field to
specify the inflow and outflow boundary conditions

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

and investigated the vortex preservation in the wake


region. With the turbulence model turned off and
significant grid refinement they were able to match the
numerical solution to the experimental measurements.
Besides the flow field which is not well
resolved, the effect of bubble dynamics on the
cavitation inception has also not been fully addressed.
Previous studies (Hsiao and Chahine 2003b, 2003c,
2004) have shown that inclusion of bubble dynamics
significantly affects the prediction of cavitation
inception for a steady-state tip vortex flow as well as
for an unsteady vortex/vortex interaction flow field.
In the current study we aim to improve the
numerical prediction of cavitation inception for a
ducted propulsor in two ways. First, a reduced
computational domain is considered which excludes
propulsor solid surfaces to reduce geometric
complexity and only encompasses the region of
interaction of the two vortices. A direct numerical
simulation is conducted for this reduced
computational domain but with initial and boundary
conditions provided by the RANS computation of the
full ducted propulsor flow field. Second, a one-way
coupled spherical bubble dynamics model developed
by Hsiao and Chahine (2003b,c) and a two-way
coupled non-spherical bubble dynamics model
developed by Hsiao and Chahine (2004) are applied
to study bubble dynamics and to predict cavitation.

the cross flow plane, 61 61 , 121 121 and 181 181 .


All grid points are evenly distributed without
stretching. This results in a uniform grid size of 3mm
in the streamwise direction and 0.5mm for the finest
grid in both cross directions. At least 34 grid points are
within the vortex core in each direction for the finest
grid since the vortex core size is about 17mm in
diameter at the trailing edge.

Figure 1. A view of the reduced computational


domain used for the current computations.

2. Numerical Approach
2.1 Flow Configuration
We consider the David Taylor Propeller
5206, a rotating ducted propulsor, which is a threebladed propeller with a constant chord of 0.3812m
from hub to tip and a tip diameter of 0.8503m and
operates in a duct of diameter 0.8636m. The detailed
propulsor geometry can be found in Chesnakas and
Jessup (2003). There have been three numerical
studies (Kim 2002, Brewer et al. 2003 and Yang 2003)
applying RANS codes to obtain a time-averaged flow
field for this ducted propulsor. They all give
reasonable agreement with the experimental
measurements.
We construct a reduced computational
domain behind the trailing edge of the propulsor blade
that encompasses the region of interaction of the two
vortices. This computational domain has a square
cross area of 0.094m 0.094m and extends from the
tip trailing edge to the downstream location 0.34m
from the tip trailing edge. Figure 1 illustrates the
location of the reduced computational domain relative
to the ducted propulsor. We consider a 4-block grid
system with 101 grid points in the streamwise
direction and three different numbers of grid points in

Figure 2. The interpolated pressure field of the


reduced computational domain.
To conduct our numerical computations in
this reduced domain, the solution of a RANS
computation obtained by Yang (2003) is interpolated
to provide the initial conditions at the grid points of
the reduced domain. We consider the case of an
advance coefficient, J=0.98, with an inflow velocity,
U = 6.96 m/s . This results in a Reynolds number

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

based on the blade tip radius and the inflow velocity of


Re = 3 106 . Figure 2 shows the interpolated pressure
contours at different streamwise locations to indicate
the position of the main vortex in the reduced domain.
Figure 3 shows the pressure contour and velocity
vectors at the inlet boundary on the x-r plane. The two
co-rotating vortices (the tip-leakage vortex and the
trailing edge vortex) can be readily seen. The strength
of the tip-leakage vortex is much larger than that of the
trailing-edge vortex.

Figure 3. The pressure contours and velocity vectors at


the inlet boundary on the x-r plane.
2.2 Navier-Stokes Computations
For the vortex interaction study in the
reduced domain, the flow is obtained via direct
numerical simulation of the Navier-Stokes Equations
without turbulence modeling. Since the present
computation is conducted on a rotating frame attached
to the rotating propeller blade, the steady-rotating
reference frame source terms, i.e. the centrifugal force
and the Coriolis force terms, are added to the
momentum equation. The resulting unsteady
incompressible
continuity
and
Navier-Stokes
equations written in non-dimensional vector form and
Cartesian notations are given as
u = 0 ,
(1)
Du
1 2
= p +
u + 2 r 2 u ,
(2)
t
Re
where u = (u, v, w) is the velocity, p is the
pressure, r is the radial position vector, is the
angular velocity, Re = u * L * / is the Reynolds
number, u* and L* are the characteristic velocity and
length, is the liquid density, and is its dynamic
viscosity.
To solve Equations (1) and (2) numerically, a
three-dimensional incompressible Navier-Stokes
solver, DF_Uncle, developed at Mississippi State
University and modified by DYNAFLOW, INC. is

applied. DF_Uncle is based on the artificialcompressibility method (Chorin 1967) in which a


time derivative of pressure is added to the continuity
equation as
1 p
+ u = 0 ,
(4)
t
where is the artificial compressibility factor. As a
consequence, a hyperbolic system of equations is
formed that can be solved using a time marching
scheme. This method can be marched in pseudo-time
to reach a steady-state solution. To obtain a timedependent solution, a Newton iterative procedure
needs to be performed at each physical time step in
order to satisfy the continuity equation. In the present
study the time-accurate solution was obtained when
the maximum normalized velocity divergence was
less than 1.010-3. Detailed descriptions of the
numerical scheme can be found in Vanden and
Whitfield (1993).
The boundary conditions for this reduced
domain are also deduced from the RANS solution.
The initial values of the pressure and velocities
interpolated from the RANS solution are imposed at
all boundaries except the inlet and outlet boundaries.
At the inlet boundary the method of characteristics is
applied with all three components of velocities
specified from the RANS solution. For the outlet
boundary all the variables are extrapolated from the
inner grid points but with the initial value of the
pressure fixed at one grid point.
2.3 Bubble Dynamics Models
Two bubble dynamics models, a spherical
model and a non-spherical model, are applied in this
study. In the spherical bubble dynamics model each
bubble is tracked by a Lagrangian scheme in the flow
field which combines the RANS solution and the
current DNS solution by oversetting the grid of the
reduced domain with the overall propulsor grid. As a
bubble is released upstream of the reduced domain
the flow field from the RNAS solution is used. Once
the bubble enters the reduced domain, the flow field
obtained from the current simulation is applied.
Bubble transport is modeled via the motion equation
described by Johnson and Hsieh (1966) while the
bubble dynamics is simulated by solving a surface
Averaged Pressure (SAP) Rayleigh-Plesset equation
developed by Hsiao and Chahine (2003a,b).
The non-spherical bubble dynamics model is
embedded in the unsteady Navier-Stokes solver, DFUNCLE, with appropriate free surface boundary

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Pressure coefficient along vortex center


-4

DNS 61x61 grid


DNS121x121 grid
DNS 181x181 grid
RANS

-6

Cp

conditions and a moving Chimera grid scheme. Since


unsteady Navier-Stokes computations are timeconsuming, this non-spherical model is combined with
the spherical model mentioned above. The spherical
model is used to track the bubble during its capture by
the vortex and the non-spherical model is turned on
only when the bubble size exceeds a preset limit value.
When the non-spherical model is turned on, the flow
field due to the spherical bubble motion and volume
change is superimposed on the liquid phase flow field
solution to provide an initial condition for the
unsteady viscous computation. This model allows the
bubble to deform non-spherically and a full two-way
interaction between the bubble and the flow field can
be obtained. Detailed description of this model and
numerical implementations can be found in Hsiao and
Chahine (2001,2004).

-8

-10

0.1

0.2

0.3

s/C

0.4

0.5

0.6

0.7

Figure 4. Comparison of the pressure coefficient


variation along the vortex center for three different
DNS grids and for the RANS solution.

3. Results and Discussion


3.1 Liquid Flow
The simulation of vortex interaction in the
reduced computational domain is conducted with the
turbulence model turned off. The mean flow solution
is specified at the inlet boundary. Additional unsteady
turbulent fluctuations from upstream will be
simulated in future efforts. The two-vortex interaction
is then simulated for different discretizations.
Three levels of grid resolution as described
in Section 2.1 were tested. Although unsteady
computations were conducted for this study, all three
cases converged to a practically steady-state solution.
The instability due to strong vortex/vortex interaction
as shown in Hsiao and Chahine (2003c) is not
observed in the current simulation. This is probably
due to a relatively weak trailing-edge vortex. Hsiao
and Chahine have shown that the interaction between
the two co-rotating vortices becomes weaker as the
relative strength of the main vortex is increased. To
further resolve the instability due to a weak
interaction, further grid refinement may be required.
Figure 4 shows a comparison of the resulting
pressure coefficient, Cp, along the vortex center line
for these three cases. It is seen that as the grid is
refined, Cpmin approaches about -11 at a location 0.35
chord length downstream from the tip trailing edge.
The solutions of the 121121 and 181181 grids are
quite close. Since the 121121 grid only yields a
small difference in the minimum pressure as
compared to the finest grid, this grid solution was
used for subsequent bubble dynamics computations
for the sake of CPU time reduction.

Figure 5. Iso-pressure surfaces equivalent to


cavitation extent at various cavitation numbers as
obtained by the RANS solution and the current DNS
solution with the 121121 grid.
The pressure coefficient along the vortex
center obtained with the 121121 grid is also
compared with the RANS solution and shown in
Figure 4. Major fundamental differences are seen
between these two results. The RANS computation
predicts Cpmin=-8.2 at s/C=0.1 while the current
simulation shows Cpmin=-11 at s/C=0.35. This is
probably due to excessive vortex diffusion and
dissipation. The comparison is also made by showing
various iso-pressure surfaces in Figure 5. This is

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

similar to visualizing a cavitating vortex at different


cavitation numbers. The current results seem to agree
with the experimental observation much better
because the experimental videos show a longextended fully cavitating vortex core at = 5.6 (see
Figure 6), and also indicate cavitation inception at
10.8 at about 0.35 chord length downstream of the
trailing edge.

Figure 6. Fully developed cavitation in the vortex


core at = 5.6 (Chesnakas and Jessup 2003).
It is important to examine the flow field near
the location where the pressure reaches the minimum
value. In our previous study (Hsiao and Chahine
2003c), we showed that the two-co-rotating vortices
periodically approach each other during the vortex
merger. As they move closer, the flow in the axial
direction is accelerated and results in a decreased
pressure in the vortex center. It is found that as the
axial velocity reaches a maximum value, the pressure
in the vortex center will drop to its minimum. This is
also observed in the current simulation. The
computed Cp and axial velocity along the vortex
center line are shown in Figure 7.

3.2 Area of Bubble Capture: Window of


Opportunity
The window of opportunity through which
a nucleus needs to enter to be captured by the vortex
and generate strong acoustic signals has been studied
for a finite-span tip vortex flow by Hsiao and Chahine
(2003b). It is also important to determine this
window of opportunity for the current flow field
because with the knowledge of the location and size
of this small window, we are able to distribute and
follow nuclei more efficiently. Near inception the size
of the window of opportunity is strongly related to
the probability of the cavitation events.
To establish the window of opportunity a
rectangular release area was specified ahead of the tip
leading edge of the propulsor on the x-r plane with
165 nuclei of a given size released from a 1511 grid
point array. Figure 8 illustrates the location of the
release area related to the propulsor blade. The
cavitation number was specified high enough (=12)
such that the maximum growth size of a nucleus was
less than 10 %. Each nucleus was tracked and the
minimum pressure coefficient it encountered during
its travel was recorded and assigned to the release
grid point. This enables us to plot a contour of the
minimum encountered pressure coefficient for the
release grid points and to obtain the window of
opportunity for each case.
Figure 9 shows contours of minimum
encountered pressure coefficient for four different
nuclei sizes, R0 = 5, 10, 20 microns. The contours are
blanked out for the release points where the nucleus
collides with the propeller surface. It is seen that the
size of the window of opportunity becomes smaller
and its location shifts closer to the propeller pressure
side surface when the nuclei size decreases.

3.5

-8

3.3

Vs

Cp

3.4

Cp
Axial velocity

-9

3.2

-10

3.1

0.1

0.2

0.3

0.4

s/C

0.5

0.6

0.7

Figure 7. Pressure coefficient and axial velocity as a


function of the distance from the tip trailing edge.

Figure 8. The location of the release area for


establishing the window of opportunity.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Figure 9. Contours of encountered Cpmin for nuclei


with R0=5, 10, and 20 microns.
3.3 Single Bubble Dynamics for Prediction of
Cavitation Inception
Experiments conducted by Chesnakas and
Jessup (2003) with a high speed video camera and a
sensitive hydrophone captured the bubble and its
emitted acoustic signal during sub-visual cavitation
events. According to the duration of the acoustic
signal, the cavitation events were categorized into
popping and chirping events. They stated that the
popping event has a very short duration of
noticeable acoustic signal less than 0.3ms and that the
bubble virtually remained spherical when its size was
less than 0.1mm in diameter. The chirping event has
a much longer duration ranging from 0.3 to 10ms, and
the bubble has an elongated shape. They found that all
the cavitation inception events occurred near or behind
a location 0.5 chord length downstream of the tip
trailing edge.
To simulate the cavitation events, we
investigated the bubble behavior and the emitted
acoustic signal for different initial nuclei sizes at
different cavitation numbers.
We found that
popping cavitation events can be observed at a
cavitation number just slightly smaller than the
negative minimum pressure coefficient, -Cpmin=11.

Figure 11 shows the bubble size variation and emitted


acoustic signals for an initial nucleus size, R0=20m at
=10.85. It is seen that the maximum bubble size is
about 0.1mm in diameter and the noticeable acoustic
signal only lasts about 0.3ms. As the cavitation
number is reduced the bubble grows to a much larger
size and the duration of the acoustic signal is much
longer as shown in Figure 12 for R0=20m at
=10.75.
Figures 11 and 12 also show the pressure
encountered by the bubble during its journey. There is
a small delay in time for the bubble to grow to its
maximum size after encountering the minimum
pressure. This delay significantly increases when a
cavitation event is produced by a small size nucleus.
An example of such a cavitation event is shown in
Figure 13 for R0=5m at =10.3. To illustrate where
the cavitation event occurs in the flow field, the bubble
trajectory and size variations are plotted with the
propulsor blade and iso-pressure surface as shown in
Figure 14. It is seen that for the larger R0 the cavitation
event occurs at a location slightly earlier than the
experimental observation while the smaller R0 grows
to its maximum size near a location 0.5 chord length
downstream of the tip trailing edge.
The bubble dynamics is also studied with the
non-spherical model. Figure 15 compares the bubble
shapes obtained with the spherical and the nonspherical model for R0=20m at =10.75. It is seen
that both models predict almost the same maximum
growth size. The non-spherical model also shows that
the bubble elongates in the axial direction and
becomes a cylindrical shape as it grows. However, for
R0=20m at =10.85 the bubble remains almost
spherical at its maximum size as shown in Figure 16.
For both cases the bubble starts to collapse after
reaching its maximum size. The non-spherical
computations, however, fail to continue once strong
deformations develop over the bubble surface during
the collapse.

Figure 11. The bubble radius, emitted acoustic


pressure signal and encountered pressure during a
cavitation event for R0= 20m at =10.85.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Spherical Model

Non-Spherical Model

Figure 12. Bubble radius, emitted acoustic pressure


and encountered pressure during a cavitation event for
R0= 20m at =10.75

Figure 15. Computed bubble sizes and shapes of both


spherical and non-spherical modes plotted with two
levels of iso-Cp for R0= 20m at =10.75.

Figure 13. Bubble radius, emitted acoustic pressure


and encountered pressure during a cavitation event for
R0= 5m at =10.3.
R0=20m
=10.85

R0=20m
=10.75

Figure 16. Computed bubble sizes and shapes of nonspherical modes for R0= 20m at =10.85
s/C=0
Cp=-10.9

s/C=0

Cp=-5.6

Cp=10 9

Cp=-5.6

s/C=0.5

s/C=0.5

blow up

blow up

R0=5m
=10.3

s/C=0
Cp=-5.6
Cp=-10.9
s/C=0.5

Figure 14. The bubble trajectories and size variations


during the cavitation event for three cases.

3.4 Multiple Bubble Dynamics for Prediction of


Cavitation Inception
In order to simulate a real nuclei flow field as
exists in nature or in the waters of a cavitation tunnel,
Hsiao and Chahine (2004b) used a statistical nuclei
distribution model and showed that the nuclei size
distribution has a strong influence on the prediction of
cavitation inception. Since the nuclei size distribution
is not available for the experiments conducted on the
ducted propulsor flow, we have selected two very
different nuclei size distributions and compared their
effect on the prediction of cavitation inception. The
first nuclei size distribution contains larger nuclei sizes
ranging from 2.5 to 25m while the other one contains
smaller nuclei sizes ranging from 2.5 to 10m (see
Figure 17). In both cases we randomly released the
nuclei from a 0.02m0.03m window. A total of 600
nuclei were released within 0.4 second. The nuclei size
distribution for the both cases is shown in Figure 17.
As the nuclei travel in the computational
domain, the resulting acoustic signals are monitored.
The acoustic pressure is monitored on the shroud wall
at a location 0.5 chord length downstream of the tip

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

trailing edge. A series of computations were conducted


at different cavitation numbers for both nuclei
distributions to obtain acoustic signals for conditions
above and below cavitation inception. Figure 18
illustrates the acoustic signals for the larger nuclei size
distribution at three different cavitation numbers.
600
500
Nuclei Distribution 1
Nuclei Distribution 2

Number

400
300
200

100
0
2.5E-06

5.0E-06

7.5E-06

1.0E-05

1.5E-05

2.0E-05

2.5E-05

Nuclei Size (m)

Figure 17. The number of nuclei released versus nuclei


size for two different nuclei size distributions
considered in this study.
Cavitation Number =10.90
Void fraction =9.0e-10

Acoustic Press (pa)

600

400

From the results shown in Figures 18, we can


define a cavitation inception number based on the
number of cavitation events per unit time exceeding a
certain value. Here a cavitation event is defined
arbitrarily when a cavitating bubble emits an acoustic
signal higher than 100pa. The curve of the number of
cavitation events per second versus cavitation number
is shown in Figure 19. It can be seen that there is a
critical cavitation number above which no cavitation
events occur. For nuclei size distribution No. 1 (larger
bubbles) an abrupt rise in the number of cavitation
events is seen when the cavitation number is below the
critical cavitation number. Based on these curves one
can determine the cavitation inception number for both
cases by defining a criterion. For example, if 10 events
per second is defined for cavitation inception, then a
cavitation inception number i=10.89 for the larger
nuclei size distribution and i=10.6 for the smaller
nuclei size distribution can be deduced from Figure
19. Chesnakas and Jessup (2003) defined the
cavitation inception criterion as one event per second
and obtained a cavitation inception number about 11.
This inception number is very close to the critical
cavitation number (10.9) for the larger nuclei size
distribution, but these results are subject to the two
criteria selected: amplitude of the peak and number of
peaks per unit time.

200

80
0.1

0.2

Time (Sec)

0.3

0.4

Number of events per second

Cavitation Number =10.85


Void fraction =9.0e-10

Acoustic Press (pa)

600

400

70

Nuiclei Distribution 1
Nuclei Distribution 2

60
50
40
30
20
10

200

0
10.2

10.3

10.4

10.5

10.6

10.7

10.8

10.9

11

Cavitation Number
0

0.1

0.2

Time (Sec)

0.3

Figure 19. Number of events per second versus


cavitation number for the nuclei size distributions
shown in Figure 17.

0.4

Cavitation Number =10.80


Void fraction =9.0e-10

Acoustic Press (pa)

600

4. Conclusions

400

200

0.1

0.2

Time (Sec)

0.3

0.4

Figure 18. The acoustic signals for the large size


nuclei distribution case at three different cavitation
numbers.

A direct numerical simulation of the two


interacting vortices in a ducted propulsor flow field
was conduced in a reduced computational domain to
address the grid resolution issue in RANS
computation. It was found that vortex diffusion and
dissipation were significantly reduced with grid
refinement. The resulting solutions illustrated with
iso-pressure surfaces agree much better than RANS
computations with experimental observations for
fully developed cavitation in the vortex core and for
cavitation inception number and location. No

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

instability was seen due to a weak vortex/vortex


interaction between the tip-leakage vortex and the
trailing-edge vortex in the simulations. Further grid
refinement may be required to resolve any such
instability.
The location and size of the window of
opportunity through which a nucleus needs to enter to
be captured by the vortex was identified for different
nuclei sizes.
From the study of single bubble dynamics we
showed that the characteristics of the acoustic signals
and bubble shapes as well as the location of cavitation
inception resemble those observed experimentally. A
multiple bubble dynamics model was also applied to
study the effect of nuclei size distribution and to
predict cavitation inception in real flow field
conditions. Different nuclei size distributions and
definitions of the cavitation inception event were
found to influence the cavitation inception number.
However, the range of cavitation inception number
(i11) was found to agree much better than previous
studies (i 5) with the experimental measurements.
ACKNOWLEDGMNETS
This work was conducted at DYNAFLOW,
INC. (www.dynaflow-inc.com) and was supported by
the Office of Naval Research under contract No.
N00014-04-C-0110 monitored by Dr. Ki-Han Kim.
The RANS solution of the full propulsor flow was
provided by C. I. Yang from NSWCD and the
experimental data was provided by Christopher J.
Chesnakas from NSWCD. Their cooperation on this
study is greatly appreciated.

REFERENCES
[1] Brewer, W.H., Marcum, D.L., Jessup, S.D.,
Chesnakas, C., Hyams, D.G., Sreenivas, K., An
Unstructured RANS Study of Tip-Leakage
Vortex Cavitation Inception, Proceedings of the
ASME Symposium on Cavitation Inception,
FEDSM2003-45311, Honolulu, Hawaii, July 610, 2003.
[2] Chen, A.L., Jacob, J.D., Savas, O., Dynamics of
Co-rotating Vortex Pairs in the Wakes of
Flapped Airfoils, J. Fluid Mech., vol. 382, 1999,
pp. 155-193.
[3] Chesnakas, C.J., Jessup, S.D., Tip-Vortex
Induced Cavitation on a Ducted Propulsor,
Proceedings of the ASME Symposium on
Cavitation
Inception,
FEDSM2003-45320,
Honolulu, Hawaii, July 6-10, 2003.

[4] Chorin, A. J., A Numerical Method for Solving


Incompressible Viscous Flow Problems, Journal
of Computational Physics, Vol. 2, 1967, pp. 1226.
[5] Dacles-Mariani, J., Zilliac, G.G., Chow, J.S.,
Bradshaw, P., Numerical/Experimental Study of
a Wingtip Vortex in the Near Field, AIAA
Journal, Vol. 33, No. 9, 1995, pp.1561-1568.
[6] Devenport, W.J., Vogel, C.M., Zsoldos, Flow
Structure Produced by the Interaction and Merger
of a Pair of Co-Rotating Wing-Rip Vortices, J.
Fluid Mech., vol. 394, 1999, pp. 357-377.
[7] Hsiao, C.-T., Pauley, L.L., Numerical Study of
the Steady-State Tip Vortex Flow over a FiniteSpan Hydrofoil, ASME Journal of Fluid
Engineering, Vol. 120, 1998, pp. 345-349.
[8] Hsiao, C.-T., Chahine, G.L., Liu, H.L., Scaling
Effects on Prediction of Cavitation Inception in
a Line Vortex Flow, ASME Journal of Fluids
Engineering, Vol. 125, 2003a, pp. 53-60.
[9] Hsiao, C.-T., Chahine, G.L., Scaling of Tip
Vortex Cavitation Inception Noise with a
Statistic Bubble Dynamics Model Accounting for
Nuclei Size Distribution, Proceedings of the
ASME Symposium on Cavitation Inception,
FEDSM2003-45315, Honolulu, Hawaii, 6-10
July, 2003b.
[10] Hsiao, C.-T. and Chahine, G.L., Effect of
Vortex/Vortex Interaction on Bubble Dynamics
and Cavitation Noise, Fifth International
Symposium on Cavitation CAV2003, Osaka,
Japan, November 1-4, 2003c.
[11] Hsiao, C.-T., Chahine, G.L., Prediction of
Vortex Cavitation Inception Using Coupled
Spherical and Non-Spherical Models and NavierStokes Computations, Journal of Marine Science
and Technology, Vol. 8, No. 3, 2004, pp. 99-108.
[12] Johnson, V.E., Hsieh, T., The Influence of the
Trajectories of Gas Nuclei on Cavitation
Inception, Sixth Symposium on Naval
Hydrodynamics, 1966, pp. 163-179.
[13] Kim, J., Sub-Visual Cavitation and Acoustic
Modeling for Ducted Marine Propulsor, Ph.D.
Thesis, 2002, Department of Mechanical
Engineering, The University of Iowa, Adviser F.
Stern.
[14] Oweis, G., Ceccio, S., Cheskanas, C. Fry, D.,
Jessup, S., Tip Leakage Vortex (TLV)
Variablity from a Ducted Propeller under Steady
Operation and its Implications on Cavitation
Inception, 5th International Symposium on
Cavitation, Osaka, Japan, November 1-4, 2003.
[15] Yang, C.I., Jiang, M., Chesnakas, C.J., and
Jessup, S.D., 2003, "Numerical Simulation of Tip

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Vortices of Ducted-Rotor", NSWCCD-50-TR2003/46


[16] Vanden, K., Whitfield, D. L., Direct and
Iterative Algorithms for the Three-Dimensional
Euler Equations, AIAA-93-3378, 1993.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Christopher Chesnakas
Naval Surface Warfare Center, Carderock Division,
USA
The work is an interesting & valuable
contribution to our understanding of the tip-vortex
cavitation inception problem. You state that your
calculations now well match the experiments. There
are, however, some observations in the measurements
that I dont see reproduced in your calculations.
First, the experiment shows the cavitation events to
increase in number very gradually as cavitation
number is lowered. Your calculations seem to show
the bubble events to increase rather abruptly.
Second, the experiments show that as is lowered,
the character of the bubble cavitation events
gradually changes. At high , only single bubble
cavitation is observed. These events, pops, are
very short in duration (<1ms) and produce a very
short acoustic signal (<0.5ms). At lower , elongated
cavitation bubbles, and small bubble trails are
observed, and the acoustic character of the noise
signal changes. The bubble trails may last for several
ms, and the acoustic signal is much longer (>1ms)
and has a totally different character. These chirps
are quite distinct from the pops, and begin to occur
when the event rate is still quite low. These
phenomena, I believe, are not yet fully described.

and better experimental information is needed to


compare with the corresponding numerical
simulations, i.e. initial bubble size, better
visualization of the bubble dynamics, and better
signal capture and synchronization.

Figure a. Bubble behaviour in a vortex flow field at


low cavitation number. Three-dimensional view of
the bubble just before splitting predicted by
2DYNAFS.

AUTHORS REPLY
The increase in the rate of cavitation events
when the cavitation number is reduced depends on
the nuclei size distribution. The calculations
presented in this paper did not always show the
bubble events to increase abruptly. As shown in
Figure 19, the number of cavitation events for the
nuclei distribution #2 case increases much more
gradually than that of the nuclei distribution #1 case
as the cavitation number is reduced. To make good
comparison between numerical result and
experimental measurement, a reliable nuclei size
distribution needs to be provided.
Our computations capture both what you
call pop events and chirp events. This paper was
not concerned with showing these or to address nonspherical bubble elongation, even though we have
extensively addressed this elsewhere [a-c]. Figure 11
in the paper shows the short duration sharp signals or
pops. Figure a in this response shows how the
simulations also show that bubbles could
significantly elongate at low cavitation numbers.
Figure b shows the corresponding longer duration
oscillations or chirp. This is obviously qualitative,

Figure b. Corresponding long duration of


acoustic signals resembling chirp.
[a] Hsiao, C.-T., Chahine, G. L., Prediction of
Vortex Cavitation Inception Using Coupled Spherical
and Non-Spherical Models and Navier-Stokes
Computations, Journal of Marine Science and
Technology, Vol. 8, No. 3, pp. 99-108, 2004.
[b] Choi, J.-K., Chahine, G.L., Noise due to
Extreme Bubble Deformation near Inception of Tip
Vortex Cavitation, Physics of Fluids, Vol.16, No.7,
July 2004.
[c] Choi, J.-K and Chahine, G. L., Non-Spherical
Bubble Behavior in Vortex Flow Fields,

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Computational Mechanics, Vol. 32, No. 4-6, pp.281290, December 2003

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics


St. Johns, Newfoundland & Labrador, CANADA, 8-13 Aug 2004

Bubble drag reduction at large scales and high Reynolds


numbers
Wendy Sanders, Jinhyun Cho, Eric Winkel, Elizabeth Ivy, Robert Etter1
David Dowling, Marc Perlin, Steve L. Ceccio
(Department of Mechanical Engineering, University of Michigan, USA)
(1Carderock Division, Naval Surface Warfare Center, USA)

ABSTRACT

INTRODUCTION

Skin friction in liquid flows may be reduced


when bubbles are present near solid surfaces.
Potential applications may occur at downstream
distance-based Rex on the order of 1011 while
laboratory experiments to date have achieved only
several million. This paper presents results for Rex as
high as 210 million from skin-friction drag reduction
experiments conducted in the US Navy's William B.
Morgan Large Cavitation Channel (LCC). A nearzero-pressure-gradient flat-plate turbulent boundary
layer was generated on a 12.9-m-long hydraulically
smooth flat plate that spanned the test section. The
test surface faced downward and air was injected at
volumetric rates to 0.38 m3s1 through one of two
flush-mounted 40-micron sintered-metal injectors
that spanned the test model at upstream and
downstream locations. Spatially- and temporallyaveraged shear stress and impedance-based void
fraction measurements are reported. For the measured
flow conditions, the mean bubble diameter was ~300
microns. At the lowest freestream speed (6 ms1),
buoyancy lifted the air bubbles to the plate surface
where they formed a quasi-continuous gas film that
lead to essentially 100% skin-friction drag reduction
to the farthest downstream measurement location. At
higher speeds, the bubbles remained distinct and
skin-friction drag reduction was observed when the
+
bubbly mixture was within 300 lo of the plate surface

Three techniques that involve a quasi-steady


energy expenditure at constant speed to achieve skin
friction reduction are suction (or blowing) through
the surface; injection of a polymeric solution from
the surface; and injection of gas bubbles through the
surface. This final technique, referred to as BDR (for
bubble-induced skin-friction drag reduction), is
discussed herein.
Prior BDR experiments (see e.g. Merkle and
Deutsch 1992) have shown that this technique can
reliably produce drag reductions of greater than 50%,
and as high as 80% to 90%, at Rex of several million
on flat plates of length about 1 meter. Hence air
injection is a promising technique to achieve drag
reduction for commercial and military watercraft if it
remains effective at much higher Reynolds numbers,
Rex ~ 109 to 1010, and for much longer distances. This
paper presents the results of an experimental study
that reduces significantly the Rex and length
differences between model and prototype scales.
This paper is organized as follows: a review
of prior BDR studies; a description of the setup,
measured quantities, and the experimental program; a
description of the impedance void fraction
measurement technique; a presentation of the main
results (i.e. drag reduction due to air injection, and
attendant
impedance-based
void
fraction
measurements); and a final section that summarizes
and presents conclusions.

even when the bubble diameter was more than 100 lo


+

where lo is the viscous wall unit of the equivalent


boundary layer without air injection. Skin-friction
drag reduction did not persist when shear-induced
bubble migration forced bubbles farther from the
plate. The measurements suggest that a combination
of density reduction and turbulence modification is
responsible for bubble-based skin-friction drag
reduction.

REVIEW OF PRIOR BDR STUDIES


McCormick & Bhattacharyya (1973)
reported the first successful BDR experiments at
Reynolds numbers (based on hull length) to 1.8x106.
About the same time, Soviet researchers (Migirenko
& Evseev 1974; Bogdevich & Malyuga 1976;
Bogdevich & Evseev 1976) observed that the
maximum level of drag reduction occurred

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

immediately downstream of gas injection and did not


persist with increasing downstream distance.
Beginning in the early 1980s, BDR has been studied
extensively at the Applied Research Laboratory at the
Pennsylvania State University. Madavan et al. (1984)
made BDR measurements in a zero-pressure-gradient
turbulent boundary layer at Rex as large as 107. The
same researchers also investigated the influence of
the pore size of the air injector (Madavan et al.
1985a) and found BDR to be relatively independent
of injector pore size from 0.5 to 100 microns.
Deutsch & Castano (1986) showed that BDR on an
axisymmetric body improved with increasing flow
speed as the influence of bubble buoyancy was
reduced. Fontaine & Deutsch (1992) used five
different gases with varying densities to show that
BDR did not depend on the chemical composition of
the injected gas. Using a flat plate, Pal et al. (1988)
found that bubble size did not affect BDR for bubble
radii from 75 to 600 microns. Clark & Deutsch
(1991) analyzed the effects of adverse and favorable
pressure gradients on BDR using an axisymmetric
body, and found that a weak favorable pressure
gradient significantly inhibited drag reduction, while
a weak adverse pressure gradient caused separation
and large reductions in the skin friction drag at low
gas injection rates. A comprehensive review of the
first two decades of BDR research is provided by
Merkle & Deutsch (1992). Recently, BDR research
has also been conducted in Japan. Kodama et al.
(2000) presents BDR results from a sea trial.
Takahashi et al. (2001) examined BDR on a flat
surface at Rex to 25 million and found that variations
in bubble size and boundary layer thickness did not
significantly influence the level of drag reduction.
Kawamura et al. (2003) found a similar insensitivity
to bubble size for bubble radii of 250 to 1000
microns at similar Rex values. The new higherReynolds-number results presented here confirm
many of these prior findings while providing unique
results for BDR persistence.
A brief review of the outer-variable scaling
for BDR is necessary for our discussions later. Skin
friction between a solid surface and a flowing liquid
is commonly reported in normalized form as the skin
friction coefficient, C F , defined as CF = w where
1
U 2
2

the usual way: u* = w , with uo being the


friction velocity of the flow without air injection.
The skin friction ratio depends on the
volumetric rate of air injection, Qa, flow speed U,
x,
boundary
layer
downstream
distance
characteristics, and possibly injector and bubble
characteristics. Madavan et al. (1984), as well as
other researchers, showed a correlation of drag
reduction data when the non-dimensional gas-flow
rate is given by Q a SU , where S is a reference area,
typically the wetted surface of the porous bubble
injector. Madavan et al. (1985a) introduced a
different normalization of the gas injection rate,
Qa/(Qa+Qw) where Qw is the volume flux of water in
the boundary layer when air is not injected;
Qw = U o o* b . and * are the boundary layers
99% and displacement thicknesses, respectively,
without air injection. For hydraulically smooth
surfaces, Deutsch et al. (2003) proposed yet another
scaling parameter, Qa (Qa + Ubo ) , using the single
liquid phase boundary layer momentum thickness, .
There have been many analytical and
numerical efforts to determine and understand BDR
mechanisms. Madavan et al. (1984) considered the
polymer- and/or particulate-flow hypothesis of
Lumley (1973, 1977) as a means of explaining BDR
through increased viscosity and decreased density in
the near-wall bubble laden flow. Mari (1987)
explored the implications of changing the bulk
density and viscosity beyond the sub-layer on the
mean boundary layer profile during BDR, while
Legner (1984) used a simple stress model that
included density reduction and turbulence
modification. Madavan et al. (1985b) modeled the
bubbly boundary layer as a homogeneous fluid with
spatially varying density and viscosity, and found
agreement with the experimental results of Madavan
et al. (1985a). This model employed an a priori
assumption that the presence of bubbles will
significantly influence the production of turbulence
only through variation of the eddy viscosity or
mixture properties. Bubble splitting requires energy
that may be supplied by turbulence (Hinze 1955), so
it provides a potential means of turbulent energy
reduction that could lead to reduced skin friction.
Meng & Uhlman (1998) proposed bubble splitting as
a mechanism for BDR.
Direct numerical simulation (DNS) of
turbulent bubbly flows has been conducted at
relatively low Reynolds numbers. Druzhinin &
Elghobashi (1998) performed a DNS of a
microbubble-laden mixing layer to examine how
bubbles accumulate in a region of concentrated
vorticity. Vance et al. (2003) used DNS and large-

w is the skin friction, is the liquid density, and U is


the local flow speed at the edge of the boundary
layer. C Fo will denote skin friction measured without
bubbles, and BDR results will be presented in the
usual way as a skin friction ratio, CF CFo . In
addition, the flows friction velocity u* is defined in

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

two-component LDV at x = 10.7 cm mounted on


multi-axis traverse (these LDVs are described in
Bourgoyne et al. 2003). The experiments were
conducted at three flow speeds corresponding to 6
ms1, 12 ms1, and 18 ms1 in the LCCs empty test
section.
Compressed air was injected through one of
two gas injectors having spans of 2.65 m located at x
= 1.32 m, the upstream injector or UI, and at x = 9.79
m, the downstream injector or DI. The compressed
air, filtered to remove particles as small as 1 micron,
was supplied to each injector (cross-section shown in
figure 2) through a 40-port manifold with a port
spacing of approximately 6.5 cm. The compressed
and filtered air was metered using five independent
Sierra Instruments regulating flow meters (0.09 m3s-1
capacity). The experiments involved four volumetric
airflow rates: Qa = 0.05, 0.09, 0.19, and 0.38 m3s1.
Several diagnostic systems were employed
to make measurements of the single-phase and
bubbly flows. The single-phase flow measurements
included LDV, static pressure, and shear stress. The
bubbly flow measurements included bubble images,
near-wall electrical impedance, and shear stress.
During the experiments the water temperature ranged
from 17.5C to 23.3C.
Static pressure was measured at multiple
downstream locations on the wall of the test section
above and below the model using a Paroscientific
Digiquartz (model 740) pressure transducer that was
calibrated with a Druck DP601 standard. These static
pressure measurements were used to calculate the
streamwise pressure gradient, the free-stream flow
velocity outside the plates boundary layer, and the
free-stream acceleration.
The average wall shear stress, w = Df/A, was
measured directly using six floating-element straingauge force balances that measured the skin-friction
drag force, Df, on round area A having a 15.24 cm
diameter. These sensors were custom-designed and
fabricated at the Applied Research Laboratory at the
Pennsylvania State University. These six force
balances were all located at the same spanwise
coordinate (approximately 1/3 span) at x = 1.96 m,
3.41 m, 5.94 m, 7.43 m, 9.23 m, and 10.68 m (see
figure 1). The sensor located at x = 5.94 m did not
function properly for most of the injected flow
conditions, and in these cases, the data from this
sensor were removed from the presentation of results.
Strain gauge amplifiers (Vishay 2310) were
employed for excitation, nulling, and amplification of
the force balance signals. Data were collected at a
rate of 100 Hz for variable lengths of time ranging
from 10 seconds to several minutes. The force
balances were individually calibrated at least once
per test day.

eddy simulation (LES) to model features of a


microbubble cloud, and found that most
microbubbles reside within the flows coherent
structures. Xu et al. conducted a channel flow DNS
using a force coupling model between the bubbles
and the liquid and concluded that initial bubble
seeding, density reduction, bubble-turbulence
interactions, and bubble-wall proximity were all
important for BDR. Fernndez et al. (2003)
performed DNS to compute the interaction of
deformable bubbles with concentrated vortices and
found that migration of microbubbles into streamwise
vortices can modify the near-wall velocity
fluctuations.
EXPERIMENTAL SET-UP, QUANTITIES
MEASURED, AND THE EXPERIMENTAL
PROGRAM
The experiments were conducted in the US
Navys Large Cavitation Channel (LCC), the worlds
largest low-turbulence recirculating water tunnel (see
Etter and Wilson 1992, Park et al. 2003). The
nominal test section dimensions are 13 m (length) by
3.05 m x 3.05 m (width and height). Test section flow
speeds and pressures can be set from 0.5 to 18.3 ms1
and from 3.5 to 410 kPa, respectively. The LCCs
free stream turbulence level is below 0.5%.
The test model, shown schematically in
figure 1, was a rigid flat plate (L = 12.9 m long, 3.0 m
wide, 18.4 cm thick) designed to span the width and
the length of the LCCs test section. Its leading edge
was a 4 to 1 ellipse and the test surface of the model
faced downward so that gravitational buoyancy
forced bubbles toward the test surface. The test
surface was made from 304 stainless steel and was
polished fully assembled to a root-mean-square
surface roughness of k 0.4 m. A distributed
roughness boundary layer trip, consisting of a film of
epoxy embedded with 120-micron diameter particles
arranged randomly with a separation of 2 to 5 mm,
was applied from 2.5 cm x 27.5 cm across the
entire span of the model. The models trailing edge
was an 8 wedge truncated with a 45 asymmetric
bevel where its thickness decreased to ~2 cm. The
models thickness produced a test section area
blockage of 6%. The model was mounted with its
center-plane 5.7 cm below the test sections vertical
centerline. During testing, minor pressure differences
between the top and bottom of the model produced a
maximum vertical displacement of 2 mm at 50%
span.
Nominal flow speeds were monitored by a
single-component laser Doppler velocimeter (LDV)
at the fixed location x = 3.2 cm, y = 44.8 cm, and z =
76.2 cm from the LCC wall (25% span); and by a

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

As is common for strain gauge equipment,


the force-balance outputs under zero load tended to
drift mildly over periods of several hours. Thus, noflow output measurements at a known pressure were
recorded several times daily so that an accurate
compensation for zero drift could be made. The
combined error varied between force balances, but
was generally within 5% of the final skin friction
measurement.
A Sony XC-55BB progressive-scan blackwhite miniature digital camera system and a VCL12YM C-mount lens were used to view the bubbles
closest to the surface of the plate. An external trigger
was employed with the camera system, allowing for a
shutter-speed to as small as 10 sec. The cameras
resolution was 659 x 494 pixels with a pixel
dimension of 7.4 x 7.4 microns. The cameras field of
view was 11.0 mm by 8.3 mm with a focus depth-offield of 2.1 mm (measured in water). The camera
looked downward from the model, and was mounted
on a remote-controlled micro-traverse (Optosigma
model SGSP60-5ZF) capable of 1-micron
incremental motion over a range of 5 mm. This
traverse allowed the cameras focus field to be placed
at different distances from the surface. Bubble
images were recorded at x = 1.96 m and 10.68 m
using a 4-Watt argon-ion laser to provide
illumination via an optical fiber. A digital image
acquisition system (BitFlow Inc., model RAV-HCI110-VNS) was used to acquire the images at three
fps, and Video Savant (V3.0, IO Industries) digital
video recording software was used to process and
evaluate the images to determine average bubble
sizes, bubble locations (relative to the test surface),
bubble velocities, and an estimate of the local void
fraction. Bubble radii were measured graphically for
each acceptable (i.e. in focus) bubble. A test target
was used to determine the images pixel spacing: 18
microns/pixel. Bubble distances from the surface of
the model were deduced from the cameras vertical
position, and its measured focal characteristics. To
obtain bubble velocity measurements, the cameras
shutter speed was increased to 100 or 200 s to create
bubble streaks.
The effects of the near-wall bubble
concentration on the friction drag reduction were
assessed using the information obtained from the
bubble images. It was not possible to deduce void
fraction, , from the two-dimensional bubble images.
Alternatively, the bubble area ratio Abubble/Atotal (that
usually over-predicts ) was exacted from the images.
Relative void fraction in the boundary layer
was estimated using eight electrical impedance
probes located at x = 1.07, 1.96, 3.41, 5.94, 7.43,
9.23, 10.68, and 11.50 m. Each probe consisted of

two brass electrodes 3.2 mm in diameter, spaced 6.4


mm apart (centerline to centerline), flush-mounted in
a non-conducting pvc disk. These instruments
measured the local electrical impedance of the flow
using a balanced AC bridge operated at 50 kHz. The
demodulated bridge signal was low-pass filtered and
sampled at 10 kHz. Each probe measured the gaswater-mixture impedance in its immediate vicinity,
and these measurements can be related to the local
void fraction (see Ceccio & George 1996).
Immediate vicinity as used implies an approximately
hemispherical volume having a radius on the order of
the electrode spacing (6.4 mm) with measured
impedance changes being proportional to void
fraction in this volume. The calibrations necessary
for absolute void fraction measurements were not
conducted, thus only the relative change in
impedance with varying injection rates and
downstream position are available and reported.
THE IVFM (IMPEDANCE VOID FRACTION
METER) MEASUREMENT TECHNIQUE
To more easily quantify the void fraction
immediately adjacent to the plate, an impedance
technique was developed and utilized since the void
fraction of the medium can be derived by measuring
the impedance of the medium. The instrument works
as follows. Two measuring electrodes are placed in
the area of interest and an AC field is excited
between them. A bridge detects any change in
impedance from the reference condition. The
difference signal is demodulated and filtered. The
output signal is proportional to the impedance
change. Since the change in bulk impedance is
correlated with the void fraction, this signal can be
used to calculate the void fraction of the bubble water
mixture.
The IVFM creates an electrostatic field with
no charge density within the calculation domain (i.e.
the adjacent fluid layer), and thus the electric
potential satisfies the Laplace equation. A
temperature field equation with no heat generation is
equivalent to that of the electrostatic field with no
charge density. Temperature is equivalent to electric
potential while thermal conductivity is equivalent to
electric conductance. Therefore, the IVFM
characteristics can be simulated by solving the heat
conduction problem. Herein commercial software,
FLUENT, is used to simulate the IVFM by
calculating the heat conduction problem with
appropriate boundary conditions.
An estimation of the position of the bubbles
was performed through numerical simulation
utilizing the area ratio information obtained near the
wall using the bubble images. A simplified

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Maxwells equation can be used to determine the


conductance of the mixture if both void ratio and
conductance of the water are known. The simplified
equation is

c =

2 1 2 mc
.
mc + 2 1

DRR = f ( , h, d ; l )

where is the air concentration ratio,


h is the mixture location,

(1)

d is the average bubble diameter, and


l is the length scale related to the turbulent
boundary layer.

Rewriting this equation for mc, the mixture


conductivity, yields

mc =

2 1 2 1 c
c + 2

(3)

After non-dimensionalization of the parameters,

(2)

h d
DRR = f ( , , ) .
l l

where c is the void fraction and 1 is the


conductance of the water. The area ratio is substituted
for the void fraction, and using the conductance of
water (deduced from the impedance at the balanced
bridge state), the conductance of the mixture is
calculated. Simulation was performed by estimating a
thickness of the water layer above the plate surface,
and assuming a mixture above that level. The
estimated position of this interface was varied until
the impedance change predicted by the simulation
matched closely that of the experimental
measurement. This determined the position of the
waterair-water-mixture interface at that time.
During the HIPLATE experiment the void
ratio ranged from 10% to 50%, and mixtures were
located between 0.5mm and 5.0mm from the wall
(except for a few cases). Therefore, using simulation
the impedance change within these bounds of void
fraction and mixture location was calculated. Within
this range, scaled impedance changes are strongly
affected by mixture location, but only very weakly
influenced by void fraction. This is evident in figure
3. The estimated relationship between the scaled
impedance and the mixture location using the least
square method was performed under the assumption
that the scaled impedance change is independent of
the void ratio. The results suggested that the scaled
impedance change is inversely proportional to the
1.578 power of the mixture location. This implies
that the scaled impedance change can be used for the
mixture location.
The physical parameters that contribute to
the drag reduction ratio (DRR) on smooth flat
surfaces could be categorized into two groups. One
group consists of parameters related to air injection
including void fraction, mixture location, and average
bubble size. The other comprises parameters related
to the turbulent boundary layer at the measurement
location. One plausible functional form would be:

(4)

The
physical
parameters
affecting
impedance change include the void ratio and the
mixture location as well as the sensor geometry and
the water conductance. Since the geometry of the
sensor was fixed once chosen, and the water
conductance was that of Memphis city water during
the HIPLATE experiment, physical parameters
influencing the impedance change (IC) can be
reduced to the void ratio and the mixture location, or
IC = g ( , h)

(5)

where is the air void fraction,


h is the mixture location.
Combining the simulation result that the
scaled impedance change is inversely proportional to
the 1.578 power of the mixture location impedance,
the change can be expressed as
IC = g ( , h) = g 1 ( ) h 1.578 .

(6)

Taking Equation (6) to the 1/1.578 power of


the impedance change and multiplying by the length
scale of the turbulent boundary layer, the modified
impedance change (MIC) becomes
MIC = IC 1 / 1.578 l = g 1 ( )1 / 1.578

l
.
h

(7)

However, the appropriate length scale of the


turbulent boundary layer is not yet determined. If the
length scale used in MIC is the proper length scale
for drag reduction, MIC will be correlated highly to
DRR regardless of the injection locations. The
integral length scale momentum thickness ( 0 ), and
the inner layer length scale wall unit ( l 0 ) measured

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

during the HIPLATE experiment were used as length


scales for MIC.

velocities. The interested reader is referred to this


paper for these results.
Near-wall phase composition measurements
and persistence are discussed next. The results
obtained from bubble images indicate that the
location of the bubbles are significant parameters for
BDR. Images also qualitatively indicate that there is a
relationship between the concentration of bubbles
near the wall and the degree of drag reduction. The
relationship between the measured drag reduction
and the observed area ratio, Ab /Atotal is shown in
figure 7 for the 12 and 18 ms1 cases. The 6 ms1 case
was not included due to air-film formation. Figure 7
indicates that an increase in area ratio (which is
monotonically related to void fraction) results in
improved drag reduction. The substantial difference
in drag reduction between similar area ratios (e.g. 12
ms1 and 18 ms1 with upstream injection at x = 1.96
m) is attributable to the position of the bubbles with
respect to the surface. These observations also
indicate that the void-fraction drag-reduction
relationship is not critically dependent on the
thickness of the boundary layer as set by the location
of injection.
Note that the largest area ratio (18 ms1 with
upstream injection at x = 1.96 m) does not correspond
to the case of the greatest drag reduction. Rather, it is
the combination of large void fraction, likely in
conjunction with small bubbles, in close proximity to
the wall that generates drag reduction.
The area ratio results in figure 7 also argue
that BDR is not merely an effect of reduced density.
The area ratio is an over-estimate of the local volume
fraction. Consider mono-disperse spheres in the facecentered-cubic configuration. The volume ratio of the
spheres is 0.74, while an image of the spheres taken
at a planar boundary and processed with the
algorithm described above would yield an area ratio
of 0.79. Inspection of figure 7 finds that an area ratio
of ~30% yielded a drag reduction of ~45%. This
suggests that the observed BDR cannot be solely due
to a reduction in density, since the skin-friction is
reduced in higher proportion than the reduction in
near-wall density.
Measurements of the local impedance of the
bubbly flow using the IVFM technique provided an
alternate assessment of the relative void fraction in
the near-wall region of the boundary layer. Figure 8
shows the mean impedance values at multiple
locations along the surface of the model during
upstream injection for the 6 ms1 (figure 8a), 12 ms1
(figure 8b) and 18 ms1 (figure 8c) flow speeds. The
impedance probe located at x/L = 0.08 is upstream of
injection, and is used to represent the baseline
impedance for the flow. In each case depicted in
figure 8, the mean impedance increases sharply

RESULTS
The characteristics of the single-phase
boundary layer were measured in preliminary tests
without bubble injection. A mild favorable pressure
gradient was measured for all flow conditions. It was
shown also that the plate was hydraulically smooth.
The first results presented are those of the
surface shear stresses. Spatially averaged surface
shear stress measurements, without air injection, are
presented in figure 4 as CFo vs. Rex. A least-squares
fit of the measured data yields a power-law
relationship between CFo and Rex,
CF o = (0.025 0.006) Re x ( 0.149 0.017) .

(8)

Figure 4 also shows the skin-friction relationship of


Schultz-Grunow (1941) that was derived from
experiments to Rex 107, and extrapolated to Rex
109. Equation (8) falls 5% below the Schultz-Grunow
(1941) relationship over the range of Reynolds
numbers investigated. To suppress bias errors, the
power-law fit (8) was used for CFo in the skin friction
ratio, CF /CFo, throughout the following BDR results.
Figure 5 presents a comparison of the
present drag reduction results to prior BDR studies
on smooth flat surfaces. Here, CF /CFo is plotted
versus an estimated-flux volume fraction of the gas,
Qa/(Qa+Qw). Overall, the agreement is not
particularly good across the breadth of the flat-plate
BDR data, with the exception of the trend that CF
/CFo approaches unity for small abscissa values (
0.1). It can be concluded from figure 5 that BDR
generally increases with increasing gas injection.
Figure 6 provides a different method of
normalizing the air flow rate in BDR experiments.
Figure 6 shows the present data along with the results
of Deutsch et al. (2003). Here, the momentum
thickness, o, replaces *, and there is an
improvement in the agreement between the upstream
injection results at 12 ms1 and 18 ms1, and the
results of Deutsch et al. (2003). The upstream
injection results for 6 ms1 and the downstream
injection results for all flow speeds do not follow the
primary trend. The 6 ms1 results are not expected to
agree since they involved an air film on the model
surface.
It is noted that in a separate manuscript
(Sanders et al.) the following results are presented:
bubbly flow characteristics and mean bubble sizes;
bubble locations relative to the surface; bubble size
distributions and coalescence; and near-wall bubble

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

following injection due to the sudden presence of air


bubbles in the flow, and the mean impedance is
consistently greater for larger Qa. For the 18 ms1
cases (figure 8c), the initial surge of impedance,
particularly with Qa = 0.38 m3s1 injection, rapidly
dissipates as the impedance level returns nearly to its
baseline value by x/L = 0.26. Figure 8d presents BDR
as a function of downstream distance for the 12 and
18 ms1 cases. The drag reduction diminishes as the
bubbly flow moves downstream. The data trends of
the near wall impedance measurements are mirrored
in that of the relative drag reduction.
Impedance measurements made during
downstream air injection are shown in figure 9. Only
the sixth, seventh and eighth probes (located at x/L =
0.72, 0.83 and 0.91) that are downstream of the
downstream injection are presented. The 6 ms1 data
show evidence of an air film, particularly for the high
air injection rates. The 12 and 18 ms1 results show
the initial surge of local mean impedance
immediately downstream of the injection point,
followed by a rapid decrease. From these results it is
evident that injection into a thicker boundary layer
does not significantly alter or improve the persistence
of BDR.
In figure 10 the drag reduction ratio (DRR),
defined as the reduced drag force due to air injection
divided by the drag force without air injection, as a
function of impedance change, is presented for both
upstream and downstream injection during the
HIPLATE experiment. Although the drag reduction
ratio exhibits increases as the impedance change
becomes large for both injection cases, there is a
significant difference in slopes between the two.
Since the impedance change is a strong indicator of
the mixture location, a plausible explanation might be
that the drag force is reduced the closer the mixture
gets to the wall, but the degree of drag reduction is
highly dependent on the local length scale of the
boundary layer.
Figures 11 and 12 show the DRR variations
with MIC (modified impedance change) scaled by the
momentum thickness and scaled by the wall unit.
DRR is correlated better to MIC scaled by the wall
unit for both injection locations. This implies that the
mixture location scaled by the wall unit is a proper
dimensionless parameter in microbubble drag
reduction and that MIC scaled by the wall unit can
serve as a viable indicator of DRR. Average bubble
size divided by the wall unit is also a dimensionless
parameter of the drag reduction ratio. However,
considering the strong correlation between DRR and
MIC, that is nearly independent of average bubble
size, suggests that average bubble diameter does not
play an important role in drag reduction.

To scale bubble migration, a simple oneway coupled bubble tracking model was developed.
The model and the results gleaned from it can be
found in Sanders et al.
SUMMARY AND CONCLUSIONS
The skin friction drag reduction resulting
from the introduction of air bubbles into a high
Reynolds number flat-plate turbulent boundary layer
has been investigated experimentally. The singlephase and multiphase surface shear stress was
measured, along with the near wall bubble
characteristics. A new technique, IVFM, was
discussed and implemented, and shown to estimate
void fraction well.
The data presented suggest that the amount
and persistence of microbubble drag reduction are
dependent on both inner and outer boundary layer
characteristics. The reduction of skin-friction drag
was scaled with an average boundary layer gas flux
employing the relative momentum thickness and was
found consistent with prior data from Deutsch et al.
(2003).
The amount of drag reduction strongly
depends on the near-wall void fraction. At low
speeds, the injected bubbles coalesced to form a
nearly continuous gas film beneath the model
surface. As the freestream speed is increased, the
injected bubbles remain distinct. When the bubbles
remained near the surface (i.e. within 300 wall units),
the friction drag was reduced substantially.
Examination of the near-wall bubble void fraction
implies that a combination of density reduction and
turbulence modification is responsible for the
reduction in friction drag.
The presence of high shear near the plate
surface leads to the depletion of bubbles in the near
wall region. Near wall impedance measurements by
IVFM correlate reasonably with the shear stress
reduction, demonstrating that the drag reduction is
lost when the bubbles migrate from the near-wall
region. This formation of a liquid layer between the
bubbly mixture and the surface leads to a loss of drag
reduction. Also, entrainment of liquid into the
boundary layer leads to a dilution of the bubbly
mixture with a resulting loss of drag reduction.
ACKNOWLEDGEMENTS
The authors of this paper wish to
acknowledge the significant contributions of Kent
Pruss, Paul Tortora, and Hans van Sumeren of the
University of Michigan; Ivo VanderHout of the Delft
University of Technology; Bruce Hornaday, Dr. J.
Michael Cutbirth, and the LCC technical staff from

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

the Naval Surface Warfare Center Carderock


Division; and Duncan Brown of the Johns Hopkins
University Applied Physics Laboratory. This effort
was sponsored by the Defense Advance Research
Projects Agency (Dr. Lisa Porter, Program Manager)
and the Office of Naval Research (Dr. L. Patrick
Purtell, Program Manager) under ONR contract
number N00014-01-1-0880. The content of this
document does not necessarily reflect the position or
the policy of the United States Government, and no
official endorsement should be inferred.

Proceedings of ASME Fluids Engineering


Division Summer Meeting 2003, 1-4.
Fontaine, A. A. & Deutsch, S., (1992) The influence
of the type of gas on the reduction of skin
friction drag by microbubble injection,
Experiments in Fluids 13, 128-136.
Hinze, J. O. (1955) Fundamentals of the
Hydrodynamic Mechanism of Splitting in
Dispersion Processes, A.I.Ch.E. Journal 1 (3),
289-295.
Kawamura, T., Kakugawa, A., Kodama, Y.,
Moriguchi, Y. & Kato, H. (2003) Effect of
bubble size on the microbubble drag reduction
of a turbulent boundary layer, Proceedings of
ASME Fluids Engineering Division Summer
Meeting 2003, 1-8.
Kodama, Y., Kakugawa, A., Takahashi, T. &
Kawashima, H. (2000) Experimental study on
microbubbles and their applicability to ships for
skin friction reduction, International Journal of
Heat and Fluid Flow 21, 582-588.
Legner, H. H. (1984) A simple model for gas bubble
drag reduction, Physics of Fluids 27 (12),
2788-2790.
Lumley, J. L. (1973) Drag reduction in turbulent
flow by polymer additives, Journal of Polymer
Sciences: Macromolecular Review 7, 263-290.
Lumley, J. L. (1977) Drag reduction in two phase
and polymer flows, Physics of Fluids 20 (10),
S64-S70.
Madavan, N. K., Deutsch, S. & Merkle, C. L. (1984)
Reduction of turbulent skin friction by
microbubbles, Physics of Fluids 27 (2), 356363.
Madavan, N. K., Deutsch, S. & Merkle, C. L. (1985a)
Measurements of local skin friction in a
microbubble-modified turbulent boundary
layer, Journal of Fluid Mechanics 156, 237256.
Madavan, N. K., Merkle, C. L. & Deutsch, S.
(1985b) Numerical investigations into the
mechanisms of microbubble drag reduction,
Journal of Fluids Engineering 107, 370-377.
Mari, J. L. (1987) A simple analytical formulation
for
microbubble
drag
reduction,
Physicochemical Hydrodynamics 8 (2), 213220.
McCormick, M. E. & Bhattacharyya, R. (1973)
Drag reduction on a submersible hull by
electrolysis, Naval Engineers Journal 85, 1116.
Meng, J. C. S. & Uhlman, J. S. (1998) Microbubble
formation and splitting in a turbulent boundary
layer for turbulence reduction, International
Symposium on Seawater Drag Reduction, p.
341-355.

REFERENCES
Bogdevich, V. G. & Evseev, A. R. (1976) The
distribution on skin friction in a turbulent
boundary layer of water beyond the location of
gas injection, Investigations of Boundary
Layer Control (in Russian), Thermophysics
Institute Publishing House, 62.
Bogdevich, V. G. & Malyaga, A. G. (1976) Effect
of gas saturation on wall turbulence,
Investigations of Boundary Layer Control (in
Russian), Thermophysics Institute Publishing
House, 49.
Bourgoyne, D. A., Hamel, J. M., Ceccio, S. L. &
Dowling, D. R. (2003) Time-averaged flow
over a hydrofoil at high Reynolds number,
Journal of Fluid Mechanics 496, 365-404.
Ceccio, S. L. & George, D. L. (1996) Review of
electrical impedance techniques for the
measurement of multiphase flows, Journal of
Fluids Engineering 118, 391-399.
Clark III, H. & Deutsch, S. (1991) Microbubble skin
friction reduction on an axisymmetric body
under the influence of applied axial pressure
gradients, Physics of Fluids A 3 (12), 29482954.
Deutsch, S. & Castano, J. (1986) Microbubble skin
friction reduction on an axisymmetric body,
Physics of Fluids 29 (11), 3590-3597.
Deustch, S., Moeny, M., Fontaine, A. & Petrie, H.
(2003) Microbubble drag reduction in rough
walled turbulent boundary layers, Proceedings
of ASME Fluids Engineering Division Summer
Meeting 2003, 1-9.
Druzhinin, O. A. & Elghobashi, S. (1998) Direct
numerical simulations of bubble-laden turbulent
flows using two-fluid formulation, Physics of
Fluids 10 (3), 685-697.
Etter, R.J, and M.B. Wilson (1992) "The Large
Cavitation Channel," Proc. the 23rd American
Towing Tank Conf., New Orleans.
Fernndez, A., Lu, J. & Tryggvason, G. (2003)
Bubble effects on wall shear in vertical flows,

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Merkle, C. L. & Deutsch, S. (1992) Drag reduction


in liquid boundary layers by gas injection,
Progress in Astronautics and Aeronautics 123,
351-412.
Migirenko, G. S. & Evseev, A. R. (1974) Turbulent
boundary layer with gas saturation, Problems
of Thermophysics and Physical Hydrodynamics
(in Russian).
Pal, S., Merkle, C. L. & Deutsch, S. (1988) Bubble
characteristics and trajectories in a microbubble
boundary layer, Physics of Fluids 31 (4), 744751.
Park, J. T., Cutbirth, J. M. & Brewer, W. H.
Hydrodynamic performance of the large
cavitation channel (LCC), Proceedings of
ASME Fluids Engineering Division Summer
Meeting 2003, 1-14.
Sanders, W.C., Dowling, D.R., Perlin, M. & Ceccio,
S.L. (submitted) Bubble Friction Drag
Reduction in a High Reynolds Number Flat
Plate Turbulent Boundary Layer, Journal of
Fluid Mechanics.
Schultz-Grunow, F. (1941) New frictional resistance
law for smooth plates, NACA Technical
Memorandum 17 (8), 1-24.

Takahashi, T., Kakugawa, A., Nagaya, S.,


Yanagihara, T. & Kodama, Y. (2001)
Mechanisms and scale effects of skin friction
reduction by microbubbles, 2nd Symposium on
the Smart Control of Turbulence, 1-9.
Vance, M. W., Sugiyama, K., Takagi, S. & Squires,
K. D. (2003) Microbubble transport in
turbulent channel flow, Proceedings of ASME
Fluids Engineering Division Summer Meeting
2003, 1-8.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Shear Stress Sensors

x = 11.50 m

Single Point Electrodes


x = 9.23 m

Bubble Camera

x = 7.43 m

x = 10.68 m
Downstream Injector,
x = 9.79 m

x = 3.41 m
x = 1.07 m
x = 5.94 m

y
x

x = 1.96 m
Upstream Injector, x = 1.32 m
x = 0.00 m

Figure 1. HIPLATE schematic showing locations of the shear stress sensors (located 50.8 cm from the
spanwise centerline), single point electrodes (located 25.4 cm from the spanwise centerline), and the bubble
camera (located 77.5 cm from the spanwise centerline at x = 1.96 m and x = 10.68 m). The streamwise position
of each measurement location is indicated relative to the leading edge of the model. The model is presented here
with the test surface facing upward, although the model was oriented with the test-surface facing downward
during the experiment.

10

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Flow
40 micron porous
stainless steel

Contracting slot

3 screens
0.4 mm thick

4.3 cm
3.0 cm

Gas In
1.3 cm
4.8 cm

Figure 2. Cross-sectional schematic of the gas injectors. The injector measures 2.65 m in the spanwise
direction. The gas inlet at the bottom of the injector consists of forty 1.3-cm ports spaced evenly across the span
of the injector. Three brass perforated plates with 0.5-mm diameter holes provide the pressure drop to evenly
distribute the gas across the injector. The contracting slot has a 10 taper leading to the 40-micron pore-size
sintered stainless steel.

11

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

1
10% air concentration ratio
30% air concentration ratio
50% air concentration ratio

Scaled Impedance

Power ( )

-1.578

Scaled Impedance = 0.4483H

0.5

0
0

Mixture location H (mm)


Figure 3. Scaled impedance change as a function of mixture location (void ratio from 10% to 50% and mixture location
between 0.5mm and 5.0mm from the wall).

12

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

0.0030
Sensor 1, x/L = 0.15
Sensor 2, x/L = 0.26
Sensor 3, x/L = 0.46
Sensor 4, x/L = 0.58
Sensor 5, x/L = 0.72
Sensor 6, x/L = 0.83
Power law fit
Schultz-Grunow

0.0025

CFo 0.0020

0.0015

CFo = 0.025Re0.149
x
0.0010
0.0E+00

5.0E+07

1.0E+08

1.5E+08
Rex

2.0E+08

2.5E+08

Figure 4. Skin friction coefficient, CFo, in the absence of bubbles as a function of downstream-based Reynolds
number. A power-law least-squares fit was applied to the data, and the resulting equation is represented by the solid
line. The dashed line is the resistance law of Schultz-Grunow (1941).

13

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

-1

1.2

#
1.0
+

#
#+
+
##
#+
#

#
#
##
+
##

++

##
#
#
+#
+
#
0.8
#
+####
++
++
#++##
###

CF
C Fo 0.6

0.4

0.2

0.0

0.0

0.2

HIPLATE, U = 6 ms , UI
HIPLATE, U = 6 ms-1, DI
-1
HIPLATE, U = 12 ms , UI
HIPLATE, U = 12 ms-1, DI
HIPLATE, U = 18 ms-1, UI
HIPLATE, U = 18 ms-1, DI
Bogdevich et al., U = 2 ms-1
Bogdevich et al., U = 4 ms-1
Bogdevich et al., U = 6 ms-1
MDM, U = 5 ms-1, plate below TBL
MDM, U = 5 ms-1, plate above TBL
MDM, U = 11 ms-1, plate below TBL
MDM, U = 11 ms-1, plate above TBL
-1
MDM, U = 17 ms , plate below TBL
MDM, U = 17 ms-1, plate above TBL
-1
Deutsch et al., U = 7 ms
-1
Deutsch et al., U = 10 ms
-1
Kodama et al., U = 5 ms
-1
Kodama et al., U = 7 ms
-1
Kodama et al., U = 10 ms

0.4

0.6

0.8

Qa

Qa + Ub( o o )
Figure 5. The skin friction ratio presented as a function of the volumetric fraction of gas flow rate for the
present work, compared to the results of previous researchers. The data from Bogdevich et al. represent those
acquired by Soviet researchers (1974-1976) in plate-up and plate-down experiments. MDM denotes the data
reported by Madavan, Deutsch & Merkle (1984, 1985a).

14

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

HIPLATE, U = 6 ms-1, UI
-1
HIPLATE, U = 6 ms , DI
-1
HIPLATE, U = 12 ms , UI
-1
HIPLATE, U = 12 ms , DI
HIPLATE, U = 18 ms-1, UI
-1
HIPLATE, U = 18 ms , DI

1.2

1.0

0.8

CF
C Fo

-1

Deutsch et al., U = 7 ms
Deutsch et al., U = 10 ms-1

0.6

0.4

0.2

0.0

0.0

0.2

0.4

0.6

0.8

1.0

Qa
Q a + Ub o

Figure 6. The skin friction ratio presented as a function of the volumetric fraction of the gas flow rate using the
momentum thickness of the unmodified boundary layer, o. The smooth-wall data presented by Deutsch et al.
(2003) are also plotted.

15

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

12 ms-1, UI, x = 1.96 m


18 ms-1, UI, x = 1.96 m
12 ms-1, DI, x = 10.68 m
18 ms-1, DI, x = 10.68 m
12 ms-1, UI, x = 10.68 m
18 ms-1, UI, x = 10.68 m

1.0

0.9

0.8

CF
C Fo
0.7

0.6

0.5
0.0

0.1

0.2

0.3

0.4

0.5

0.6

Ab
Atotal
Figure 7. The measured skin friction ratio as a function of the imaged area ratios, given by Ab /Atotal, where Ab
is the area of the image covered by focused bubbles and Atotal is the total image area for 12 and 18 ms-1. UI =
upstream air injection at x = 1.32 m; DI = downstream air injection at x = 9.79 m.

16

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Injection

Injection
200

200
U=6
U=6
U=6
U=6

3 -1

ms
3 -1
ms
3 -1
ms
3 -1
ms

U=12
U=12
U=12
U=12

175

Mean Impedance [kohm]

Mean Impedance [kohm]

175

-1

ms , UI, Q a=0.05
-1
ms , UI, Q a=0.09
-1
ms , UI, Q a=0.19
-1
ms , UI, Q a=0.38

150
125
100
75
50
25

3 -1

ms
3 -1
ms
3 -1
ms
3 -1
ms

150
125
100
75
50
25

0
0.0

0.2

0.4

0.6

(a)

0.8

1.0

0.0

0.2

0.4

(b)

L
Injection

0.6

0.8

1.0

Injection
1.0

200
U=18
U=18
U=18
U=18

175

ms-1, UI, Q a=0.05


-1
ms , UI, Q a=0.09
-1
ms , UI, Q a=0.19
-1
ms , UI, Q a=0.38

m3s-1
3 -1
ms
3 -1
ms
3 -1
ms

-1

0.8

125
100
75

3 -1

12 ms , 0.05 m s
-1
3 -1
12 ms , 0.09 m s
-1
3 -1
12 ms , 0.19 m s
-1
3 -1
12 ms , 0.38 m s
-1
3 -1
18 ms , 0.05 m s
-1
3 -1
18 ms , 0.09 m s
-1
3 -1
18 ms , 0.19 m s
-1
3 -1
18 ms , 0.38 m s

150

1-(CF/CFo)

Mean Impedance [kohm]

-1

ms , UI, Q a=0.05
-1
ms , UI, Q a=0.09
-1
ms , UI, Q a=0.19
-1
ms , UI, Q a=0.38

0.6

0.4

50

0.2
25

0.0

0
0.0

(c)

0.2

0.4

0.6

0.8

1.0

0.0

0.2

(d)

0.4

0.6

0.8

x
L

Figure 8. (a-c) Mean near-wall impedance as a function of the normalized downstream distance. These graphs
represent measurements recorded during upstream injection at (a) 6 ms-1; (b) 12 ms-1; and (c) 18 ms-1. The
vertical bars represent the maximum and minimum values of the impedance range during the sampling interval
(15 seconds). The sensor located at x/L = 0.08 is located just upstream of injection (x/L = 0.11), and is used to
represent the baseline impedance for the flow. (d) Measured drag reduction during upstream injection at 12 and
18 ms-1, presented as a function of the normalized downstream distance.

17

Copyright National Academy of Sciences. All rights reserved.

1.0

Twenty-Fifth Symposium on Naval Hydrodynamics

Injection

Mean Impedance [kohm]

100

U=6 ms-1, DI, Qa=0.05 m3s-1


-1
3 -1
U=6 ms , DI, Qa=0.09 m s
U=6 ms-1, DI, Qa=0.19 m3s-1
U=6 ms-1, DI, Qa=0.38 m3s-1

Injection

-1

3 -1

-1

3 -1

U=12 ms , DI, Qa=0.05 m s


-1
3 -1
U=12 ms , DI, Qa=0.09 m s
-1
3 -1
U=12 ms , DI, Qa=0.19 m s
-1
3 -1
U=12 ms , DI, Qa=0.38 m s

75

U=18 ms , DI, Qa=0.05 m s


U=18 ms-1, DI, Qa=0.09 m3s-1
U=18 ms-1, DI, Qa=0.19 m3s-1
-1
3 -1
U=18 ms , DI, Qa=0.38 m s

50

Injection

25

0
0.6

0.8

1.0

0.6

0.8

1.0

0.6

0.8

1.0

x
L
Figure 9. Measured mean impedance of the near-wall flow during downstream injection, as a function of the
normalized downstream distance. The three impedance sensors are located at x/L = 0.72, 0.83 and 0.91, and the
injector is located at x/L = 0.76.

18

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

0.7
Upstream Injection
Downstream Injection

Drag reduction ratio (1-Cf/Cf0)

0.6
0.5
0.4
0.3
0.2
0.1
0.0
0

10

15

20

25

30

35

Impedance change (k)


Figure 10. The drag reduction ratio as a function of impedance change for upstream and downstream injection.
0.7
Upstream Injection
Downstream Injection

Drag reduction ratio (DRR)

0.6

0.5
DRR = 0.3165 MIC - 0.0325
2
R = 0.7684

0.4

0.3

0.2

0.1

0
0.00

0.20

0.40

0.60

0.80

1.00

1.20

1.40

1.60

1.80

Modified Impedance Change (MIC) using momentum thickness


Figure 11. The drag reduction ratio as a function of the modified impedance change using momentum thickness for the
upstream and downstream injection.

19

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

0.7
Upstream Injection
Downstream

Drag reduction ratio (DRR)

0.6

0.5

0.4
DRR = 0.8774MIC - 0.0088
R2 = 0.9041

0.3

0.2

0.1

0
0.00

0.10

0.20

0.30

0.40

0.50

0.60

0.70

0.80

Modified Impedance Change (MIC) using wall unit


Figure 12. The drag reduction ratio as a function of the modified impedance change using wall units for
upstream and downstream injection.

20

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics


St. Johns, Newfoundland and Labrador, CANADA, 8-13 August 2004

Tip Vortex Cavitation Inception Study Using the Surface


Averaged Pressure (SAP) Model Combined with a Bubble
Splitting Model
Jin-Keun Choi, Chao-Tsung Hsiao, and Georges L. Chahine
(DYNAFLOW, INC., U.S.A.)
near and above the cavitation inception number (Figure
2).
On the other hand, our recent studies with a more
precise non-spherical axisymmetric method (Choi and
Chahine, 2003a, 2003b, 2004, Choi, Chahine, and
Hsiao, 2003a, 2003b) revealed that the bubble nucleus
behavior at cavitation numbers just below the
cavitation inception number is associated with extreme
bubble elongation and splitting. Furthermore, these
studies showed that the bubble splitting is followed by
strong reentrant jets in the resulting offspring bubbles
(Figure 3) and by the emission of very high pressure
signals. A pressure signal is generated at the bubble
splitting; however, much stronger pressures may be
subsequently generated during the impact of the
reentrant jets and the collapse of the offspring bubbles.
The fact that the pressure signal from the subsequent
behavior of the offspring bubbles is much higher than
that from the collapse (splitting) of the original bubble
suggests the need to include the bubble splitting in the
SAP model in on-going and future studies. In addition,
bubble splitting obviously results in the modification of
the original nuclei size distribution, and this effect is
included in the present upgraded SAP model.
In order to avoid the higher computational cost of
non-spherical methods and, at the same time, to include
bubble splitting effects in the multiple bubble
simulations, we need to identify the conditions for
occurrence and to model the bubble splitting. Through
an extensive series of non-spherical simulations under
various conditions, we have developed rules for the
conditions under which a bubble splits, the timing of
the splitting, and the characteristics of the resulting
offspring bubbles. These rules are then implemented in
the upgraded SAP spherical model resulting in
dynamic modification of the bubble nuclei sizes and
distribution during a SAP simulation.
In this paper, we illustrate the method for the study
of cavitation inception in the flow field of a tip vortex.
First, the non-spherical numerical studies on bubble

ABSTRACT
Non-spherical simulations of bubble nuclei in a tip
vortex flow indicate bubble elongation followed by
splitting and reentrant jet formation. The associated
noise is much higher than the pressure signal from the
collapse of a spherical bubble. A major difficulty in
applying such a non-spherical method to a field of
nuclei is the prohibitively expensive cost of
corresponding computations. In this paper, we attempt
to overcome this difficulty by performing simulations
with an improved Surface Averaged Pressure (SAP)
spherical model that accounts for bubble splitting.
Non-spherical numerical studies are used to develop
splitting criteria and characterize the resulting noise
and offspring sub-bubbles. These criteria are
implemented in the SAP spherical model, and
numerical results with single bubble and with a
realistic field are then presented. The effect of the
inclusion of bubble splitting on the cavitation inception
prediction is then investigated.
INTRODUCTION
Conducting numerical cavitation inception experiments
using a distribution of nuclei as in a real liquid has
been very successful in providing a numerical tool to
study tip vortex cavitation inception and its scaling.
We have developed a Surface Averaged Pressure
(SAP) spherical bubble dynamics model (Hsiao,
Chahine, and Liu, 2003, Hsiao and Chahine, 2003a,
2003b) that has provided bubble dynamics results very
similar to those obtained using a 3-D two-way
interaction model (Figure 1). The SAP model is
especially appropriate for simulations with a large
number (~103 used so far) of bubble nuclei because of
its reasonable memory and computational time
requirements. The success of this model in the study of
cavitation inception is partly due to the fact that the
spherical approximation is valid at cavitation numbers

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

bubble splitting model based on axisymmetric


simulations of bubble nucleus under various conditions
in a typical tip vortex flow field of a hydrofoil. In this
section, we briefly summarize our observations of the
bubble splitting and the characterization of the
phenomena.

splitting leading to the development of splitting criteria


and characterization of the resulting offspring bubbles
are introduced and the implementation of the model
within the context of the SAP spherical model is
described. Numerical results with single and multiple
bubble nuclei are then presented, and the effect of
inclusion of bubble splitting on the cavitation inception
prediction is discussed.

0.006

Max. Equivalent Radius for Re=2.88E06, R o=50m

SAP Spherical Model


3DynaFS+UNCLE

0.005

Max. R EQ [m]

0.004

0.003

0.002

0.001

0
2.1

2.2

2.3

2.4

2.5

2.6

2.7

Figure 3: Bubble splitting and jet development of a nucleus


captured by a tip vortex (in time sequence from top left to
bottom right). (from Choi and Chahine (2004))

2.8

Cavitation Number

Figure 1: Comparison of maximum bubble sizes from SAP


spherical model and those from a non-spherical bubble
model. Note the excellent agreement between the SAP
spherical model and the 3-D two-way interaction model
(3DYNAFS+DF_UNCLE) from Hsiao and Chahine (2003b).

BUBBLE SPLITTING MODEL

The tip vortex flow field used in the development


of the splitting model is that of an elliptic hydrofoil of
1 m chord and 1.5 m half span subject to 2.88 m/s
inflow. The splitting model is derived from
observations of the equivalent radius1 at which a
bubble splits and the equivalent radii of the subbubbles. As shown in Figure 4, these radii form a
common set of curves regardless of the initial nucleus
size within the range of 10 to 100 m. We have also
found that if the curves are normalized by the
equivalent radius of the bubble just before the splitting,
the ratios are constant over a wide range of cavitation
numbers as shown in Figure 5. The ratio of the
maximum equivalent radius, Rmax, to the equivalent
radius just before the splitting is very close to 1.06, and
the radii of the two sub-bubbles just after the split are
0.95 Rmax and 0.55 Rmax regardless of the initial bubble
size. These observations are the basis for the bubble
splitting criterion and the initial size of the resulting
sub-bubbles that are used below.
The axial location of the sub-bubble centers
obtained from numerous computations with various
initial bubble nuclei size is shown in Figure 6. The
sub-bubble centers normalized by the equivalent radius
at splitting, Rsplit, fall on the same curves throughout
the range of cavitation numbers we have studied. The
center position of the downstream sub-bubble is

In our previous work (Choi and Chahine, 2003a, Choi,


Chahine, and Hsiao, 2003), we have developed a

1 The equivalent radius is defined as the radius of a sphere of

Maximum Acoustic Pressure at 1.53 m from the Tip


100

10

-1

P acoustic / P amb

10-2

10-3

10-4

10

-5

10-6
0.75

0.8

0.85

0.9

0.95

1.05

/ -Cpmin

Figure 2: The maximum acoustic pressure as a function of


cavitation number. Note that the bubble behavior near and
above the cavitation inception (approx. 2.58 in this case) is
quasi-spherical. (from Choi and Chahine (2004))

the same volume.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

constant and are -1.0 Rsplit and 4.4 Rsplit, respectively for
the upstream and the downstream sub-bubbles.

observed to be near 4.4 Rsplit, and the axial location of


the upstream sub-bubble varies from approximately
-1.0 Rsplit for the low cavitation numbers to 0.0 at the
cavitation number where splitting starts to occur.

Center of Gravity of Sub-bubbles, Re=2.88x10 6


6

Equivalent Radii, Re = 2.88x10 6, R o = 10, 20, 50, 100 m

CG sub 1 / R EQ

0.02

Maximum

CG / R EQ

0.015

10 m
20 m
50 m
100 m

R EQ [m]

at splitting

10 m
20 m
50 m
100 m

0.01

sub-bubble 1
0.005

1.6

1.8

CG sub 2 / R EQ

no splitting
for 100 m

sub-bubble 2
0
1.4

no splitting
for 50 m

2.2

2.4

2.6

-2
1.4

2.8

1.6

1.8

2.2

Cavitation Number

2.4

2.6

2.8

Figure 6: The axial locations of the sub-bubbles just after the


splitting normalized by the equivalent radius just before the
splitting, Rsplit. Four initial nuclei sizes, 10, 20, 50, and 100
m are considered.

Cavitation Number

Figure 4: The equivalent radii predicted by the axisymmetric


boundary element method (2DYNAFS) which allows bubble
elongation and splitting. Four initial nuclei sizes, 10, 20, 50,
and 100 m, are considered, and the maximum radii, radii at
splitting, and two radii of the sub-bubbles are shown.

Based on the observations of the bubble splitting


described above, the following bubble splitting model
for low cavitation numbers (or smaller nuclei) was
developed. The model consists of a splitting criterion
and initial conditions for the sub-bubbles.

Ratio of Radii, Re = 2.88x10 6, Ro = 10, 20, 50, 100 m


1.2

Bubble splitting criterion: After a bubble has grown


explosively (<inception), it will split at the beginning of
the collapse once its radius reaches 95% of the
maximum radius.

R max / R split
1

R sub 1 / R split
Radii Ratio

0.8

10 m
20 m
50 m
100 m

0.6

Initial condition of sub-bubbles: The radii of the


larger and smaller sub-bubbles are respectively, 0.95
and 0.55 of the radius at splitting, Rsplit. The initial gas
pressure in each sub-bubble is that of the original
bubble pressure at splitting, and the initial radial
velocities are zero. The initial location of the larger
sub-bubbles is 1.0 Rsplit upstream of the original bubble
center, and that of the smaller one is 4.4 Rsplit radii
downstream of the original bubble center.

R sub 2 / R split

0.4

0.2

0
1.4

1.6

1.8

2.2

Cavitation Number

2.4

2.6

2.8

Figure 5: The ratios of the maximum equivalent radius, Rmax,


and the equivalent radii of the two sub-bubbles, Rsub1, Rsub2,
relative to the equivalent radius just before the splitting, Rsplit,
for the four initial nuclei sizes, 10, 20, 50, and 100 m.

REENTRANT JET NOISE MODEL


The peak-to-peak values of the acoustic pressure
signal are shown in Figure 7. The acoustic pressure
data predicted with 10, 20, 50, and 100 m nuclei also
fall on a common curve regardless of the initial bubble
sizes. The maximum pressure peaks predicted by
2DYNAFS are always observed at the development of
the jets in the sub-bubbles following the bubble
splitting. We can use this observation to develop a
model for the reentrant jet noise.

The radial velocity of the sub-bubbles just after the


split was found to be close to zero for both subbubbles. It was also found that the initial locations of
the sub-bubbles depend weakly on the Reynolds
numbers but more strongly on the cavitation numbers.
For cavitation numbers in the mid-low range (1.7
2.2), the locations of the sub-bubbles relative
to the bubble center just before the splitting are fairly

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Pressure [Pa]

Field Point Pressure at Re=2.88x10 6, with Ro=10, 20, 50, and 100 m
10

10

10

10

10

10

10

Peak-to-peak pressure at (r,z)=(1.5297, 0.0) m

10 m
20 m
50 m
100 m

-1

1.4

1.6

1.8

2.2

2.4

2.6

2.8

Cavitation Number

Figure 7: Acoustic pressure signals predicted by the


axisymmetric boundary element method (2DYNAFS) which
allows bubble elongation and splitting. Data for the four
initial nuclei sizes, 10, 20, 50, and 100 m form a common
curve.

Figure 8: Typical pressure signal from the reentrant jet after


a bubble splitting.

Field Point Pressure for Re=2.88E06, R o=20m


10

Peak-to-peak pressure at (r,z)=(1.5297, 0.0) m

10

-2

10

-3

10

-4

10

-5

p ( ) = exp ( 2.5217 + 12.7754 ) ,


t1 ( ) = exp ( 2.7698 2.0205 ) ,

Pressure [Pa]

10

10

10

t1 = exp( -2.7698 - 2.0205 )


t2 = exp( -2.3703 - 5.7600 )

Time difference [s]

p = exp( -2.5217 + 12.7754 )

The modeling of the jet noise is based on the


observations of a typical acoustic pressure from the
reentrant jet as shown in Figure 8. To model the timing
of the jet noise, the time delay between the splitting
and the zero crossing of the jet noise is defined as t1,
and the time interval between the minimum and the
maximum of the jet noise is defined as t2. These two
time intervals and the peak-to-peak values of the
acoustic pressure signal are plotted for a range of
cavitation numbers in Figure 9. These data are fit into
exponential curves resulting in the following
expressions:

Pressure due to jet


Time from split to signal
Time from min to max
100
1.6

1.7

1.8

1.9

2.1

2.2

2.3

2.4

10-6
2.5

Cavitation Number

Figure 9: Exponential curve fit of the acoustic pressure from


the jet and the two time intervals of the jet noise.

(2)

t2 ( ) = exp ( 2.3703 5.7600 ) .

Then, the acoustic pressure signal due to the jet is


modeled with the following sine signal.
p jet ( t ) =

t t1 ( )
p ( )
sin
, t1 t2 t t1 + t2 (3)
2
t

(
)
2

The jet noise from this model is compared with the


original pressure signal for =2.10 in Figure 10. The
duration of the signal from the model is shorter than
the original signal at this cavitation number because
the curve fit for t2 underestimates the duration of the
signal as shown in Figure 9. We will improve this
approximation in our next studies.

Figure 10: Comparison of the pressure signal from the


reentrant jet between the 2DYNAFS prediction and the model
equation (3) at =2.10.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

SINGLE SPLITTING OF A NUCLEUS

8
6

Re=2.88x10 , R o=20m, =1.70

The improved SAP spherical model is first applied


to a 20 m nucleus at =1.70. The tip vortex flow field
used here is the same flow field in which the splitting
model was developed (Choi and Chahine, 2003a, Choi,
Chahine, and Hsiao, 2003). The resulting time history
of the bubble radius is shown in Figure 11 and the
associated acoustic pressure signal is shown in Figure
12. When the nucleus grows to its maximum size and
reaches 95% of the maximum radius, it splits into two
sub-bubbles as expected from the model explained
earlier. The larger sub-bubble shrinks monotonically
after the split, while the smaller sub-bubble begins with
a small growth followed by violent collapses and
rebounds. The acoustic pressure has an initial peak
corresponding to the initial growth of the nucleus,
followed by a sharp jet noise after the split. Then, there
are two contributions to the acoustic pressure, one for
each sub-bubble. The resulting signal from the subbubbles has a smaller maximum peak than that without
splitting.
Another simulation of 20 m nucleus at a higher
cavitation number, =2.10, is shown in Figure 13. In
this case, both of the sub-bubbles experience collapse
and rebound cycles. However, the combined pressure
signal still has smaller peaks than that from the
simulation without the splitting if the strong jet noise
of about 800 Pa appearing at 0.066 s is excluded. In
Figure 14, the bubble behavior at the cavitation
inception number is shown. In this case, the larger subbubble is more active, yet producing weaker signal
than the simulation without splitting.

0.01

Pressure [Pa]

4
2
0
-2
-4

no split
with split (sub-bubbles)
with split (sum)

-6
-8
0.05

0.075

0.1

0.125

Time [s]

0.15

0.175

Figure 12: Acoustic pressure signals predicted with 20 m


nucleus at =1.70.

Figure 13: Bubble radius and the acoustic pressure signal as


functions of time predicted with 20 m nucleus at =2.10.

Re=2.88x10 , R o=20m, =1.70

0.008
no split
with split

Radius [m]

0.006

0.004

0.002

0
0.05

0.075

0.1

0.125

Time [s]

0.15

0.175

Figure 11: Bubble radius history predicted with 20 m


nucleus at =1.70.

Figure 14: Bubble radius and the acoustic pressure signal as


functions of time predicted with 20 m nucleus at =2.26
(cavitation inception).

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

The acoustic pressure signal from this simulation is


shown in Figure 16. Compared to the acoustic signal
without splitting, the signal with splitting is much
noisier due to the jet noise, but the contribution from
the sub-bubble collapses is weaker than the signal
without splitting.

Based on the results of the three cavitation


numbers, we can arrive at the following conclusions
for a single split of a nucleus. Visual detection of
cavitation, which is based on detecting the maximum
bubble size, is not affected by the splitting because the
sub-bubble sizes are always smaller than the maximum
size observed before the inclusion of splitting in the
model and never grow larger than this. The acoustic
pressure signal with splitting has a strong jet noise, but
the resulting signals of the collapsing sub-bubbles are
smaller than such a signal of one big bubble without
splitting. This is more obviously seen in Figure 15,
where the maximum of the resultant acoustic pressure
signal from sub-bubble collapses for each cavitation
number is always slightly smaller than the maximum
peak observed from original SAP simulations without
splitting.
Field Point Pressure for Re=2.88E06, R o=20m
Peak-to-peak pressure at (r,z)=(1.5297, 0.0) m

Pressure [Pa]

104

10

10

10

10

10

-1

10

-2

1.6

Figure 16: Acoustic pressure signal from 100 nuclei of size


20 m with splitting compared to that without splitting.

no split
with split (jet noise)
with split (sub-bubbles)

1.7

1.8

1.9

2.1

2.2

2.3

2.4

Cavitation Number

Figure 15: Maximum of the acoustic pressure signal vs.


cavitation numbers predicted with 20 m nucleus.

MULTIPLE
SPLITTING

NUCLEI

WITH

SINGLE
Figure 17: Amplitude spectrum of the acoustic pressure
signal from 100 nuclei of size 20 m with splitting compared
to that without splitting. Also shown is the influence of the
inclusion of the reentrant jet noise on the results.

Simulations with multiple nuclei can provide an


insight for the cavitation noise in a real flow field. The
improved SAP spherical model is applied to the same
tip vortex flow field at =2.10 with a set of 100 nuclei
of the same size 20 m. The nuclei are distributed
randomly in space within a rectangular box of 0.1 m x
0.01 m x 0.006 m just in front of the blade tip and the
computation stops when all these bubbles have gone
through the 0.01 m x 0.006 m window.2 This
distribution corresponds to a void fraction of 5.59x10-7.

The amplitude spectrum can be obtained from the


pressure signal, p(t), through the finite Fourier
transformation.
T

P ( f , T ) = p ( t ) e i 2 ft dt

(4)

2 Note that the same nuclei population is assumed

where T is the time duration for the finite Fourier


transformation and f is the frequency. The amplitude

everywhere in the flow domain, but only the nuclei coming


through this window has the opportunity to grow explosively.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

spectrum of the acoustic signal is shown in Figure 17.


When the spectrum with splitting but without the jet
noise is compared with the spectrum with splitting, the
noise contents of frequency lower than 50~100 kHz
have decreased and those at higher frequency have
increased. This redistribution of the frequency contents
is due to the smaller sub-bubbles created by the bubble
splitting. If the jet noise is considered, the spectrum has
a peak at 25 kHz. This frequency corresponds to the
duration of the jet noise, 40 s, at this cavitation
number.

radii obtained for a cavitation number of 1.70 show


several radii curves due to successive splits. None of
the sub-bubbles grow larger than the first maximum
radius reached by the initial nucleus. The
corresponding acoustic pressure signal shown in Figure
20 becomes very complex with many new small
signals appearing as the sub-bubbles become smaller
and smaller. The amplitude spectrum of the combined
signal is shown in Figure 21. Compared to the
spectrum without splitting, the spectrum with multiple
splitting has richer contents over all frequencies above
a couple of hundreds kHz.

MULTIPLE SPLITTING
The bubble splitting model used above is based on
the non-spherical axisymmetric bubble simulations
starting from the first cycle of the bubble growth. In
such simulations, the numerical computations usually
cannot be continued through the collapse and rebound
of the sub-bubbles because of the touchdown of the
developed reentrant jets and the instabilities of the
collapsing bubbles. Therefore, the non-spherical
simulations cannot tell if the newly born sub-bubbles
would split again or not. The first cycle usually has the
strongest growth because it occurs just after the
nucleus passes the minimum pressure region of the tip
vortex. The bubble growth in the second and later
cycles is weaker than the first because the pressure
becomes higher and higher as the bubble moves
downstream. Would the sub-bubble placed a little
downstream of the minimum pressure region of the tip
vortex split again? This question can be answered by
performing a non-spherical simulation from the second
cycle of the SAP spherical simulation because the local
pressure encountered by the bubble nucleus a bubble
cycle downstream would be very similar to that of the
sub-bubble after the splitting.
In Figure 18, two 2DYNAFS simulations are
shown, one starting from the beginning of the first
cycle and the other from the beginning of the second
cycle of the SAP prediction. Even though the local
pressure encountered by the bubble in the second cycle
is higher than that of the first, the bubble elongates and
splits in the same way. The only major difference
between the two cases is the amount of the growth, i.e.
about 400 times in the first cycle vs. about twice in the
second cycle. The characteristics of the splitting in the
second cycle are very similar to those in the first cycle
as summarized in Table 1.
Because the sub-bubbles also split under very
similar criteria as the first splitting, we can apply the
same bubble splitting model to successive splitting of
the sub-bubbles. Results from such a simulation are
shown in Figure 19 for a 20 m nucleus. The bubble

0.01

Re = 2.88e6, R o = 20 m, = 2.10

2DynaFS
SAP
2DynaFS (2nd Cycle)

Equivalent Radius [m]

0.008

0.006

0.004

0.002

0
0.05

0.06

0.07

0.08

Time [sec]

0.09

0.1

Figure 18: Two 2DYNAFS simulations, one starting from


the beginning of the first cycle and the other from the second
cycle of the bubble growth predicted by SAP with 20 m
nucleus at =2.10.

Table 1: Characteristics of the bubble splitting in the first


and the second cycles.

Req,split / Req,max
Req,1 / Req,split
Req2 / Req,split

1st cycle
0.95
0.95
0.55

2nd cycle
0.94
0.97
0.43

The successive splits of sub-bubbles can also be


applied to simulations in the field of nuclei. Following
the nuclei distribution used by Hsiao and Chahine
(2003a), 142 nuclei of size distributions between 5 and
50 m, as shown in Figure 22, are randomly distributed
in the box-like volume of 0.1 m x 0.01 m x 0.006 m
just in front of the blade tip. The void fraction based on
the bubble distribution and the box-like volume is
1.46x10-6.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

50

Number of nuclei

40

Cavitating bubbles
Number of bubbles

30

20

10

0
5

10

15

20

30

40

50

Nuclei Size (m)

Figure 22: Distribution of 142 nuclei of sizes from 5 to 50


m and the resulting number of nuclei that cavitate in the
simulation.

Figure 19: Bubble radii of the sub-bubbles from a 20 m


nucleus with multiple splitting at =1.70.

Figure 23: Acoustic pressure signal from 142 nuclei of size


distribution from 5 to 50 m with multiple splitting at =2.10
compared to that without splitting.

Figure 20: Acoustic pressure signals of the sub-bubbles and


their sum from a 20 m nucleus with multiple splitting at
=1.70.

Figure 24: Amplitude spectrum of the acoustic pressure


signal from 142 nuclei of size distribution from 5 to 50 m
with multiple splitting at =2.10 compared to that without
splitting.

Figure 21: Amplitude spectrum of the acoustic pressure


signal from 20 m nucleus with multiple splitting at =1.70.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

The acoustic pressure signal obtained from the


simulation is shown in Figure 23, and its amplitude
spectrum is shown in Figure 24. The pressure signal
obtained from the distribution of nuclei sizes with
multiple splits has smaller peaks, and has increased
higher frequency contents (approximately higher than
10 kHz) and less low frequency contents
(approximately lower than 10 kHz) than that without
splits.

ACKNOWLEDGEMENT

CONCLUSIONS

REFERENCES

In order to further develop a tool to study scaling of


cavitation inception and noise, we have improved the
Surface Averaged Pressure (SAP) spherical model by
combining it with a bubble splitting model derived
from non-spherical bubble simulations. The method
accounts for a cascade of splits of the sub-bubbles and
includes the reentrant jet noise. During the
development of the model, we have found that the
bubble can lead to splitting even further downstream of
the minimum pressure region, allowing successive
multiple splitting of sub-bubbles. Our numerical
simulations show that at least the second splitting of
sub-bubbles has characteristics very similar to those of
the first splitting. This enhanced SAP model is applied
to single and multiple nuclei simulations in a typical tip
vortex flow field.
Through the applications of the enhanced SAP
model to various cases, we have concluded that the
visual detection of the cavitation is not affected by the
successive nuclei splitting because the sub-bubbles
never grow larger than the first maximum size within
the range of our study. However, the bubble splitting
has the following effects on the acoustic detection of
cavitation: (a) The noise from reentrant jets after the
split dominates the pressure signal from the rebounds
of the original nucleus and sub-bubbles created from it
through splitting. The jet noise contributes to a distinct
peak in the spectrum. (b) The resulting acoustic
pressure from the rebounds of the sub-bubbles is
smaller than the pressure from the rebounds of the
original nucleus without splitting. However, compared
to the spectrum without splitting, the sub-bubble noise
has an amplitude spectrum with more high frequency
content and less low frequency content. This trend of
shifting of frequency contents in the spectrum is a
direct result of the bubble population redistribution
from the large size original nucleus to many small size
sub-bubbles. We are currently working on the
expansion of the model and applications to a wider
range of cavitation numbers, initial nuclei distributions,
and Reynolds numbers.

Choi, J.-K. and Chahine, G.L., A Numerical Study on


the Bubble Noise and the Tip Vortex Cavitation
Inception, Journal of Ship and Ocean Technology,
Vol.7, No.3, Sep. 2003a, pp.13-33.

This work was conducted at DYNAFLOW, INC.


(www.dynaflow-inc.com). The research has been
supported by the Office of Naval Research under the
contract number N00014-04-C-0110 monitored by Dr.
Ki-Han Kim. The authors greatly appreciate this
support.

Choi, J.-K. and Chahine, G.L., Non-spherical Bubble


Behavior in Vortex Flow Fields, Computational
Mechanics, Vol. 32, No. 4-6, Dec. 2003b, pp. 281-290.
Choi, J.-K. and Chahine, G.L., Noise due to Extreme
Bubble Deformation near Inception of Tip Vortex
Cavitation, Physics of Fluids, Vol.16, No.7, Jul. 2004.
Choi, J.-K., Chahine, G.L., and Hsiao, C.-T.,
Characteristics of Bubble Splitting in a Tip Vortex
Cavitation, Proceedings of the 5th International
Symposium on Cavitation, CAV2003, Nov. 2003,
Osaka, Japan.
Hsiao, C.-T. and Chahine, G.L., Effect of Vortex /
Vortex Interaction on Bubble Dynamics and Cavitation
Noise, Proceedings of the 5th International
Symposium on Cavitation, CAV2003, Nov. 2003a,
Osaka, Japan.
Hsiao, C.-T. and Chahine, G.L., Prediction of Vortex
Cavitation Inception Using Coupled Spherical and
Non-Spherical
Models
and
Navier-Stokes
Computations, Journal of Marine Science and
Technology, 2003b (in print).
Hsiao, C.-T., Chahine, G.L., and Liu, H., Scaling
Effects on Prediction of Cavitation Inception in a Line
Vortex Flow, Journal of Fluid Engineering, Vol. 125,
2003, pp. 53-60.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Neil Bose
Memorial University, NL, Canada
Have the authors done work on the impact
and collapse of these free and split bubbles on a
surface such as a strut in the flow? Is your model
able to predict the pressures incident on the surface?
AUTHORS REPLY

Boundary Element Simulation of the ONR SnayGoertner Bubble Benchmark Problems, SAVIAC
Proceedings of the 67th Shock and Vibration
Symposium, Monterey, CA, 1996.
Chahine, G. L., Duraiswami, R., and Kalumuck, K.
M., Boundary Element Method For Calculating 2-D
and 3-D Underwater Explosion Bubble Loading On
Nearby
Structures
Including
Fluid-Structure
Interaction Effects, NSWC Dahlgren Division
Report NSWCDD/TR-93/46, 1997.

The authors would like to thank Prof. Bose


for his questions. We have done extensive studies on
the bubble behavior near solid surfaces in the past.
Particularly relevant examples are collapsing bubbles
near fixed and responding solid bodies (Chahine and
Perdue, 1989, Kalumuck et al., 1995, Chahine et al.,
1996, 1997) and spark generated bubbles between
walls (Choi and Chahine, 2003a, 2004). In the case
of an explosion bubble near a solid body, the bubble
behavior with reentrant jet and the impact pressure
due to the jet on the body surface were studied. Very
good correlation between the measured and predicted
time history of pressure was reported in Chahine et
al. (1997).
More recently, we observed spark generated
bubble behavior between two walls and the acoustic
noise from the splitting and subsequent reentrant jet
development as reported in Choi and Chahine
(2003a, 2004). The predicted bubble behavior was
well correlated to the observed behavior, and the
acoustic noise at the moment of splitting was
observed in the experiments.
REFERENCES
Chahine, G. L. and Perdue, T. O., Simulation of the
Three-Dimensional Behavior of an Unsteady Large
Bubble near a Structure, Drops and Bubbles, ed.
T.G. Wang, Proc. A.I.P. Conference, 197:169-187,
1989.
Chahine, G. L., Zhang, S., and Duncan, J. H., The
Final Stage of Bubble Collapse near a Rigid Wall,
Journal of Fluid Mechanics, Vol.257, pp147-181,
1993.
Kalumuck, K. M., Duraiswami, R., and Chahine, G.
L., Bubble Dynamics Fluid-Structure Interaction
Simulation by Coupling Fluid BEM and Structural
FEM Codes, Journal of Fluids and Structures,
Vol.9, pp.861-883, 1995.
Chahine, G. L., Prabhukumar, S., and Duraiswami,
R., Bubble Dynamics near A Cylindrical Body: 3-D

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Christopher Chesnakas
Naval Surface Warfare Center, Carderock Division,
USA
First, a comment: The acoustic spectra of the
splitting bubbles and the jet model looks most like
the experimental data, and I would recommend
concentrating future efforts on that model.
Second, a question: What happens to the
bubbles after the jet forms?
Is the bubble
annihilated? Or, are more bubbles formed, and does
the process continue?
AUTHORS REPLY
The authors would like to thank Dr.
Chesnakas for his valuable comment and questions.
We are glad to hear that the predicted cavitation noise
spectrum with the bubble nuclei splitting/jet model
looks like the experimental observation. We will
continue similar studies in the near future.
We have put efforts to continue the
simulation after the touchdown of the reentrant jet.
In axisymmetric simulations, the bubble turns into a
toroidal bubble and continues to oscillate responding
to the surrounding pressure changes. However, the
authors believe that the bubble could break into
smaller
bubbles
following
non-symmetrical
instabilities which we do not describe in the
axisymmetric code, but could study with our 3-D
code, 3DYNAFS. The break-up process may be
continued until the bubbles become very small and
the surface tension starts to overcome excessive
deformation.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics


St. Johns, Newfoundland and Labrador, CANADA, 8-13 August 2004

Hydroelastic Modeling of Surface-Piercing Propellers


Yin Lu Young, Princeton University, USA
Previous Experimental Investigations

ABSTRACT
A three-dimensional potential-based boundary
element method (BEM) is used for the hydrodynamic
analysis of unsteady ventilated flows around surfacepiercing propellers. For cases without blade vibration, the developed BEM is able to predict forces and
ventilation patterns that compared well with experimental measurements. However, as demonstrated
in past experimental and numerical studies, hydroelastic effects become important for surface-piercing
propellers operating at high advance coefficients. In
this work, two hydroelastic models are presented for
the approximation of the hydroelastic forces due to
blade vibration. Overviews of the formulation and
validation studies with experimental measurements
are presented.
INTRODUCTION
For high-speed vessels, surface-piercing (or
partially-submerged) propellers are more efficient
than fully submerged propellers due to the reduction
of appendage drag by elevating most of the propeller
assembly (e.g. shafts, struts, hub, etc.) above the
water surface. However, surface-piercing propellers
are not as commonly employed compared to waterjets because: 1) reduced efficiency when operating in
low-speed conditions, 2) high levels of blade stress,
especially at the blade entry phase where the impact
forces are mostly absorbed by the thin blade leading
edge, and 3) fatigue and vibration issues due to the
cyclic loading and unloading of the blades. These
problems are often a result of improper design of
the propulsion system due to the lack of a reliable
performance prediction method and the scarcity of
systematic performance data. On the other hand,
a properly designed propulsion system with surfacepiercing propeller(s) can be significantly more fuelefficient than other types of propulsors for high-speed
vessels. Thus, additional systematic experimental
studies and the development of a reliable performance prediction method are crucial to the design
and wide application of surface-piercing propellers.

Most of the past experimental studies focused


on the study of time-averaged thrust, torque, bending moment, and transverse forces.
Examples
include investigations by (Shiba 1953, Hadler &
Hecker 1968, Shields 1968, Hecker & Crown 1970,
Brandt 1973, Kruppa 1972, Hecker 1973, Alder &
Moore 1977, Rains 1981, Rose & Kruppa 1991,
Kruppa 1992, Rose et al 1993, Wang 1995, Ferrando & Scamardella 1996, Nozawa & Takayama
2002a). Based on past experimental studies, it
can be generally concluded that there exist four
major operating regimes for surface-piercing propellers: base-cavitating, partially-ventilated, transitional, and fully-ventilated:
Base-Cavitating: Mixture of vapor- and airfilled cavities that start near the blunt blade
trailing edge (i.e. most of the blade is fully wetted), and close before reaching the free surface
due to difficulty of the propeller in achieving the
design performance (Vorus 2003). As a result,
the efficiency of the propeller is significantly reduced due to high viscous and cavitation drag.
Partially-Ventilated: The air cavities start near
the blunt trailing edge and vent to the free surface. When the advance coefficient is reduced, a
partial cavity filled primarily with liquid vapor
may also develop. In this flow regime, the extent
and volume of the air cavity, as well as the timeaveraged thrust and torque coefficients, tend to
increase with decreasing advance coefficient.
Transitional: When the advance coefficient is
further reduced from the partially-ventilated
regime, the propeller enters the transition
regime. This flow regime is highly unstable,
and is often accompanied by violent oscillatory
forces. The air cavities start to spread toward
the blade leading edge and tend to fluctuate in
shape and size. In addition, a sudden drop in
thrust and torque coefficients often occur due to
the spread of air cavities toward the blade leading edge.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Fully-Ventilated: When the advance coefficient


is further reduced, the propeller enters the fullyventilated regime. This flow regime is characterized by continuous ventilated cavities that start
near the leading edge on the suction side of each
blade and vent to the atmosphere. This flow
regime is relatively stable and the blade trailing
edge remains ventilated at all times. The thrust
and torque coefficients tend to decrease with the
advance coefficient due to the dominance of cascade effects.
An illustrative drawing of the propeller
thrust/torque
characteristics
and
cavitation/ventilation patterns in the four major flow
regimes is shown in Fig. 1.
Recently, experimental studies have also been
conducted to examine the time-dependent hydrodynamic loads, and resulting stresses on the propeller
blades, shaft, and hull structure. Examples include
investigations by (Dobay 1970, Olofsson 1996, Miller
& Szantyr 1998, Dyson 2000, Dyson et al 2000,
Nozawa & Takayama 2002b). The objectives of the
studies were to investigate the physics behind the
large transverse forces and the source of vibratory
forces that are often observed in surface-piercing propellers. Based on these studies, it can be observed
that there exist four phases in each blade cycle: inair phase, blade-entry phase (from the entry of the
blade leading edge to the total submergence of the
entire blade), in-water phase, and blade exit phase
(from the initial exist of the blade leading edge to the
complete exit of the entire blade). The intensity and
the duration of the phases depend on the immersion
ratio and the operating flow regime. These studies
have also found that the large transverse forces experienced by surface-piercing propellers are due to the
asymmetry of the time-history of the horizontal and
vertical components of the tangential forces on the
individual blades, and the source of vibratory forces
is resonant blade vibration. Blade vibration can result in amplified fluctuations on measured forces and
ringing of the ventilated cavities. Examples of fluctuating blade forces (which increases with increasing
Froude number) and ringing pattern on the ventilated cavity surface due to blade vibration are shown
in Fig. 10 and 2, respectively.
These systematic studies are extremely valuable
to advance the understanding of the physics, and to
characterize and quantify the performance of surfacepiercing propellers. However, these studies are also
extremely expensive, difficult, and time consuming
to perform. The test must be carried out in a vari-

able pressure free-surface tunnel that permits highspeed operations to explore the various flow regimes.
The free surface must be clearly defined. A multicomponent dynamometer is needed to measure primary and secondary forces. Special equipments are
also needed to simultaneously provide realistic conditions for cavitation inception while maintaining constant water density. A high-performance blade dynamometer is needed to capture the effect of blade
vibration and time-dependent hydrodynamic loads.
Moreover, special considerations are needed to address scale issues concerning flow and blade vibration characteristics. Thus, there is also a strong
need for the development of reliable, versatile, and
robust computational tools to predict the hydrodynamic and hydroelastic performance of surfacepiercing propellers.

Previous Theoretical Hydrodynamic


Analysis
The first known method for the analysis
of surface-piercing propellers was developed by
(Yegorov & Sadovnikov 1961) using a blade element
method based on two-dimensional hydrofoil theory.
To simplify the analysis, they ignored the effect of adjacent blades, cavities, and wake vortex sheets (Olofsson 1996). Later, (Oberembt 1968) applied a lifting
line approach that included the effect of immersion,
but the propeller was assumed to be lightly loaded
such that no natural ventilation of the propeller and
its vortex wake occurred. Furuya developed a liftingline approach that included the effect of propeller
ventilation in (Furuya 1984, Furuya 1985). He applied the image method to account for free surface
effects, and assumed the face portion of the blades
to be fully wetted and the back portion of the blades
to be fully ventilated starting from the blade leading
edge. The blades were reduced to a series of lifting lines, and method was combined with a 2-D water entry-and-exit theory developed by (Wang 1977,
Wang 1979) to determine thrust and torque coefficients. Furuya compared the predicted mean thrust
and torque coefficients with experimental measurements obtained by (Hadler & Hecker 1968). In general, the predicted thrust coefficients were within acceptable range compared to measured values. However, there were significant discrepancies with torque
coefficients. Furuya (Furuya 1984, Furuya 1985) attributed the discrepancies to the effects of nonlinearity during the blade entry phase, absence of the blade
and cavity thickness representation in the induced
velocity calculation, uncertainties in interpreting the

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

experimental data, and limitations of the lifting-line


theory.
In 1991, (Vorus 1991) extended the conventional
propeller theory given in (Lewis 1989) to determine
steady forces and moments on a surface-piercing propeller. He assumed that the steady sectional lift coefficient (CL ) is linearly related to the increment
blade section angle of attack () as in the case of a
fully-ventilated flat plate: CL = 4 . The effects
of rake, skew, and inclination in the geometry were
considered. The method was shown to be an easy to
use design tool for trade-off studies and optimization
of blade geometry (Vorus 1991).
An unsteady lifting surface method was employed by (Wang et al 1990a) for the analysis of 3-D
fully ventilated thin foils entering into initially calm
water. The method was later extended by (Wang
et al 1990b, Wang et al 1992) to predict the performance of fully ventilated partially submerged propellers with its shaft above the water surface. Similar
to (Furuya 1984, Furuya 1985), the method assumed
the flow to separate from both the leading edge and
trailing edge of the blade, forming on the suction
side a cavity that vents to the atmosphere. The
predicted results were compared with experimental
measurements by (Hadler & Hecker 1968) and numerical predictions by (Furuya 1984, Furuya 1985).
The predictions were within reasonable agreement
with experimental values for a propeller with limited data range. However, substantial discrepancies
were observed with experimental values and numerical predictions by (Furuya 1984, Furuya 1985) for
another propeller.
A 3-D vortex-lattice lifting surface method developed by (Kudo & Ukon 1994, Kudo & Kinnas
1995) for the analysis of supercavitating propellers
was extended to treat surface-piercing propellers.
However, the method performed all the calculations
assuming the propeller to be fully submerged, then
multiplied the resulting forces with the propeller submergence ratio. As a result, only an estimate of the
mean forces was obtained while the complicated phenomena of blades entry to, and exit from, the water
surface were completely ignored.
A 2-D time-marching boundary element method
(BEM) was developed by (Savineau & Kinnas 1995)
for the analysis of the flow field around a fully ventilated partially submerged hydrofoil. The negative
image method was used to account for free surface
effects. However, this method only considered the
hydrofoils entry to, but not exit from, the water surface.
In all of the above mentioned numerical meth-

ods, the blades were assumed to be rigid, and thus


the effects of blade vibration were completely ignored.

Previous
Analysis

Theoretical

Hydroelastic

In the past, the modified cantilever beam theory (Taylor 1933) was the most commonly method
used to determine static blade stresses. The theory
assumed the blade to be a cantilever beam loaded by
thrust and torque distributed linearly over the radius
(Schoenherr 1963). Later, modifications were made
to include the effects of rake, skew, and centrifugal force (Morgan 1954, Schoenherr 1963, Atkinson
1968). However, the beam theory cannot accurately
predict the stress distributions for complex blade geometries due to its simplified assumptions. Later,
more sophisticated theoretical models based on the
measured or calculated mode shapes and resonance
frequencies of propeller blades in air were developed
by (Tsushima 1972, Brooks 1980). In 1981, (Murai & Shimizu 1981a, Murai & Shimizu 1981b) developed a linearized theory for the analysis of flow
around a chordwise flexible supercavitating hydrofoil when clamped at the trailing edge and when
supported elastically, but the method was limited to
2-D. Recently, finite element methods coupled with
hydrodynamic models have also been employed for
the analysis of dynamic blade stresses. These include the works of (Genalis 1970) using thin shell elements, (Atkinson 1973) using thick shell elements,
and (Kuo & Vorus 1985) using 3-D isoparametric
brick elements. Nonetheless, all of the above mentioned methods were developed for analysis of fully
submerged, non-cavitating propellers.
Hydroelastic analysis of cavitating propellers is
more complicated due to the presence of the unknown and time-dependent cavity surfaces, and the
interaction with elastic blade motion. The problem
is even more difficult for surface-piercing propellers
because both the added mass and damping effects
of the surrounding fluid vary with space (i.e. nodedependent) and blade angle (i.e. time-dependent)
since the blade enters and exists the fluid domain
in each revolution. However, most FEMs are not
designed to accommodate time-varying point masses
for each node. Thus, the space- and time-dependent
added forces due to fluid-structure interaction need
to be iteratively postulated as known values on the
right-hand-side of the equation of motion in the
FEM.
Recently, (Dyson 2000, Dyson et al 2000) pre-

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

sented a numerical model for the hydroelastic analysis of surface-piercing propellers. In their work, a 3-D
FEM (PAFEC) was used to model the structure. A
combination of 2-D semi-loof shell elements and 3-D
brick elements were used. The added mass effect of
the surrounding water was included by distributing
point masses across the surface of the blade according to a profile based potential flow theory around
a flat plate. The total damping ratio was taken to
be 0.05. However, to simplify the analysis, (Dyson
2000, Dyson et al 2000) used an assumed instead
of evaluated hydrodynamic load model. Thus, the
method cannot provide an accurate description of
the dynamic stresses.

OBJECTIVE
The ultimate objective of this work is to fully
couple a 3-D BEM with a 3-D FEM to determine
the time-dependent hydroelastic response of surfacepiercing propellers. The formulation and preliminary results for the 3-D BEM/FEM coupling are presented. To validate the coupling methodology, and
to access the sensitivity of the solution to the accuracy of added mass and hydrodynamic damping, a
simplified hydroelastic model using a single-degreeof-freedom (SDOF) system is introduced. The results shown that the single-degree-of-freedom system
provides a fast and practical mean of estimating the
hydroelastic response of surface-piercing propellers.

Hydrodynamic Analysis
In the present method, the hydrodynamic part
of the analysis is performed using a 3-D BEM. The
low-order potential based BEM was first developed
for the analysis of marine propellers in steady flow
by (Lee 1987, Kerwin et al 1987) and unsteady flow
by (Hsin 1990, Kinnas & Hsin 1992). The method
was then extended for the analysis of flow around
2-D partially cavitating and supercavitating hydrofoils (Kinnas & Fine 1991) and 3-D partially cavitating hydrofoils (Fine & Kinnas 1993). In (Kinnas
& Fine 1992), the method was named PROPCAV
(PROPeller CAVitation) for its added ability to analyze 3-D unsteady flow around cavitating propellers.
Later, (Mueller & Kinnas 1999) modified the method
to search for midchord cavitation on either the back
or the face of propeller blades. Most recently, the
method has been further extended to predict simultaneous face and back cavitation on conventional
fully submerged propellers (Young & Kinnas 2001a),

supercavitating propellers (Young & Kinnas 2001b,


Young & Kinnas 2003d), as well as surface-piercing
propellers (Young & Kinnas 2001b, Young & Kinnas
2003e) in non-axisymmetric inflow.
The method assumes the propeller to be a rigid
solid body which rotates at a constant angular velocity () in an unbounded fluid, and only sheet cavities (either vapor- or gas-filled) are considered. A
schematic drawing of the problem definition is shown
in Fig. 3.
The inflow wake is expressed in terms of the absolute (ship fixed) system of coordinates (X, Y, Z),
and is assumed to be the effective wake, i.e. it includes the interaction between the vorticity in the
inflow and the propeller (Kinnas et al 2000, Choi
2000). The inflow velocity, ~qin , with respect to the
propeller fixed coordinates (x, y, z), can be expressed
as the sum of the inflow wake velocity, ~qw , and the
propellers angular velocity
~ , at a given location ~x:
~qin (x, y, z, t) = ~qw (x, r, B t) + ~ ~x (1)
p
where r = y 2 + z 2 , B = arctan(z/y), and ~x =
(x, y, z). The resulting flow is assumed to be incompressible and inviscid. Hence, the total velocity, ~q,
can be expressed in terms of ~qin and the perturbation
potential :
~q(x, y, z, t) = ~qin (x, y, z, t) + (x, y, z, t)

(2)

where satisfies the Laplaces equation in the fluid


domain (i.e. 2 = 0). Note that the propeller fixed
coordinates system is used in analyzing the flow and
that the irrotational flow assumption is applied only
to the perturbation flow field.
The perturbation potential, p , at every point p
on the combined wetted blade surface SW B (t), ventilated cavity surface SC (t), and free surface SF (t),
must satisfy Greens third identity:


Z Z
G(p; q)
q (t)
2p (t) =
q (t)
G(p; q)
dS
nq (t)
nq (t)
S(t)
(3)
where S(t) SW B (t)SC (t)SF (t) is the combined
surfaced. The subscript q corresponds to the variable
point in the integration. G(p; q) = 1/R(p; q) is the
Greens function with R(p; q) being the distance between points p and q. ~nq is the unit vector normal
to the integration surface, with the positive direction
pointing into the fluid domain.

Boundary Conditions
The kinematic boundary condition requires the
flow to be tangent to the wetted portion of

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

the blade and hub surfaces, which forms a

:
Neumann-type boundary condition for n

= ~
qin ~n
n

(4)

The dynamic boundary condition requires that


the pressure everywhere on the ventilated cavity surface to be constant and equal to the atmospheric pressure, Patm . Defining a (Po
Patm )/( 2 n2 D2 ) = 0 as the ventilated cavitation
number, the total ventilated cavity velocity, ~qc ,
can be expressed as follows:
2

|~qc | = n2 D2 a + |~
qw | + 2 r2 2gY 2

(5)
t

where is the fluid density and r is the distance


from the axis of rotation. Po is the pressure far
upstream on the shaft axis; g is the acceleration of gravity and Y is the ship fixed coordinate. n = /2 and D are the propeller rotational frequency and diameter, respectively. It
can be shown that the dynamic boundary condition renders a Dirichlet boundary condition for
(Kinnas & Fine 1992, Young & Kinnas 2003e).
For ventilated cavities, a = 0 and thus the
first term at the right-hand-side of Eq. 5 goes to
zero. However, the first term is purposely left in
the equation to allow for future modeling of simultaneous vapor- and air-filled cavities, which
may occur in the based-cavitating or partiallyventilated flow regimes.
The linearized free surface boundary condition
and infinite Froude number assumption require
= 0 at the free surface, which implies that the
negative image method can be used to account
for the effect of the free surface. Consequently,
only vertical motions are allowed on the free surface.
On the ventilated cavity surface, the kinematic
boundary condition requires that the total velocity normal to the ventilated cavity surface to
be zero:
D
(n h(s, v, t))
Dt



+ ~qc (x, y, z, t) (n h(s, v, t))


t

= (6)
= 0

where n and h are the curvilinear coordinate


and ventilated cavity thickness normal to the
blade surface, respectively. It can be shown that

Eq. 6 renders a partial differential equation to


determine the normal thickness of the ventilated
cavities (Kinnas & Fine 1992, Young & Kinnas
2003e).

Ventilation Detachment Condition


The present BEM is able to search for the detachment locations of the ventilated cavity using an
iterative algorithm that is based on the smooth detachment assumption (which ignores viscous and surface tension effects):
1. The ventilated cavity has non-negative thickness
at its leading edge.
2. The pressure on the wetted portion of the blade
upstream of the ventilated cavity should be
greater than the prescribed cavity pressure.
3. During the blade exit phase, the ventilated cavities are required to detach at or aft of the intersection between the blade section and the free
surface.
In the current implementation, the cavities are assumed to vent to the free surface and the ventilated
cavities are only searched for on the suction side of
the blade surface. On the pressure side of the blade
surface, the ventilated cavities are assumed to detach
from the blade trailing edge. However, it is possible
to use the current algorithm to search for ventilated
cavity detachment locations on the pressure side of
the blade. In addition, the current algorithm can also
be modified to treat surface-piercing propellers operating in base-cavitating conditions by adjusting the
value of the cavity pressure in the dynamic boundary
condition.

Implementation
Since air loading of the propeller is small compared to the hydrodynamic load, Greens formula
(Eqn. 3) is only solved for the total number of submerged blade and wake panels. The values of and

n are set equal to zero on the blade and wake panels


that are above the free surface. The solution algorithm for partially submerged propellers is similar to
that for fully submerged supercavitating propellers,
which was given in detail in (Young & Kinnas 2002,
Young 2002, Young & Kinnas 2003a, Young & Kinnas 2003e).

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Hydroelastic Analysis
In this work, two approaches are presented for
the hydroelastic analysis. The first approach couples
the 3-D BEM with a consistent 3-D FEM to model
the fluid structure interaction. The blade is discretized into 3-D solid elements using the commercial
software ABAQUS. The fluid-structure coupling algorithm follows the linear field decomposition shown
in (Vorus 1981, Kuo & Vorus 1985). The problem is
solved in the time domain using an implicit Newmark
algorithm. In the second approach, the blade is simplified as a single-degree-of-freedom (SDOF) system
with a lumped mass, and the resulting ordinary differential equation representing the equilibrium equation of motion is solved using an explicit DormandPrince Runge-Kutta (4,5) formula in MATLAB. An
overview of the formulation for both approaches are
shown in this section.

BEM/FEM coupling
For hydroelastic analysis, the vibratory blade
motion is defined as displacements (~) superposed
on the rigid blade motion:

~ = ~x + ~(~
, t)

(7)

where
~ and ~x denote the position vectors to the deformed and undeformed blade surface, respectively.
Similar to the hydrodynamic formulation for
rigid blades, the perturbation flow field is assumed to
be incompressible, inviscid, and irrotational. Thus,
the perturbation velocity can be represented as the
gradient of the perturbation potential , where
2 = 0. Assuming linearity, can be decomposed
into two parts:
(~
, t) = (~x, t) + (~
, t)

(8)

where (~x, t) denotes the perturbation potential due


to rigid blades rotating in non-uniform wake, and
(~
, t) denotes the perturbation potential due to the
vibrating blades in uniform wake.
Applying Taylors expansion and ignoring
higher order terms, the perturbation velocity can be
approximated as follows(Young & Kinnas 2003c):
= +

(9)

Thus, the total velocity ~v can be written as:


~v (~
, t) = ~qin (~x, t) + (~
, t) = ~
q (~x, t) + (~
, t)
(10)

where ~q(~x, t) = ~qin (~x, t) + (~x, t) (i.e. Eqn. 2) is


the fluid velocity due to rigid blades rotating in nonuniform wake; (~
, t) is the fluid velocity due to
vibrating blades in uniform wake.
The total pressure (Pt ) acting normal to the
blade surface can be computed using Bernoullis
equation:
Pt

=
=

Pt Po
(11)


1
1 2 2
1 2
2
|~qw | + r gY
|~v |

2
2
t
2

where

t
|~v |2

=
=


+
t
t
|~q|2 + 2~q + ||2

(12)
(13)

The last term in Eqn. 13 can be dropped since the


vibratory displacements are assumed to be small.
Thus, Pt can also be decomposed into two parts:
Pt = P + Pv
where
P =

(14)


1
1
1 2
2
|~qw | + 2 r2 gY
|~q| (15)
2
2
t
2



Pv =
~q (16)
t

P is the hydrodynamic pressure if the blades were


perfectly rigid; Pv is the transient fluid pressure
distribution due to blade vibration.
Both and must satisfy the Laplace equation.
Hence, Greens third identity can also be applied to
determine the unknown values of . Applying the
negative image method, Eqn. 3 can be rewritten as
follows for :


Z Z
G(p; q)
q (t)
G(p; q)
dS (17)
2p (t) =
q (t)
nq (t)
nq (t)
S(t)


Z Z
G(p; q)
q (t)

q (t)
G(p; q)
dS
nq(t)
nq(t)
S(t)
The subscripts q and q correspond to integration
points on the real and image integration surfaces,
respectively. Note that the negative sign in front of
the second integral is due to the equal and opposite
strengths of the real and image singularities. Rewriting Eqn. 17 in matrix form yields:
 

[A] {} = [B]
(18)
n

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

where [A] and [B] are the dipole and source influence coefficient matrices, respectively. It should be
noted that the influence coefficient matrices are timedependent for surface-piercing propellers due to the
variation of the blade submergence with blade angle.
The advantage of this approach is that the influence
coefficients, [A] and [B], are the same as those for
the hydrodynamic analysis, thus result in substantial savings in CPU time and memory.
Multiplying both sides of Eqn. 18 by [A]1 yields
the solution for {}(Young & Kinnas 2003c):
 

(19)
{} = [C]
n
where [C] = [A]1 [B].
The values of
n on the right-hand-side (RHS)
of Eq. 19 is obtained by relating the perturbation
velocity normal to the vibrating blade surface,
n ,
to the normal component of the solid body velocity
at the element centroids (Young & Kinnas 2003c):
  ( ~ )

=
~n
(20)
n
t
Defining [T ] as the transformation matrix which
relates the normal velocities at the element centroids
to the element nodal velocities:
~
{
~n} = [T ]{u}

(21)

By combining Eqs. 19, 20, and 21, the following


expression for can be obtained:
{} = [C][T ]{u}

(22)

The unknown nodal velocities, {u},


of the solid
blade elements are determined by solving the dynamic equilibrium equation of motion:
[M ]{
u} + [D]{u}
+ [K]{u} = {F } + {f }

(23)

where [M], [D], and [K] are the structural mass,


damping, and stiffness matrices, respectively. {
u},
{u},
and {u} are the nodal acceleration, velocity, and
displacement vectors, respectively.
{F } is the nodal force vector due to rigid blades
rotating in nonuniform inflow, and {f } is the nodal
force vector due to blade vibration:
Z
{F } = [N ]T {P }dS
(24)
Z
{f } = [N ]T {Pv }dS
(25)

where [N ] is the shape function for transforming the


surface tractions (pressure acting normal to the elements) to consistent nodal forces.
{P } is the pressure vector due to rigid blades
rotating in non-uniform inflow. Thus, it is the same
as that computed from the BEM method as explained in the Hydrodynamic Analysis section. On
the other hand, {Pv } is the pressure vector due
to blade vibration, which depends on the unknown
structural displacements.
The partial time derivative of can be computed as follows:
 

= [C][T ]{
u}
(26)
t
Thus, the pressure vector due to blade vibration
(Eqn. 16) can be rewritten as follows:
{Pv } = [C][T ]{
u} [QD][C][T ]{u}

(27)

where [QD] is the matrix operator representing ~q .


The first term on the RHS can be seen as the added
mass, and the second term on the RHS can be seen
as the hydrodynamic damping. Thus, the dynamic
equilibrium equation of motion can be rewritten as
follows:
([M ] + [Mh ]) {
u} + ([D] + [Dh ]) {u}
+ [K]{u} = {F }
(28)
where [Mh ] is the hydrodynamic added mass and
[Dh ] is the hydrodynamic damping due to fluidstructure interaction:
Z
[Mh ] = [N ]T [C][T ]dS
(29)
Z
[Dh ] = [N ]T [QD][C][T ]dS
(30)
Note that both [Mh ] and [Dh ] depend on time because the influence coefficient matrix [C] changes
with blade submergence for surface-piercing propellers.
The natural frequencies and mode shapes of the
blades can be computed as follows:

2 ([M ] + [Mh ]) + ([D] + [Dh ]) + [K] {} = 0
(31)
where is the eigenvalue and {} is the eigenvector, the mode of vibration. Equation 31 needs to
be solved at each blade angle because both [Mh ]
and [Dh ] change with blade submergence for surfacepiercing propellers. Hence, the natural frequencies
and mode shapes also change with blade angle.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

BEM/SDOF coupling
The BEM/FEM coupling is not an easy task
due to the time- and space-dependent added mass
and hydrodynamic damping terms. Thus, in order
to validate the underlying coupling scheme, a much
simplified structural model using a SDOF system is
introduced. In this approach, the blade is simplified as a massless beam with a lumped mass SDOF
system, as shown in Fig. 4.
First, consider the SDOF system in air subject
to a transient load F . The displacement u is governed by the dynamic equilibrium equation of motion:
m
u + du + ku = F
(32)
where m, d, and k are the structural mass, damping,
and stiffness of the approximated blade in air. u,
u,
and u are the acceleration, velocity, and displacement, respectively, of the SDOF system. The natural frequency (dry ) and damping coefficient (dry )
of the SDOF system in air are defined as follows:
r
k
(33)
dry =
m
d
dry =
(34)
2mdry
Next, consider the same blade completely surrounded by water, the resulting equilibrium equation of motion is the same as Eq. 32 but with m
replaced by m + mh and d replaced by d + dh . mh
and dh represent the added mass and hydrodynamic
damping, respectively, when the blade is fully submerged. Thus, the natural frequency and damping
coefficient of the SDOF system in water are defined
as follows:
s
k
wet =
(35)
(m + mh )
wet

d + dh
2(m + mh )wet

(36)

Finally, if the submergence of the blade varies


with time (as in the case of a surface-piercing propeller), then both the added mass and hydrodynamic
damping will also vary with time (i.e. blade angle). Thus, the equilibrium equation of motion can
be written as:
[m + mh (t)] u
(t) + [d + dh (t)] u(t)
+ ku(t) = F (t)
(37)
Hence, the natural frequency and damping coefficient
also vary with time (i.e. blade angle) for a surface-

piercing propeller represented by a SDOF system:


s
k
(t) =
(38)
[m + mh (t)]
(t)

d + dh
2[m + mh (t)](t)

(39)

where (t) = dry and (t) = dry in the in-air


phase, and (t) = wet and (t) = wet in the inwater phase, as shown in Fig. 14. The values of dry
and wet can be obtained either from FEM frequency
analysis or from experimental excitation tests, and
standard values for dry and wet can be assumed.
During the blade entry and blade exit phase, it is
logical to assume that (t) and (t) are both smooth
functions that vary between the limiting values. In
particular, it is reasonable to assume that both (t)
and (t) vary like a sine function in the blade entry
and blade exit phase since the blade moves along the
arc defined by the propeller radius. Thus, using the
assumed functions for (t) and (t), the motion of
the SDOF can be obtained by solving Eq. 37. Finally,
the hydrodynamic force due to blade vibration, f (t),
can be approximated as follows:
f (t) = [mh (t)
u + dh (t)u]

(40)

The total hydrodynamic load is defined as F (t)+


f (t). F (t) is the hydrodynamic force for rigid blades,
which is computed using the BEM described in the
Hydrodynamic Analysis section.

Results
To validate the treatment of surface-piercing
propellers, numerical predictions for propeller model
841-B are compared with experimental measurements collected by (Olofsson 1996). A photograph
of the surface-piercing propeller and the velocity distribution at the propeller plane is shown in Figs. 5
and 6, respectively. The axial velocity is zero at the
free surface because a flat plate was placed in front
of the propeller to provide a well-defined free surface. Details of the experiments are given in (Olofsson 1996), and are summarized here for the sake of
completeness.

Summary of Experiment by Olofsson


In (Olofsson 1996), Olofsson presented a very
thorough series of experimental studies to determine
the time-averaged and dynamic performance of propeller model 841-B. The four-bladed high-speed partially submerged propeller was designed based on sea

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

trials on a board a 13 m twin screw planing test


craft. The diameter of the model-scale propeller is
250 mm. The experiments were conducted at the
KaMeWa free surface cavitation tunnel in Sweden.
The tunnel was equipped with a large de-aerating
chamber downstream of the test section. The deaerating chamber was used to eliminate the occurrence of air-liquid mixture in the test section when
testing to maintain constant water density. The tunnel has also been equipped with a special device for
micro air bubble seeding to provide constant cavitation inception as well as realistic propeller thrust
and torque without hysteresis in partially cavitating
conditions. The blade dynamometer was a 4-bladed,
single flexure, 5-component dynamometer developed
by KaMeWa. A schematic diagram of the flexureunit used to measure the reaction load is shown in
Fig. 7. The blade-flexure dynamometer was especially designed to capture the true effect of fundamental blade vibration by keeping the fundamental vibration frequency of the bare flexure and the
fundamental vibration frequency of the fixed-to-hubattached (or rigid-attached) blade to be well above
unity. The sum of the hydrodynamic load plus the
inertial and gravitational load due to true blade vibration is obtained by subtracting the inertial and
gravitation load imposed by the propeller rotation
and rigid-blade motion (due to relative displacement
and rotation of the flexure unit) from the reaction
load measured by the blade-flexure dynamometer.
In the experiment, only one blade tip immersion ratio (h/D = 0.33) was considered, and the influence of Froude and cavitation number at different
advance coefficients, J = V /nD, was systematically
investigated. Tests with different shaft yaw and inclination angles were also performed. Three scale
models were examined: (ratio of full-scale diameter and model scale diameter) = 1, 3, and 9. The
model scale was require to have the same cavitation
2
number (
= (Po Pv )/0.5V ) and Froude number
(Fr = V / gD) as the full scale:
=

Patm Pv 1 1
2
0.5gD
Fr

(41)

where Pv is the vapor pressure, and D is the model


scale diameter.
To verify the finite element model, shake tests
at partial immersion were conducted to measure the
natural frequencies of the blade-flexure. An actual
model 841-b was manufactured, and the dynamometer was clamped to a rigid diving apparatus and the
blade-flexure was set at different angular positions
with the blade being immersed in water step-by-

step. The measured natural frequencies of the bladeflexure versus angular position is shown in Fig. 8.
The measured fundamental frequencies of the bladeBF
flexure in air and in water are dry
= 1013 Hz and
BF
wet = 425 Hz, respectively. Olofsson also calculated the fundamental frequencies of the bare flexure
F
BF
(dry
) and the blade-flexure (dry
) in air using FEM,
F
BF
and found dry = 2457 Hz and dry
= 1043 Hz, respectively. He also found that the ratio of the fundamental frequency of the bare flexure and the rigidlyF
B
attached blade in air, dry
/dry
, to be approximately
2.1. In the current hydroelastic model, the blades are
assumed to be rigidly-attached at the root. Assuming the ratio of the fundamental frequencies of the
rigidly-attached blade and the blade-flexure in air is
B
BF
B
BF
the same as in water (i.e. dry
/dry
= wet
/wet
),
the measured fundamental frequencies of the rigidlyattached blade in air and in water can be inferred as
B
B
dry
= 1138 Hz and wet
= 477Hz, respectively.

BEM results
The predicted and measured time-averaged
thrust (KT ), torque (KQ ), and efficiency () of propeller model 841-B in the ship-fixed coordinates is
shown in Fig. 9. The lines and symbols in Fig. 9
represent the numerical predictions and experimental
measurements, respectively. For each advance coefficient, multiple experimental data points are shown
to depict the effect of Froude number and scale ratio.
The time-averaged results compared well with experimental values because the discrepancies caused by
resonant blade vibration act as fluctuations superimposed on the basic load in the experimental values,
which will not appear in the time-averaged results
due to the time-averaging process.
Comparisons of the predicted and measured
time-history of the individual dynamic blade loads
for J = 1.0 and J = 1.2 are shown in Figs. 10 and
11, respectively. Also shown in Figs. 10 and 11 are
the effect of Froude number, Fr , onthe measured dynamic blade loads. Since Fr = V / gD, the advance
speed V increases with increasing Fr , and thus the
system is more susceptible to resonant blade vibration. This is evident via the increase in the magnitude of the humps (amplified fluctuations superimposed on the basic load) with increasing Froude
number for J = 1.0, as shown in Fig. 10. However,
the use of the BEM alone will not be able to capture the humps due to the rigid blade assumption.
Thus, more discrepancies can be observed at higher
Froude numbers for the same advance coefficient. On
the other hand, for a fixed value of V (or Fr ), the pro-

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

peller rotational frequency n decreases with increasing advance coefficient J = V /nD. Hence, for the
same Fr , the susceptibility of the blade to resonant
vibration decreases with increasing J. To demonstrate that the humps observed on the time-history
of the blade loads are due to resonant blade vibration, the harmonic axial coefficients normalized by
the mean value are plotted against harmonics of the
propeller angular velocity n = n /2 in Fig. 12. As
expected, the harmonic coefficients are amplified at
the frequency corresponding to the fundamental frequency of the blade in water.
The comparison of predicted and observed ventilation patterns at J = 1.2 and Fr = 6 are shown
in Fig. 13. The ventilation patterns compared very
well because the blade is behaving like a rigid body
at higher advance coefficients.

BEM/SDOF results
Since the SDOF model assumed the blade to
be rigidly attached to the hub, dry and wet are
taken as 1138 Hz and 477 Hz, respectively. The assumed variation of the fundamental frequency of the
blade with blade angle is depicted in Fig. 14 along
the measured frequencies. The damping coefficients
in air and in water, dry and wet , were assumed to be
0.03 and 0.05, respectively. (t) is assumed to vary
in a similar trend as (t). It should be noted that
wet = 0.05 is a reasonable value based on the study
of propeller excitation in water by (Hylarides 1978).
An example of the predicted motion of the SDOF
system for J = 0.8 is shown in Fig. 15. The predicted
hydrodynamic load due to the rigid blade (F (t)),
transient load due to the vibrating blade (f (t)), and
total hydrodynamic load (F (t) + f (t)) are compared
with experimental measurements for different values
of advance coefficient (J) and Froude number (Fr ) in
Figs. 16 to 22. In general, the results agreed well with
experimental measurements, and that the inclusion
of the added forces due to hydroelastic effects did improved the numerical prediction. Most importantly,
the maximum axial force coefficient predicted using
the present method compared well with experimental
measurements.
It should be noted that discrepancies at the
blade entry phase and the blade exit phase for J =
0.8 shown in Figs. 16 and 17 are due to increase in
intensity and height of jet sprays, and rise in the
overall free surface elevation due to cavity displacement effects. The later is evident in the experimental measurement shown in Figs. 16 and 17, where
the blade appears to carry load from approximately

68 degrees to 290 degrees while the designed loading


region is from 88 degrees (when the blade leading
edge first enters the water) to 272 degrees (when the
blade completely exists the water). This clearly indicates an increase in overall free surface elevation. In
addition, more discrepancies can be observed at the
blade entry phase than at the blade exit phase. This
is due to the fact that strong jets develop near the
blade leading edge at the instant of impact, which
lead to very high slamming forces at the blade entry
phase. However, the effects of jet sprays and increase
in overall free surface elevation cannot be captured in
the current model due to the application of the negative image method in the BEM. This implies that
the prediction of F (t), the rigid-blade hydrodynamic
load, must first be improved to account for nonlinear
free surface effects in order to improve the prediction
of total hydrodynamic blade loads, F (t) + f (t). The
hydrodynamic load due to resonant blade vibration,
f (t), represents the fluctuating component of the total hydrodynamic load, and thus only acts in the
loading region defined by the BEM.
To validate the BEM/SDOF model, an ideal
F (t) which resembles the measured load minus the
fluctuations due to resonant blade vibration for J =
0.8 is applied. A plot of the ideal rigid-blade hydrodynamic load along with the measured total hydrodynamic forces at different Froude numbers are
shown in Fig. 23. Also shown in Fig. 23 is the
rigid-blade hydrodynamic load obtained from the
BEM for comparison purposes. Using the ideal F (t),
the predicted total hydrodynamic load for J = 0.8
at Fr = 4 and Fr = 6 are compared with the
measured values in Fig. 24. As shown in the figure, the amplitude and the frequency of the humps
compared much better with experimental measurements. This quick study validates the use of the
BEM/SDOF model to estimate the hydroelastic axial force due to blade vibration. Since the underlying
concept of the BEM/SDOF coupling is the same as
the BEM/FEM coupling, this study also validates
the presented BEM/FEM coupling algorithm. Another important implication of the result is that improving the BEM model to account for nonlinear free
surface effects is crucial to obtain accurate prediction of the hydro- and elasto-performance of surfacepiercing propellers at low advance coefficients.
To investigate the sensitivity of the solution to
the damping coefficient, comparison of the predicted
total hydrodynamic force coefficient for different values of dry and wet is shown in Fig. 25. As expected, the solution is more sensitive to changes in
the hydrodynamic damping, wet , than the struc-

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

tural damping, dry . To investigate the sensitivity


of the solution to the fundamental frequency in water (wet ), comparison of the predicted total hydrodynamic axial force coefficient for different values of
wet is shown in Fig. 26. As expected, the amplitude
and the width of the humps increased with decreasing wet . Comparison of Figs. 25 and 26 indicates
that the solution is more sensitive to changes in wet
than wet . In addition, the results suggest that the
assumed values of dry = 0.03 and wet = 0.05 are
reasonable. Thus, for the fully BEM/FEM coupling,
the structural and hydrodynamic damping matrices
can be lumped and expressed as a function of the
total mass and stiffness matrix:
([D] + [Dh ]) = ([M ] + [Mh ]) + [K]

(42)

where and are the Rayleigh damping factors. To


account for the variation of [Dh ] with time, should
also vary with time. The application of Eq. 42 will
significantly simplify the modal analysis and transient calculation for the finite element method.

of the flexure unit, which is not considered in the


current FEM model.
To evaluate the natural frequencies of the blade
in water, the global added mass matrix, Eq. 29, is
computed via the BEM analysis. Contrary to the
structural mass matrix ([M ]), the added mass matrix ([Mh ]) is not symmetric and is fully populated.
To simplify the analysis, the added mass matrix is
lumped at each BEM node for each degree of freedom k by multiplying the consistent added mass matrix with the unit acceleration vector corresponding
to that degree of freedom:
MhL |i,k =

3N
XN
j=1

where
n
o 
j,1 = 1 0

n
o 
j,2 = 0 1

n
o 
j,3 = 0 0

j,k ;
Mh |i,j

(43)

1 0

0 1

0 0

0 1

0 0

1 0

0 0

1 0

0 1

BEM/FEM results
To accommodate the classic sharp leading edge
and thick trailing edge of surface-piercing propellers,
a combination of 15-node quadratic triangular prisms
and 20-node quadratic bricks are used to represent
the blade in the FEM analysis. Quadratic elements
were selected to avoid FEM problems due to shear
locking (occurs in first-order, fully integrated elements subject to bending due to numerically induced
shear strains) and hourglassing (occurs in first-order,
reduced-integration elements in stress and displacement analysis due to zero strain at the integration
point). To reduce the computational time, reduced
integration is employed for the 20-node quadratic
bricks. To avoid errors due to interpolation, the corner nodes of the FEM elements correspond to the
corner nodes of the BEM elements. The dependence
of the predicted blade frequency in air for propeller
M841B is shown in Fig. 27. The first number indicates the number of BEM elements in the chordwise
direction (which correspond to two times the number
of FEM elements), and the second number indicates
the number of BEM (and number of FEM) elements
in the spanwise direction. Also shown in Fig. 27 are
the measured natural frequencies in air obtained by
Olofsson (Olofsson 1996). As shown in the figure,
the numerical predictions converged and compared
well with experimental measurements. Measured frequencies beyond the third mode are not shown in the
figure because those are influenced by the flexibility

k = 1, 2, 3

T
T
T
(44)

and N N is to total number of BEM nodes. Thus,


the non-zero entries in the resultant diagonal lumped
added mass matrix can be written as:

MhL |1,1 MhL |1,2 MhL |1,3 MhL |2,1 MhL |2,2 MhL |2,3
MhL |N N,1

MhL |N N,2

MhL |N N,3

It should be noted that the hydrodynamic damping matrix, [Dh ], is currently neglected in the modal
analysis. Comparison of the predicted versus the
measured natural frequencies for the first two modes
as a function of the blade angle is shown in Fig. 28.
The natural frequencies are normalized by the fundamental frequency in air. As shown in the figure,
the predicted natural frequencies compared well with
experimental measurements.

CONCLUSIONS
A 3-D BEM has been extended for the analysis
of surface-piercing propellers. Numerical and experimental validation studies of the BEM for surfacepiercing propellers were presented in (Young & Kinnas 2003a, Young & Kinnas 2003e, Young & Kinnas
2003c, Kinnas & Young 2003). Due to the rigid blade
assumption, the BEM alone cannot capture the effect of blade vibration. To account for the fluidstructure interaction, two hydroelastic models are

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

presented. The first model couples the BEM with


a 3-D FEM, and the second model couples the BEM
with a SDOF system. The predicted natural frequencies as a function of blade angle using the BEM/FEM
approach compared well with experimental measurements. The predicted time-dependent axial force coefficient using the BEM/SDOF approach also compared well with experimental measurements. However, there are some discrepancies at the blade entry and exit phase for cases with low advance coefficient due to the effects of jet sprays, and rise in
overall free surface elevation. Results indicated that
1) the presented hydroelastic models are valid, 2) the
BEM model must be improved to account for nonlinear free surface effects in order to obtain a better estimate of the performance of surface-piercing
propellers at low advance coefficients, 3) hydroelastic forces are more sensitive to changes in the added
mass than the hydrodynamic damping, and 4) despite the simplicity of the BEM/SDOF model, it provides a practical mean of estimating the axial forces
due to blade vibration. Additional validations for
the full BEM/FEM approach are underway. Also in
progress is the modeling of nonlinear free surface effects at the entry phase using a 2-D BEM with the
exact free surface boundary conditions. The ultimate
goal of this research is to develop a robust and reliable tool to predict the hydro- and elasto-dynamic
performance of surface-piercing propellers.
REFERENCES
Alder, R. and Moore, D. 1977 Performance
of an inclined shaft partially-submerged propeller
operating over a range of shaft yaw angles. Technical report. SPD-802-01. Naval Ship Resesarch and
Development Center.
Atkinson, P. 1968 On the choice of method for the
calculation of stress in marine propellers. Transactions, Royal Institution of Naval Architecture,
110.
Atkinson, P. 1973 The prediction of marine
propeller distortion and stresses using a superparametric thick shell finite-elelement model.
Transactions, Royal Institution of Naval Architecture, 115.
Brandt, H. 1973 Modellversuche mit schiffspropellern an der wasseroberfl
ache. Schiff und Hafen,
25, 5, pp. 415422.
Brooks, J. 1980 Vibrations of a marine propeller operating in a nonuniform inflow. Technical
report. Report DTNSRDC-80/056. David Taylor
Research Center. April.

Choi, J. 2000 Vortical inflow propeller interaction using unsteady three-dimensional euler solver.
Doctoral dissertation, Department of Civil Engineering, The University of Texas at Austin. August.
Dobay, G. 1970 Unsteady blade force measurements on a skewed partially-submerged propeller.
Technical report. 392-H-01. Naval Ship Resesarch
and Development Center. June.
Dyson, P. K. 2000 The modelling, testing and
design, of a surface piercing propeller drive. Doctoral dissertation, Department of Mechanical and
Marine Engineering, Plymouth University. October.
Dyson, P. K., Chudley, J. and Grieve, D.
2000 An experimental programme to determine
the mean and time varying loads imposed by surface piercing propellers. Sea Australia 2000, Sydney.
Ferrando, M. and Scamardella, A. 1996 Surface piercing propellers: Testing methodologies,
results analysis and comments on the open water
characteristics. Proceedings: Small Craft Symposium.
Fine, N. and Kinnas, S. 1993 A boundary element method for the analysis of the flow around 3d cavitating hydrofoils. Journal of Ship Research,
37, September, 213224.
Furuya, O. 1984 A performance prediction theory for partially submerged ventilated propellers.
Proceedings: Fifteenth Symposium on Naval Hydrodynamics.
Furuya, O. 1985 A performance prediction theory for partially submerged ventilated propellers.
Journal of Fluid Mechanics, 151, pp. 311335.
Genalis, P. 1970 Elastic strength of propellers an analysis by matrix methods. Technical report.
Report 3397. David Taylor Research Center.
Hadler, J. and Hecker, R. 1968 Performance of
partially submerged propellers. Proceedings: The
7th ONR Symposium on Naval Hydrodynamics.
August.
Hecker, R. 1973 Experimental performance of
a partially submerged propeller in inclined flow.
SNAME Spring Meeting, Lake Buena Vista, FL,
April.
Hecker, R. and Crown, D.
1970
Performance characteristics of partially-submerged propeller 4281 with varying number of blades at low
advance coefficients. Technical report. 249-H-12.
Naval Ship Resesarch and Development Center.
November.
Hsin, C.-Y.
1990
Development and analysis

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

of panel method for propellers in unsteady flow.


Doctoral dissertation, Department of Ocean Engineering, Massachusetts Institute of Technology.
September.
Hylarides, S. 1978 Model tests on hydrodynamics of propeller blade vibrations. Proceedings:
Propeller 78 Symposium, SNAME.
Kerwin, J., Kinnas, S., Lee, J.-T. and Shih,
W.-Z. 1987 A surface panel method for the hydrodynamic analysis of ducted propellers. Trans.
SNAME, 95.
Kinnas, S. and Fine, N.
1991
NonLinear Analysis of the Flow Around Partially or
Super-Cavitating Hydrofoils by a Potential Based
Panel Method. Proceedings: Boundary Integral
Methods-Theory and Applications, Proceedings of
the IABEM-90 Symposium, Rome, Italy, October
15-19, 1990. Springer-Verlag, Heidelberg, 289300.
Kinnas, S. and Fine, N. 1992 A nonlinear
boundary element method for the analysis of unsteady propeller sheet cavitation. Proceedings:
Nineteenth Symposium on Naval Hydrodynamics.
August, 717737.
Kinnas, S. and Hsin, C.-Y. 1992 A boundary
element method for the analysis of the unsteady
flow around extreme propeller geometries. AIAA
Journal, 30, 3, March, 688696.
Kinnas, S. and Young, Y. 2003 Modeling of
cavitating or ventilated flows using bem. International Journal of Numerical Methods for Heat &
Fluid Flow on the BEM, 13, 6, pp. 672697.
Kinnas, S., Choi, J., Lee, H. and Young,
J. 2000 Numerical cavitation tunnel. Proceedings: NCT50, International Conference on Propeller Cavitation. April 3-5.
Kruppa, C. F. L. 1972 Testing partially submerged propellers. Proceedings: 13th ITTC Report of Cavitation Committee. Appendix V.
Kruppa, C. F. L. 1992 Testing surface piercing
propellers. Proceedings: Hydrodynamics : Computations, Model Tests, and Reality. May, pp. 107
113.
Kudo, T. and Kinnas, S. 1995 Application of vortex/source lattice method on supercavitating propellers. Proceedings: 24th American Towing Tank
Conference. November 2-3.
Kudo, T. and Ukon, Y.
1994 Calculation
of supercavitating propeller performance using a
vortex-lattice method. Proceedings: Second International Symposium on Cavitation. April 5-7,
403408.
Kuo, J. and Vorus, W. 1985 Propeller blade dynamic stress. Proceedings: Tenth Ship Technology

and Research (STAR) Symposium. May 21-24, pp.


3969.
Lee, J.-T. 1987 A potential based panel method
for the analysis of marine propellers in steady flow.
Doctoral dissertation, Department of Ocean Engineering, Massachusetts Institute of Technology.
August.
Lewis, E. 1989 Vibration. Principles of Naval
Architecture, 2, Chapter 7.
Miller, W. and Szantyr, J. 1998 Model experiments with surface piercing propellers. Ship
Technology Research, 45, February, pp. 1421.
Morgan, W. 1954 An approximate method of
obtaining stress in a propeller blade. Technical report. Report 919. David Taylor Research Center.
Mueller, A. and Kinnas, S. 1999 Propeller
sheet cavitation predictions using a panel method.
Journal of Fluids Engineering, 121, June, 282
288.
Murai, H. and Shimizu, S. 1981a Hydroelastic characteristics of flexible supercavitating hydrofoils. report 1. two-dimensional theory at zero
cavitation number when clamped at trailing edge.
Bulletin of the JSME, 24, August, pp. 13791387.
Murai, H. and Shimizu, S. 1981b Hydroelastic
characteristics of flexible supercavitating hydrofoils. report 2. two-dimensional theory at zero cavitation number when supported elastically. Bulletin
of the JSME, 24, August, pp. 13881396.
Nozawa, K. and Takayama, N. 2002a Experimental study on propulsive performance of surface
piercing propeller. Journal of the Kansai Society
of Naval Architects.
Nozawa, K. and Takayama, N. 2002b Hydrodynamic performance and exciting force of surface
piercing propeller. Proceedings: Asia Pacific Workshop on Marine Hydrodynamics (APHydro 2002).
May 21-23.
Oberembt, H. 1968 Zur bestimmung der instationaren fl
ugelkrafte bei einem propeller mit aus
dem wasser herausschlagenden fl
ugeln. Technical
report. Inst. f
ur Schiffau der Universitat Hamburg,
Bericht Nr. 247. Juli.
Olofsson, N. 1996 Force and flow characteristics
of a partially submerged propeller. Doctoral dissertation, Department of Naval Architecture and
Ocean Engineering, Chalmers University of Technology, Goteborg, Sweden. February.
Rains, D. A. 1981 Semi-submerged propellers for
monohull displacement ships. Proceedings: Propeller 81 Symposium. Society of Naval Architects
and Marine Engineers, Virginia Beach, VA, May
26-27, pp. 1540.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Rose, J. C. and Kruppa, C. F. L. 1991 Surface


piercing propellers - methodical series model test
results. Proceedings: FAST91. June.
Rose, J. C., Kruppa, C. F. L. and Koushan, K.
1993 Surface piercing propellers - propeller/hull
interaction. Proceedings: FAST93. December, pp.
867881.
Savineau, C. and Kinnas, S. 1995 A numerical
formulation applicable to surface piercing hydrofoils and propellers. Proceedings: 24th American
Towing Tank Conference.
Schoenherr, K. 1963 Formulation of propeller
blade strength. Proceedings: Spring Meeting, The
Society of Naval Architects and Marine Engineers.
April 25-27.
Shiba, H. 1953 Air-drawing of marine propellers.
Technical report. 9. Transportation Technical Research Institute. August.
Shields, C. 1968 Performance characteristics of
several partially submerged supercavitating propellers. Technical report. 2723. Naval Ship Resesarch and Development Center. July.
Taylor, D. 1933 The speed and power of ships.
Technical report. U.S. Governmnet Printing Office, Wahsington, D.C.
Tsushima, H. 1972 Dynamic response of an elastic
propeller to a non-uniform inflow. Doctoral dissertation, Department of Aerospace Engineering,
Pennsylvania State University.
Vorus, W. S. 1991 Forces on surface-piercing propellers with inclination. Journal of Ship Research,
35, 3, pp. 210218.
Vorus, W. 1981 Hydrodynamic added-mass matrix of vibrating ship based on a distribution of
hull surface sources. SNAME Transactions, 89,
pp. 397416.
Vorus, W. 2003 Discussion to the authors paper
presented at the 2003 propeller and shaft symposium. September 17-18.
Wang, D. 1977 Water entry and exit of a fully
ventilated foil. Journal of Ship Research, 21, pp.
4468.
Wang, D. 1979 Oblique water entry and exit of a
fully ventilated foil. Journal of Ship Research, 23,
pp. 4354.
Wang, G., Zhu, X. and Sheng, Z. 1990a Hydrodynamic forces of a three-dimensional fully ventilated foil entering water. Journal of Hydrodynamics, 5, 2.
Wang, G., Jia, D. and Sheng, Z. 1990b Hydrodynamic performance of partially submerged ventilated propeller. Shipbuilding of China.
Wang, G., Jia, D. and Sheng, Z. 1992 Study on

propeller characteristics near water surface. Proceedings: The 2nd Symposium on Propeller and
Cavitation. September 1-4, pp. 161168.
Wang, S. Y. 1995 Systemtische analyse von modellversuchen mit teilgetauchten propellrn. Doctoral dissertation, Technischen Unversitaet Berlin
(D83).
Yegorov, I. and Sadovnikov, Y. 1961 Effect of instability on hydrodynamic characteristics of a propeller cutting the water surface. Sudostroyenige, pp. 1517.
Young, Y. 2002 Numerical modeling of supercavitating and surface-piercing propellers. Doctoral dissertation, Department of Civil Engineering, The University of Texas at Austin. May.
Young, Y. and Kinnas, S. 2001a A BEM for
the prediction of unsteady midchord face and/or
back propeller cavitation. Journal of Fluids Engineering, 123, June, pp. 311319.
Young, Y. and Kinnas, S. 2001b Numerical
modeling of supercavitating and surface-piercing
propeller flows. Proceedings: CAV 2001: Fourth
International Symposium on Cavitation. California Institute of Technology, Pasadena, CA, June
20-23.
Young, Y. and Kinnas, S.
2002
A BEM
technique for the modeling of supercavitating and
surface-piercing propeller flows. 24th Symposium
on Naval Hydrodynamics, July 8-13.
Young, Y. and Kinnas, S. 2003a Analysis of supercavitating and surface-piercing propeller flows
via bem. Journal of Computational Mechanics, 32,
5-6.
Young, Y. and Kinnas, S. 2003b Fluid and structural modeling of cavitating propeller flows. Proceedings: Fifth International Symposium on Cavitation (CAV2003). November 1-4.
Young, Y. and Kinnas, S. 2003c Numerical
analysis of surface-piercing propellers. Proceedings: 2003 Propeller and Shaft Symposium. Society of Naval Architects and Marine Engineers,
Virginia Beach, VA, September 17-18.
Young, Y. and Kinnas, S. 2003d Numerical
modeling of supercavitating propeller flows. Journal of Ship Research, 47, 1, March, pp. 4862.
Young, Y. and Kinnas, S. 2003e Performance
prediction of surface-piercing propellers. To be appeared in Journal of Ship Research.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Thurst or torque coefficient

ventilated cavity

fully
ventilated

transition
partially
ventilated

partially
cavitating

ventilated cavity

vapor/gas cavity

m+m h

vapor cavity
ventilated cavity

k, d+d h

k, d
Advance coefficient, J=V/nD

Figure 1: Schematic diagram of the four major


flow regimes of surface-piercing propellers.

in air

in water

Figure 4: Single degree of freedom approximation


of the blade system in air and in water.

Figure 2: Observed ringing pattern of the ventilated cavity for surface-piercing propeller M841B
at J = 0.8 and Fr = 6.0. Taken from (Olofsson
1996).

Figure 5: Photograph of propeller model 841-B.


Taken from (Olofsson 1996).

blades

flat plate

0.2

free surface
X, x

0.4

y/R

(x,y,z)

Z
z

velocity distribution in propeller plane

0.6

free
surface
non-axisymmetric
effective inflow
wake

0.8

ventilated cavities
1

vx:

0.25

0.5

Figure 3: A partially ventilated surface-piercing


propeller subjected to a general inflow wake. The
blade-fixed (x, y, z) and ship-fixed (X, Y, Z) coordinate systems are shown. Shown in (Young &
Kinnas 2003c).

0.75

Vx

0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 0.55 0.6 0.65 0.7 0.75 0.8 0.85 0.9 0.95 1

Figure 6: Axial velocity distribution at the propeller plane. Based on data from (Olofsson 1996).

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

0.2

0.8

0.6

KT, 10KQ

0.15

0.1

0.4

10KQ

0.05

0.2

KT

Figure 7: Flexure-unit used to measure the reaction load. Taken from (Olofsson 1996).

0.6

0.8

0
1.4

1.2

JA
Figure 9: Averaged performance characteristics
of propeller model 841-B in ship-fixed coordinates. Shown in (Young & Kinnas 2003c).
Axial force coef. for J=1, =1
0.045
BEM
FnD=2
FnD=4
FnD=6
FnD=7.7

0.04
0.035
0.03

KFx

0.025
0.02
0.015
0.01
0.005
0
0.005
0

45

90

135
180
225
blade angle, = t

270

315

360

Figure 10: Comparison of predicted (BEM) and


measured (Exp) axial blade force coefficients for
J = 1.0. Also shown is the effect of Froude number on the measured values. Propeller M841B.
Axial force coef. for J=1.2, =1
0.05
BEM
FnD=2
FnD=4
FnD=6
FnD=7.7

0.04
0.03

Figure 8: Measured natural frequencies of blade


flexure versus angular position obtained by exciting the blade-flexure with the blade being gradually immersed in water step-by-step. Taken from
(Olofsson 1996).

KFx

0.02
0.01
0
0.01
0.02
0.03
0

45

90

135
180
225
blade angle, = t

270

315

360

Figure 11: Comparison of predicted (BEM) and


measured (Exp) axial blade force coefficients for
J = 1.2. Also shown is the effect of Froude number on the measured values. Propeller M841B.

Copyright National Academy of Sciences. All rights reserved.

0.2

FnD=6, =1
wet/n=10.153

0.1
0
0
0.2

10

15

20
kn

25

40

wet/n=10.153

10

15

20
kn

25

30

35

40

FnD=6, =9
wet/n=10.153

0.1
0
0

35
FnD=6, =3

0.1
0
0
0.2

30

10

15

20
kn

25

30

35

Figure 12: Measured harmonic coefficients of propeller M841B in blade-fixed coordinate system at
J = 0.8, Fr = 6. wet /n is the ratio of the fundamental frequency of the blade in water to the
angular frequency = 2n of the propeller.

40

Figure 14: Assumed and measured natural frequencies of the blade versus angular position.
predicted motion for J=0.8

5
dis, u/D

Cfx(kn)/Cfx(0)

Cfx(kn)/Cfx(0)

Cfx(kn)/Cfx(0)

Twenty-Fifth Symposium on Naval Hydrodynamics

x 10

5
10
15
0
0.02

vel, v/nD

FnD=2
FnD=4
FnD=6
FnD=7.7

45

90

135

180

225

270

315

360

45

90

135

180

225

270

315

360

45

90

135

180
225
blade angle

270

315

360

0
0.02

acc, a/n2D

0.04
0
2
0
2
0

Figure 15: Predicted blade tip motion for propeller M841B at J = 0.8.
Axial force coef. for J=0.8 FnD=4
F(t)
f(t)
F(t)+f(t)
Exp =1
Exp =3
Exp =9

0.04

KFx

0.03

0.02

0.01

0.01
0

45

90

135
180
225
blade angle, = t

270

315

Figure 16: Comparison of predicted and measured total hydrodynamic load for propeller
M841B at J = 0.8, Fr = 4.
Figure 13: Comparison of the observed(left) and
predicted(right) ventilation patterns. Propeller
M841B. JA = 1.2. = 9. Fr = 6. Shown in
(Young & Kinnas 2003c).
Copyright National Academy of Sciences. All rights reserved.

360

Twenty-Fifth Symposium on Naval Hydrodynamics

Axial force coef. for J=0.8 FnD=6

Axial force coef. for J=1 FnD=7.7

0.06
F(t)
f(t)
F(t)+f(t)
Exp =1
Exp =3
Exp =9

0.05

0.04

0.03

0.03

KFx

KFx

0.04

F(t)
f(t)
F(t)+f(t)
Exp =1
Exp =9

0.02

0.02
0.01
0.01
0

0
0.01
0

45

90

135
180
225
blade angle, = t

270

315

0.01
0

360

Figure 17: Comparison of predicted and measured total hydrodynamic load for propeller
M841B at J = 0.8, Fr = 6.

45

Axial force coef. for J=1 F =4

315

360

nD

0.03
F(t)
f(t)
F(t)+f(t)
Exp =1
Exp =3
Exp =9

0.035
0.03
0.025

0.02
0.015

0.02
0.015

0.01
0.005

0.01

0.005

0.005

0.01
45

90

135
180
225
blade angle, = t

270

315

F(t)
f(t)
F(t)+f(t)
Exp =1
Exp =3
Exp =9

0.025

KFx

KFx

270

Axial force coef. for J=1.2 F =6

nD

0.015
0

360

Figure 18: Comparison of predicted and measured total hydrodynamic load for propeller
M841B at J = 1.0, Fr = 4.

45

90

135
180
225
blade angle, = t

270

315

360

Figure 21: Comparison of predicted and measured total hydrodynamic load for propeller
M841B at J = 1.2, Fr = 6.
Axial force coef. for J=1.2 F =7.7

Axial force coef. for J=1 F =6

nD

nD

0.045

0.03
F(t)
f(t)
F(t)+f(t)
Exp =1
Exp =3
Exp =9

0.04
0.035
0.03

F(t)
f(t)
F(t)+f(t)
Exp =1
Exp =9

0.025
0.02
0.015
KFx

0.025
KFx

135
180
225
blade angle, = t

Figure 20: Comparison of predicted and measured total hydrodynamic load for propeller
M841B at J = 1.0, Fr = 7.7.

0.04

0.005
0

90

0.02
0.015

0.01
0.005

0.01
0
0.005
0.005

0
0.005
0

45

90

135
180
225
blade angle, = t

270

315

Figure 19: Comparison of predicted and measured total hydrodynamic load for propeller
M841B at J = 1.0, Fr = 6.

360

0.01
0

45

90

135
180
225
blade angle, = t

270

315

Figure 22: Comparison of predicted and measured total hydrodynamic load for propeller
M841B at J = 1.2, Fr = 7.7.

Copyright National Academy of Sciences. All rights reserved.

360

Twenty-Fifth Symposium on Naval Hydrodynamics

Axial force coef. for J=0.8, =1

J=0.8, F =6
nD

0.05
IDEAL
BEM
F =2
nD
F =4
nD
F =6

0.04

f =1138, f =400
dry
wet
f =1138, f =477
dry
wet
f =1138, f =550

0.04

dry

Exp =1

wet

nD

KFx

0.03

0.02

0.01

0.01

0.01
0

45

90

135
180
225
blade angle, = t

270

315

360

Figure 23: Plot of the ideal and computed


rigid-blade hydrodynamic load, along with the
measured total hydrodynamic forces at different
Froude numbers. Propeller M841B at J = 0.8.
F(t)
f(t)
F(t)+f(t)
Exp =1

Fx

0.04
J=0.8 F

0.02

=4

nD

0
0

45

90

135
180
225
blade angle, = t

270

Fx

J=0.8 FnD=6

0.02

315

360

F(t)
f(t)
F(t)+f(t)
Exp =1

0.04

0.02

0.01
0

45

90

135
180
225
blade angle, = t

270

315

ABAQUS, C3D20R
20x10
20x20
20x40
30x40
40x10
Experiment

2000

1000

2.5

dry

Exp =1

360

Figure 27: Comparison of the predicted and measured natural frequencies of propeller M841B in
air.

=0.03, =0.05
dry
wet
=0.03, =0.1
dry
wet
=0.06, =0.1

0.04

315

mode number

J=0.8, FnD=6
0.05

270

3000

360

Figure 24: Comparison of predicted and measured total hydrodynamic load for propeller
M841B at J = 0.8 using the IDEAL rigid-blade
hydrodynamic load.

135
180
225
blade angle, = t

4000

90

5000

measured
predicted

2nd mode

wet

f/f_1(dry)

0.03
KFx

45

Figure 26: Dependence of the predicted axial


force coefficients on dry and wet . Propeller
M841B. J = 0.8. Fr = 6.

blade frequency in air(Hz)

KFx

0.03

0.02

1.5

1st mode
1

0.01
0.5

0
0

0.01
0

45

90

135
180
225
blade angle, = t

270

315

360

Figure 25: Dependence of the predicted axial


force coefficients on dry and wet . Propeller
M841B. J = 0.8. Fr = 6.

90

180

270

360

blade angle (degrees)

Figure 28: Comparison of the predicted and measured natural frequencies of propeller M841B in
water.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics


St. Johns, Newfoundland and Labrador, CANADA
8-13 August 2004

Effect of Operational Conditions on the Cavitation Inception


Speed of Naval Propellers
T.J.C. van Terwisga1,2, D.J. Noble3, R. vant Veer1, F. Assenberg1, B. McNeice4,
P.F. van Terwisga5
(1 MARIN, The Netherlands, 2Delft University of Technology, 3Defence R&D
Canada - Atlantic, 4Royal Australian Navy, 5Royal Netherlands Navy)

ABSTRACT

INTRODUCTION

A propeller analysis package for time domain


simulations is described that permits evaluation of the
effects of variable operating conditions on propeller
cavitation inception in a seaway. The package is based
on linear seakeeping and propeller lifting surface
theories and is combined with a simulator that models
the dynamic response of conventional twin propeller
and ship engine systems. It is developed in a
cooperative project by the Canadian, Netherlands and
Australian Navies and MARIN. In this paper, the
simulation tool is applied to a representative naval
ship and propeller design to illustrate the effect of sea
state, heading and propeller control strategy on the
cavitation inception speed.
In the sample case examined, the effect of sea
state on cavitation inception speed in bow quartering
seas is shown to agree with previous full-scale
observations and measurements, predicting a
cavitation inception speed in sea state 6 of only some
36% of that in sea state 2. A range of propeller control
strategies is examined to show that operating
propellers using a constant thrust coefficient and ship
speed control strategy can produce a cavitation
inception speed that is some 29% higher in following
seas at sea state 4 than obtained with a commonly used
propeller shaft speed control algorithm. For bow
quartering waves in the same sea state, no significant
difference in inception performance could be found
using the same controller settings. Optimization of the
controller for each direction to waves is likely to
improve inception speeds for headings other than
following seas as well.

To minimize the ship's underwater acoustic signature,


propeller design for naval ships is typically governed
by an optimization process aimed at achieving the
highest possible ship speeds for inception of
cavitation. In present day practice, this design
optimization is performed under assumptions of ideal
operating conditions with the ship moving at constant
forward speed on a straight course in calm seas. Staff
requirements, the propeller design condition, and
design evaluation criteria are all aimed at an optimized
design for these conditions. Even the acoustic
evaluation of propellers in full-scale sea trials is
typically performed in deep, sheltered waters, with
results providing no indication of the degradation in
acoustic performance obtained in actual service
conditions.
Observations in previous sea trials have
typically revealed significant reductions in cavitation
inception speeds relative to design expectations during
normal operations in a seaway or during maneuvers. In
actual service, a ship will typically not sail in calm
water on a perfectly straight course due the presence
of oblique waves in a seaway and currents. The ship
will maneuver to sail a specific course or to fulfill
specific tasks, and will also accelerate or decelerate at
certain times. The ship will foul, generating increases
in resistance, and will also experience variable added
resistance due to wind and waves that change the
loading condition of the propellers.
Current propeller design practices and
operational procedures are often found to be severely
lacking due to neglect of the effects of realistic
operating conditions. These effects can now be made
amenable to analytical evaluation and optimization
through the use of currently available ship-propellerengine simulation tools. To assess the effects of in-

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

service operating conditions on the cavitation


performance of naval ship propellers, the Canadian,
Netherlands and Australian Navies have jointly
participated with MARIN in a Cooperative Research
Navies (CRN) Project that has the following
objectives:
to determine the most important factors
contributing to the degradation in propeller
cavitation
performance
during
normal
operations such as maneuvering and sailing in a
seaway (problem definition).
to evaluate available methods for predicting the
most important effects and assemble an analysis
package that can be used to predict cavitation
performance under these conditions (software
development).
to formulate strategies for operational guidance on
existing ships and suggest improvements for the
design of new propellers and ship control systems
in order to control and maximize the operational
window for cavitation free operation under
service conditions (operational and design
strategies).

including the computation and validation of some key


results obtained for the unsteady wake field in waves.
An example of the use of the simulation tool will then
be provided to illustrate the effect of sea state, heading
and the Propeller Control Algorithm (PCA) on
cavitation inception speed. The selected ship and
propeller design for this example are considered
representative of contemporary naval ships.
Although the ship and propeller particulars
are not provided, the results of numerical simulations
in a seaway are presented and their implication for the
control and optimization of propeller cavitation
performance under in-service conditions is discussed.
IMPORTANCE OF OPERATIONAL
CONDITIONS ON CAVITATION INCEPTION
The importance of operational conditions on the
cavitation inception behaviour of a propeller is
illustrated here with full scale observations made on
board of the Hydrographic Survey Vessel, HM
Tydeman (see Figure 1).

This paper provides an update to the progress


of work in this CRN Project reported previously in a
paper presented at the 24th Symposium on Naval
Hydrodynamics in Fukuoka, Japan [Kuiper et al.
2002]. That paper investigated the changes in the fullscale propeller inflow of a Fremantle Class Patrol Boat
of the Royal Australian Navy between a straight run
and a turn at constant rudder setting based on LDV
measurements. Results were also provided for the
effect of the measured inflow changes on propulsion
and cavitation predictions with a lifting surface code.
For this particular ship, changes to the transverse
inflow velocity components were found to have a
dominant effect on the power absorption and
cavitation performance of the propellers.
Development of tools for the prediction of
ship propeller and engine performance under realistic
operating conditions will help to meet the growing
demand by navies for integrated system evaluation by
means of modeling and simulation. These tools can
provide important input for evaluating the
effectiveness of tactics and decoys in torpedo defense
and anti-aircraft simulations. The simulation tools can
be applied to evaluate the effect of optimal control of
propeller cavitation and engine loads on resulting
acoustic and IR signatures.
In the following sections of this paper, results
from full-scale observations on a geographic survey
vessel are provided to demonstrate the significance of
operational conditions on propeller cavitation
performance. The philosophy and structure of the
PEASE simulation tool are then described in detail,

Figure 1 HM Tydeman
This is a survey vessel of the Royal
Netherlands Navy, with which cavitation inception
tests have been conducted in both sheltered waters and
open sea.
Calm water inception tests were conducted in
the Sognefjord (Norway) and inception tests in waves
were conducted in open sea with significant wave
heights of approximately 3.5 m and an observed
period of 6 s. Wave height and direction were
estimated by the crew and later confirmed by
hindcasting. The tests in calm water were conducted in
ideal conditions: No waves, straight course, minimal
rudder motions, no wind and no traffic.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Cavitation inception was determined from


visual observations with the MARIN CAVOBS
system. This comprises a digital camera that is
triggered with the shaft frequency, thus collecting
pictures of a particular blades passage. A description
of the system is given by Verkuyl et al. (2000).
Inception was determined by establishing
when the worst performing blade showed cavitation
over some 30% of the time. Cavitation inception was
determined for both sheet cavitation and tip vortex
cavitation. The figures quoted in this paper refer to tip
vortex cavitation inception, as it was the first type of
cavitation to occur.
After cavitation inception had been
determined for calm water conditions, straight course
ahead, similar tests were conducted for turning circles,
applying rudder angles of 10 and 20 deg to both port
and starboard. Figure 2 presents the results of the
cavitation inception tests for calm water conditions.
This figure shows that a reduction in cavitation
inception speed of some 50% occurred for rudder
angles of 20 deg.

water condition. For head seas, the cavitation


inception speed even reduces to zero.
These trials results illustrate clearly the
importance of sea state and heading on cavitation
inception speed. They also underline the relevance of
the current study, which ultimately aims at the design
and operation of the ship and propulsion system in
such a way that significant reductions in radiated noise
are attained.
9
8

Cav. Inception Speed CIS [kts]

7
6

Following seas
Calm water
Stern Qrtrng
seas

5
4

Beam seas

3
2

Bow qrtrng
seas

Head seas
0
8

10 dg PS

6
Cav. Inception Speed CIS [kts]

Figure 3 Effect of sea state 5 on inception speed

0 dg

OBJECTIVES AND SCOPE


10 dg SB

Having illustrated the importance of sea state and


heading on cavitation inception, this paper aims to
demonstrate the adverse effects of ship operations on a
straight course1 in a seaway on propeller cavitation
performance through the use of a simulation package
named PEASE (Propeller Evaluation and Analysis in a
Service Environment). This package consists of linear
ship seakeeping and propeller lifting surface
components in combination with a simulator that
solves the equations of motion, and models the
dynamics and control of a complete twin-screw,
propeller-engine system in the time domain. The paper
also describes results for a representative naval ship
design that indicate how the degradation in propeller
cavitation performance is affected by sea state,
heading and propeller control strategy.
In simulating propeller performance on a
straight course in a seaway with an emphasis on
cavitation inception, it is necessary to obtain a
sufficiently adequate prediction of the instantaneous
operating point of the propeller. This implies that
accurate predictions are needed of the parameters that
determine cavitation inception for a given propeller:

20 dg SB

20 dg PS
3

Figure 2 Effect of a turn on inception


Cavitation inception measurements in waves
are rare. During the Tydeman trials, weather
conditions worsened and the position of the ship in
open sea, provided a good environment to explore the
effect of sea state and heading on cavitation inception.
In an average sea state of 4 to 5, five series of
inception measurements have been recorded for the
following headings; head waves, bow quartering
waves, beam seas, stern quartering waves and
following waves. Figure 3 presents the results. It is
shown that in following seas, the inception speed
increases by some 17% above the calm sea value. In
the other headings, cavitation inception occurs at
significantly lower speeds when compared to the calm

A prediction model for the effects of maneuvering on


cavitation inception is currently under development.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

First, the seakeeping characteristics of the


ship at an average ship speed in a given sea state and
heading are computed in the frequency domain. This is
done with the linear seakeeping program PRECAL
(Van't Veer, 2003). This program is used to provide
the three components of the seakeeping-induced
velocities in the propeller plane. In a subsequent step,
the frequency domain wake velocities from PRECAL
are converted into time domain velocities by a
program named WAVECAL, that uses a random
phase angle distribution to realize a wave train. The
resulting wake velocities are then added to the
nominal wake field obtained from steady state model
testing in calm water.

The time dependent inflow velocity field in


which the propeller operates.
The time dependent ambient pressure.
The time dependent propeller rotation rate and
ship speed.

The unsteady wake field entering the


propellers is built up from the calm water nominal
wake field and the unsteady wake field caused by the
seaway and the hulls response. In the current PEASE
model, ship motions are first computed in the
frequency domain using linear seakeeping theory for a
given steady ship speed. The ship motion velocities
and disturbance potentials that contribute to the
resulting inflows to each propeller are then added to
the calm-sea nominal wake field to provide a series of
inflow velocity distributions at discrete time steps
during a simulation run. In the current model, the ship
motions in a seaway are decoupled from the propellership dynamics.
The computation of the ambient pressure in a
seaway is less straight-forward than it is for calm
water conditions. The time derivative of the potential
function in the Bernoulli equation, normally neglected
in steady state calm water considerations, can no
longer be neglected.
To determine the instantaneous ship speed
and propeller rotation rate, the equations of motion of
the ship in the longitudinal direction and the rotor
dynamics of the propeller-engine system are solved at
each time step. The engine dynamics are modelled by
a generic PID controller, hereafter referred to as the
Propeller Control Algorithm or PCA. This algorithm
can be programmed to enforce various propeller
control strategies, such as operation at constant rpm,
constant torque, constant power, etc., and thus
determine the instantaneous operating point of each
propeller.
Having defined the propeller operating
conditions, a detailed propeller analysis can be made
for each time step using lifting surface theory to derive
surface pressure distributions on propeller blades and
to determine whether and where cavitation occurs.
Cavitation performance in operational conditions can
then be evaluated using the percentage of time during
which the propeller is observed to be free of
cavitation.

Figure 4 Data flow chart of the PEASE suite of


programs

STRUCTURE OF PEASE SIMULATION TOOL

In the next step, the ship speed and the


propeller rotation rate for both propellers (in a twin
screw arrangement) are computed by the time domain
simulator (PEASE simulator). To this end, the
equations of motion for the longitudinal ship dynamics
and the propulsion system rotor dynamics are solved
simultaneously for both propellers. A time-averaged

Having described the philosophy behind the timedomain simulation package PEASE, this section will
focus on the structure and the elements used in the
package. A data flow chart of the PEASE simulator is
presented in Figure 4.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

total resistance as a function of speed is available,


together with the propellers open water torque
characteristic. The Propeller Control Algorithm (PCA)
determines the reaction from the combined engine
and engine control system to changes in engine
operating point. A simple PID controller is used for
this system, requiring three control coefficients Kp, Ki
and Kd that are defined for proportional, integration
and differentiation components, respectively. The
result of this block of the simulator is the
instantaneous ship speed and propeller rotation rate.
The above elements feed subsequently into
the lifting surface program ANPRO (Van Gent, 1975),
which calculates the detailed surface pressures over
the propeller blade. Also, the instantaneous thrust
calculated by ANPRO is fed back to the simulator for
the computation of the longitudinal ship dynamics in
the next time step.
With the detailed pressure distribution now
available, it is assessed whether and where cavitation
occurs on the blade. To this end, a simple cavitation
criterion is applied, where cavitation is said to occur
whenever the local pressure is lower than the vapor
pressure of seawater.
A complication in the determination of
cavitation inception arises for contemporary naval
propellers due to the fact that tip vortex cavitation
usually occurs first with increasing speed. This type of
cavitation is not modelled in current lifting surface
theory. To account for this deficiency, a correction is
made by adjusting the ambient pressure until the
computed sheet cavitation results are obtained at the
ship operating point for tip vortex cavitation inception
observed at full scale (in ideal, calm water conditions).
The adequacy of this correction is expected to worsen
as operating conditions deviate further from ideal
conditions, due to differences in the variation of
inception characteristics of tip vortex cavitation with
the propeller operating condition. However, for a first
assessment of cavitation inception in off-design
conditions, the current correction is considered to
provide a satisfactory result.
The input data sets needed for this analysis
are thus the geometrical data for both the ships hull
and the propeller, operational data (sea state, heading
and average ship speed), nominal wake field data in
steady-state, calm water conditions and data defining
the Propeller Control Algorithm.

velocity components due to waves and ship motions


can be linearly superimposed onto the nominal wake
field.
Both the steady and unsteady wake field will
be divided into components that can be assessed
independently. The total effective wake field in the
propeller plane that can be used as input for the
propeller analysis, is based on the following separation
of flow components:
eff
eff
V{total
} = V{calm water } + V{ waves }

= Vcwn + Vcwint + Vwii + Vwid + Vwir + Vsmt

(1)

The subscript cw in equation (1) indicates


calm water components, the subscript wi indicates
wave induced components and the subscript sm
indicates ship motion induced velocities.
The total calm water flow field is defined as a
summation of three components: the nominal flow
n
i
field Vcw , the propeller induced flow, Vcw and the
int

propeller flowfield related to interaction effects, Vcw .


The preferred wake field to be used for a propeller
analysis is the so called effective wake field, which is
the total wake field minus the propeller induced
velocities:

Vcweff = Vcwtot Vcwi = Vcwn + Vcwint

(2)

Typically, the total and induced flow field are


not available, and the effective wake field is built up
from the nominal wake field and the propeller-wake
interaction component, for which a code is available
(Van Gent and Hoekstra, 1985).
n
The nominal flow field, Vcw , is the flow field
caused by the ships hull and appendages without
propellers fitted. This wake field is considered to be
time independent. Possible unsteady flow phenomena
such as the effect of vortices shed by the hull are time
i
averaged. The propeller induced flow field, Vcw , is
solely due to the propeller action in a given wake field.
This part is caused by the loading or force distribution
on the blades and provides a propeller-induced
velocity that is dependent on the blades angular
position. The resulting contribution to the wake field is
therefore periodic at blade rate. The propeller-wake
int
interaction component, Vcw , is the wake field
introduced to account for interaction effects that are
lost by the breakdown into separate small components.
Phenomena that are included here are the effect of
propeller suction on the incoming hull flow and on the
vorticity distribution.

Unsteady Wake Field


Following the assumptions of linear theory, the total
wake field is assumed to be a summation of the steady
(calm water) wake field and the unsteady seakeeping
components (the contributions from the waves and the
ship motions). Aalbers and Van Gent (1984) validated
this approach with measurements, showing that the

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

The total unsteady flow field can be


decomposed into the incoming undisturbed waves
together with the diffracted and the radiated wave
velocities and into the flow velocities induced by the
ship motions in 6 degrees of freedom.
The additional wake velocity field induced by
i
the undisturbed incoming wave is denoted Vwi . The
undisturbed incoming waves will partly be reflected
by the hull, and thereby generate a diffracted wave
d
system, leading to a diffracted flow field Vwi . As the
ship sails through the seaway, it carries out
translations and rotations, giving rise to a wave system
that is radiated by the hull. The corresponding radiated
r
velocity field is denoted Vwi .
The above seakeeping components are
described in a frame of reference steadily moving with
the average ship speed. The motions of the propeller,
due to the global 6 degrees of freedom ship motions,
relative to this reference frame introduce additional
velocities. If the velocity of the propeller plane due to
ship motions is described by Vsm , the corresponding
flow field experienced by the propeller is given by:

Vsmt = Vsm

due

to

the

inflow

Finally, the transient pressure term

, due
t

to the potential flow field.


On the free-surface wave elevation a dynamic
and kinematic boundary condition should be satisfied,
prescribing the pressure on the water surface equal to
the atmospheric pressure and the flow velocity
direction of the water particles. Using the Bernoulli
equation, the surface elevation is found to be:

1 1

( Vs2 )
g t 2 g

(5)

In the present version of the PEASE code, the


second term on the right hand side of eq. (5) is
neglected, resulting in the following expression for the
unsteady potential contribution in the pressure:

(3)

= g
t

(6)

In a linearized model, it is justified to neglect


the . term.

Ambient Pressure
The calculation of the pressures in the propeller plane
has a direct effect on the calculation of the cavitation
inception number. Since the cavitation number
incorporates the ambient pressure, it is necessary to
determine the pressures in the propeller plane
accurately. The pressure in an unsteady potential field
defined by the total velocity potential , such as in
the case of a propeller operating in a seaway, can be
determined from the Bernoulli equation:

1
( Vs2 ) gz + pa
t 2

12 ( Vs2 ) ,

velocities in the propeller plane.

The resulting total effective wake field in the


propeller plane that is input for the propeller analysis
is now obtained as the sum of the individual
components as expressed in equation (1).

p =

position, taking the effect of the ship motions on


the position in the propeller plane under the calm
water plane into account.
Secondly,
the
dynamic
term

The neglect of the Vs2 term implies that there


is no pressure recovery in the wake of the hull. In
reality, pressure recovery will be somewhere in
between zero (strong viscous dissipation of energy)
and the full dynamic pressure (for potential flow). The
ratio of the pressure recovery can be determined for
example from a RANS analysis of the flow about the
hull.
The above treatment simplifies the
assessment of the unsteady pressure, and although
incorrect, it is typically used in linear theory without
introducing serious inaccuracies. This approach
however, will require further validation.

(4)

where is the total disturbance potential,


incorporating both steady and unsteady terms.
It can be seen from equation (4) that the
pressure is built up from three contributions:
First of all, the static pressure ps , given by

Propeller Rotation Rate and Ship Speed


The propeller rotation rate and the ship speed are
obtained by solving the equations of motion of the
longitudinal translational motion of the ship system
and the rotating motion of the propeller-engine
system. The equation of motion of the propeller-ship
system is given by:

gz + pa , where pa is the atmospheric


pressure. In this case, the earth fixed reference
system is used to determine the (vertical) z-

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

(M + Ma )

dVS
= Ttot (1 t ) R
dt

For a twin screw configuration, we now have


three equations with three unknowns, viz.: 1, 2 and
VS. The other parameters have to be given as input to
the simulator.
The determination of the time dependent
values of propeller torque Qj, thrust Tj and engine
torque MDEj is now described. The propeller thrust Tj
is determined from the propeller analysis by the lifting
surface code ANPRO. This analysis is also made for
the detection of cavitation inception. Contrary to the
thrust, the instantaneous propeller torque Q is
determined from the measured open water
characteristics and the instantaneous nominal wake
fraction. The measured open water characteristics for
torque are preferred because of the greater uncertainty
in the computed torque compared to the thrust
calculated by ANPRO.
The instantaneous propeller torque is
consequently determined from the open water torque
coefficient:

(7)

where:
M = displacement mass of the ship system
Ma = added mass of the ship in longitudinal
direction, that is typically frequency
dependent. This mass is given a constant
value
for
the
frequencies
under
consideration
Ttot = total thrust from all propellers
VS = ship speed
R
= total instantaneous hull resistance. This
resistance is given a time averaged value,
built up from a calm water contribution and a
time averaged drag in waves. A more realistic
model predicting a time varying drag in
waves is currently being implemented.
t
= thrust deduction value (time averaged)
It is realized that the resistance fluctuations in
waves might have an important effect on instantaneous
propeller loading and therefore on cavitation
inception. Work is currently defined to include an
algorithm that includes this time dependency of the
added resistance in waves in a future version of
PEASE.
The equation for the rotating motion of the
propeller-engine system is given by:

IP

d j
dt

where:
IP
=

= M DEj Q j

M DEj =

torque delivered by the prime mover

Qj

1
4

2j D5

where KQj = f VS ,(1 w jn ),

(9)

j
and
2

w jn is the

nominal instantaneous wake fraction for shaft system


j. The engine torque is determined by the PCA as
described below.
Apart from the PCA, the above set of
differential equations needs to define limits and initial
values to find a unique solution. To this end, the time
averaged estimates of ship speed and propeller rotation
rate VS0 and n0 provide the initial values. An important
constraint for the engine is its operational envelope,
which is often expressed in terms of torque and engine
rotation rate (see Figure 10).

(8)

polar moment of inertia of the complete


rotary system. This includes propeller,
shaft, gearbox and rotating part of the
engine. When a Power Take Off is used,
the moment of inertia of this generator
should be included as well.
shaft rotation rate

Q j = K Qj

Propeller Control Algorithm


A generic block diagram, representing the shippropeller-engine system is depicted in Figure 5
(Stapersma, 2000). This diagram shows the relation
between the translational motion dynamics of the shippropeller system and the rotational motion dynamics
of the propeller-engine system.
In the upper left part of the Figure 5, the
Propulsion
Control
Algorithm
is
depicted,
representing the Command, the Engine Control
System and the Engine. Pitch control is left out of
consideration here. The propeller operates with a fixed
pitch setting. Four different control strategies are
implemented in this PCA, each controlling the diesel
engine torque delivered to the propeller:

(often a diesel engine). The magnitude of


this torque is currently controlled by the
Propeller Control Algorithm. Four
different PCAs are used in this study,
which are elaborated in the following.
= torque exerted by the propeller on the

shaft system at the thrust bearing. It


includes bearing losses.
Subscript j indicates the shaft system location (e.g.
starboard or port).

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Command

Engine
control
system

Propulsion
control
system

Engine/
controler

Pitch
control
system

Propeller
Control
Algorithm

Mprop

Mshaft

Propeller
Torque

Propeller
Thrust

n
n

1
2 Ip

Mdt

controlling the engine on CT alone is not sufficient to


maintain a constant ship speed. This is caused by the
indifference of the thrust coefficient to variations in
ship speed. The error functions used in PCA4 are thus
defined as:
Fprop

Fship

Ship
resistance

n
Vs

Vs

eCT ji = CT SP CT j ( i 1)

Vs

1
m

eVS i = VS SP VS ( i 1)

Fdt

M DE ji = K p CT eCT ji + K pVS eVS i +

Vs
n

J =

Va

1 w

Va

nD

K i CT eCT ji dt + K d CT

Sea keeping
disturbances

PCA1
PCA2
PCA3
PCA4

=
=
=
=

dt

VALIDATION OF UNSTEADY WAKE FIELD


VELOCITIES

(10)
The validation of a simulation tool that is itself
composed of several complex components can be
cumbersome. However, the validation is still an
essential process that is required to define the
limitations and applicability of the tool. A proper
validation requires reliable data sets that preferably
allow for an end-to-end uncertainty analysis. Such an
analysis would assist in directly quantifying the
uncertainty of the final results. Most often, however,
such data sets are not available and an overall estimate
of the uncertainty must be obtained from only partial
validation studies. Validation obviously is partly
driven by opportunity, with data sets typically only
being available for portions of the simulation code.
One of the effects expected to play an
important role in cavitation inception in a seaway is
the unsteady ship wake field induced by the waves.
The wake field velocity variations affect the propeller
operating condition in two ways: on the one hand, they
affect the propeller pressure distribution and resulting
blade loading variations directly, while on the other
hand, through variations in thrust and thus ship speed,
they again affect the loading variations indirectly.
Partial validation of these unsteady velocities
can be conducted with the availability of a unique data

where:
ej = error between some user defined set point SP
and some measured process variable for shaft
system j
Kp = controller proportional coefficient
Ki = controller integration coefficient
Kd = controller differentation coefficient
Subscript j refers to the shaft system and,
except for Ki , subscript i to the time step. The error
function for the constant rotation rate is defined by:
e ji = SP j ( i 1) where:

e ji

dt

error function. The integral and differential parts of the


error function for the speed have been omitted.
In representing the engine command and the
engine system as one element in the block diagram, it
is implicitly assumed that the engine reacts
instantaneously to the new fuel rack setting that is
determined by the PID control system. However, this
would likely be a valid assumption since the
combustion process has a much shorter reaction time
than the shaft system.

A generic PID controller is used for all four


algorithms. The torque at time step i for each engine is
consequently determined from a correction to the
torque obtained in the previous time step:

de j

(11)

eVS i is only reflected in the proportional part of the

constant propeller rotation rate


constant power
constant torque
constant thrust coefficient CT and
ship speed

M DE ji = K p e ji + K i e ji dt + K d

ji

It is noted here that the error in ship speed

Figure 5 Generic block diagram of the ship-propellerengine system

deCT

= error of shaft system j at time step i

SP = set point of propeller rotation rate


j ( i 1) = actual propeller rotation rate of shaft
system j at time step i-1
The constant thrust coefficient strategy
requires multiple error functions to be used because

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

set of wave-induced wake field velocities. Aalbers and


Van Gent (1984) performed unsteady velocity
measurements with Laser Doppler Velocimetry (LDV)
for 8 positions in the propeller disk (Figure 6) in
regular head waves of 1m wave amplitude. They
found wave induced wake field variations up to 20%
of the maximum wake deficit (wmax@r=0.88) per unit
wave amplitude. This implies that unsteady wake field
variations up to some 23% of the maximum wake
fraction wmax can occur in sea state 4, with a 14%
chance of exceeding the maximum. These wake
velocity variations can significantly affect cavitation
inception.

be sufficiently well predicted to provide a realistic


simulation of the effect of a seaway on cavitation
inception.
AXIAL Velocity, pnt 4 (240 deg)

AXIAL Velocity, pnt 1 (180 deg)


0.3

0.3

PRECAL

PRECAL

uxa / (wmax A)

uxa / (wmax A)

0.2

0.15

0.1

0.2

0.15

0.1

0.05

0.05

0
0

0.5

1.5

AXIAL Velocity PHASE, pnt 1 (180 deg)

1 .5

AXIAL Velocity PHASE, pnt 4 (240 deg)

180

180

PRECAL

PRECAL

Experiments

Experiments

90

90

0
0

0.5

-90

1.5

PHASE [deg]

PHASE [deg]

0.5

WAVE FREQUENCY [rad/s]

WAVE FREQUENCY [rad/s]

-180

Experiments

0.25

Experiments

0.25

0
0

0.5

-90

WAVE FREQUENCY [rad/s]

-180

WAVE FREQUENCY [rad/s]

Figure 7 Comparison of measured and computed


axial velocities in position 1 and 4 of the propeller
plane; Comparison of amplitudes (upper half) and
phase angles (lower half).

Figure 6 Positions in propeller plane for which


velocity measurements were conducted

The prediction for the top location is good,


while the velocities in point 4 are over-predicted by
PRECAL. This can be attributed to the influence of
the shaft and brackets at which location (210 deg)
there is a clear decrease of the flow velocities in the
measurement. The influence of these appendages is
not accounted for in PRECAL. The prediction for the
shorter waves is less accurate than for longer waves.
Although in long waves the absolute surge motions
increase, the relative surge motions decrease as the
ship follows the wave contour. The wave velocity due
to the incident wave dominates the flow field. In
shorter waves the diffraction effects become more
important and since these cancel the incident wave the
measurements become very sensitive to local
disturbances.

Figure 7 shows a comparison of measured and


computed velocities at points 1 and 4 in the propeller
disk as a function of the wave frequency. Both the
amplitude of the wake velocity variation and the phase
angle (relative to the wave at the CG) are given.
Error bars about the experimental velocity
amplitudes indicate the estimated uncertainty in the
measurements. The general trend in this comparison
(which holds for most positions in the propeller disk)
is that the computations are within the estimated
experimental uncertainty of some 25% (with 95%
reliability for wave frequencies lower than
approximately 0.8 rad/s). The deviation is larger for
higher frequencies however. The prediction of the
phase angle is satisfactory.
It is noteworthy to mention that the scatter in
results is larger for the top position of the propeller
plane. This effect is attributed to the flow distortion
generated by the appendages and the complicated
viscous dominated wake this gives.
The Response Amplitude Operators (RAOs)
of the calculated ship motions for surge, pitch and roll
are in good agreement with the experimental values.
The RAO values of the local wake velocities seem to

EFFECT OF SEA STATE, HEADING AND


PROPULSION CONTROL ON CAVITATION
INCEPTION A CASE STUDY
The effect of Propeller Control Algorithm (PCA), sea
state and heading on cavitation inception have been
studied with the previously described simulation tool.
Four PCAs have been studied, involving constant

Copyright National Academy of Sciences. All rights reserved.

1 .5

Twenty-Fifth Symposium on Naval Hydrodynamics

propeller rotation rate, constant power, constant torque


and a combined constant thrust coefficient (CT) and
speed. Most of the simulations were conducted for sea
state 4 and a heading angle of 0 degrees (following
waves). The effect of sea state was investigated for a
heading of 135 degrees (bow quartering waves), where
the effect of sea states 2 and 6 were additionally
studied. The effect of heading was studied only for sea
state 4 (at headings 0 and 135 degrees). A review of
the combinations of PCA, sea states and heading
angles considered is given in Table 1.

For each of these parameters, the time


averaged value and the standard deviation were
determined. In addition, the minimum and maximum
value were determined for ship speed, propeller
rotation rate and thrust coefficient CT.
The statistical analysis furthermore produced
the number of cavitation events on the propeller,
where cavitation was considered present whenever the
local pressure coefficient was smaller than the
cavitation number at any position on the blade. A
distinction was made between pressure side and
suction side cavitation. In this way, the percentage of
the time that the blades were free of cavitation on each
side was obtained.
The Cavitation Inception Speed (CIS) in a
seaway can be determined from the statistical
cavitation data when, for a given sea state and heading
combination, a number of different ship speeds have
been simulated. Plotting the percentage of time free of
cavitation versus the average ship speed VS then
shows an almost linear relation, the percentage
cavitation free decreasing with increasing speed.
An example of this process is presented in
Figure 8 for sea state 4, heading 0 and all four PCAs.
The CIS is now defined as the maximum speed where
the propeller is 100% of the time free of cavitation,
given by the intersection between the former relation
and the 100% free of cavitation line.

Table 1 Matrix of simulation conditions


Heading
0
45
90
135
180
Note:

Sea state
4
PCA 1,2,3,4

PCA 1,2

PCA 1,2

PCA 1,2

ship speed VS
propeller rotation rate n
thrust coefficients KT and CT
quasi propeller advance coefficient J, based on
average ship speed
cavitation number n

PCA1 = propeller rotation rate control


PCA2 = power control
PCA3 = torque control
PCA4 = thrust coefficient CT and speed control

The control coefficients used in the PCAs


were chosen by a trial and error process, until stable
convergence was reached for the sea state 4, heading 0
deg condition. These coefficients were subsequently
used for all other conditions, pertaining to that PCA.
For each simulation case, a period of 900 s is
simulated using the same wave realization. The
starting condition, specified by ship speed VS0 and n0,
is estimated based on time averaged values. The ship
velocities are non-dimensionalized by a sea state and
heading dependent reference velocity (Vref(SS,H)).
This reference velocity represents the calculated
propeller inception speed, based on time averaged
values of added resistance in waves and instantaneous
wake velocities. These inception speeds resulted from
initial work, where ship speed and propeller rotation
rate were kept constant. The non-dimensional
inception speed changes reported here thus represent
the separate effect of including the ship and propeller
dynamics only, unless indicated otherwise.
For each simulation case, a statistical analysis
of the simulation has been made. In order to prevent
start-up effects from biasing the statistics, the first
100 s were omitted from the data on which a statistical
analysis was performed. Important statistical data that
were collected are:

120

Perc. time free of cav.

PCA1-rpmcontr
PCA2-Pcontrol

100

PCA3-Torque-control
PCA4-Ct/V control

80

60

40

20

0
0.6

0.7

0.8

0.9

1.1

Vs [-]

Figure 8 Non-dimensional cavitation inception speed


for four Propulsion Control Algorithms at sea state 4,
heading 0

10

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Effect of Propeller Control Algorithm


The present study shows that the Propeller Control
Algorithm (PCA) has an important effect on the
Cavitation Inception Speed (CIS) for sea state 4,
heading 0 (following seas), as presented in Figure 8.
This graph shows that the propeller rotation rate
control (PCA1) produces the worst CIS. The best
control algorithm appears to be the constant CT / ship
speed control (PCA4), resulting in a CIS that is some
29% higher than that of PCA1. Power control (PCA2)
and torque control (PCA3) show a similar intermediate
performance.
Contrary to the important effect that the PCA
has on the inception speed in following seas, its effect
becomes negligible for bow quartering seas as can be
seen in Figure 13. This difference in effectiveness of
the PCA is ascribed to the much higher wave
encounter frequency in bow quartering seas, which is
typically a factor of 5 higher for bow quartering waves
than for following seas (zero up-crossing periods
T0up = 4 s versus T0up = 20 s respectively). The
effectiveness of the control system in coping with such
different time scales depends on its frequency
characteristics and the dynamic characteristics of the
ship-propeller-engine system.
In the present study, the PID coefficients (or
gains) have been chosen to obtain a good behaviour of
the complete system in following seas. These
coefficients have subsequently been used for the
controller in bow quartering seas. It is likely that a
better selection of coefficients is achievable that will
behave better over a wider frequency range.
Considering that the frequency characteristics of the
control system will also necessitate a closer look at the
frequency characteristics of the driving engine and its
control system, this issue will need further attention in
subsequent work.
Figure 9 shows the dynamic cavitation
inception diagrams that result from the propeller
analysis at every time step considered for the rpm
control (PCA1), power control (PCA2) and thrust
coefficient / ship speed control (PCA4).

constant rpm control

sigma_n

cav. free
cav. s.s.
cav p.s.
calm water

Kt [-]

constant Power control

sigma_n

cav. free
cav. s.s.

cav p.s.
calm water

Kt [-]

Constant C T and VS control

sigma_n

cav. free
cav. s.s.
cav p.s.
calm water

Kt [-]

Figure 9 Dynamic cavitation inception Diagrams for


three Propeller Control Algorithms (PCA 1,2 and 4)
in sea state 4, heading 0 and VS=0.93
The calculated calm water inception "bucket"
has been added to these dynamic inception diagrams
for reference purposes. This latter diagram is also
constructed with the propeller analysis code ANPRO,
where the same cavitation inception criterion is used.
The calm water diagram is constructed by analyzing
the propeller at a number of different propeller
rotation rates. For each propeller rotation rate, the
advance velocity is varied and thus the propeller thrust

11

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

coefficient KT. In this way, both pressure side and


suction side cavitation inception could be detected.
Although the average ship speed is the same
for all three simulations, there is a clear difference in
number of cavitation events. It is observed that the
PCA determines the orientation of the operational
region (indicated by a cigar shaped area of operating
points) of the propeller in the cavitation inception
diagram. Clearly, the PCA4 fits best within the calm
water cavitation inception bucket.
Another noticeable observation is the
difference between calm water inception and dynamic
inception in a seaway. Since the same computer code
ANPRO has been used to determine cavitation
inception, the difference must be caused by dynamic
effects in the wakefield and the dynamic pressure.
Clearly, the determination of the operating point using
the thrust coefficient KT and cavitation number is no
longer sufficient to discriminate between cavitating
and cavitation free operations. This discrepancy is
likely to be caused by the spatial and time varying
wake field. Another possible source is a variyng
cavitation number caused by dynamic effects in the
ambient pressure. An understanding of the precise
mechanisms of these effects could possibly be
exploited in the Propeller Control Algorithm to further
delay cavitation inception in dynamic conditions.
It appears from this analysis that the most
successful control system aims at confining the
operational area to a narrow band at essentially a
constant thrust coefficient KT. The problem with such
an algorithm is that there is hardly a change in thrust
coefficient for different ship speeds, so that an
algorithm is needed that combines both the thrust
coefficient and the ship speed in one error function.
PCA4 is a first attempt at defining such an algorithm.
The ultimate control system would limit the
operational area to a single operating point within the
cavitation bucket.
Figure 10 presents the operating regions in
the engine torque-rpm diagram. The red line in these
diagrams represents the static engine limits, produced
by the engine manufacturer. If the PCAs are to be
judged from the point of view of engine management,
PCA4 also seems to offer the best possibility of not
exceeding the engine limits during operation.

M engine [Nm]

constant rpm control


45
40
35
30
25
20
15
10
5
0

p40 - MF ss4 head0 ,


Vs=.92 ,
PIDn=constant
Max eng. torque

500

1000

1500

n [rps]

M engine [Nm]

constant Power control


45
40
35
30
25
20
15
10
5
0

p40 - MF ss4 head0 ,


Vs=.92 ,
PIDP=constant
Max eng. torque

500

1000

1500

n [rps]

M engine [Nm]

constant CT and speed control


45
40
35
30
25
20
15
10
5
0

p40 - MF ss4 head0 ,


Vs=.94 , PID Ct=c,
V=c
Max eng. torque

500

1000

1500

n [rps]

Figure 10 Engine operating regions for three


Propeller Control Algorithms (PCA 1,2 and 4)

12

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Effect of Heading
The effect of heading in a seaway has been studied by
simulating the ship in a sea state 4 in both following
waves (heading 0 deg) and in bow quartering waves
(heading 135 deg). The effect of these simulations on
the cavitation behaviour is plotted in Figure 11.

constant rpm control

sigma_n

cav. free
cav. s.s.
cav p.s.

120

calm water

H0, PCA1-rpmcontrol
H0, PCA2-Pcontrol
H0,PCA4'- Ct control

Perc. time free of cav.

100

H135, PCA1-rpmcontrol
H135, PCA2-Pcontrol

80

Kt [-]
60

constant Power control

40

20

0
0.6

0.7

0.8

0.9

1.1

Vs [-]

cav. free
cav. s.s.
cav p.s.
calm water

sigma_n

Figure 11 Effect of heading on cavitation inception


Speed. All ship speeds are normalized with Vref(4,0)

It is noted that there is no significant


difference in cavitation inception speed for the two
distinct PCAs in bow quartering waves, as opposed to
the results obtained in following waves. Furthermore,
the rate of cavitation events increases faster in
following waves than in bow quartering waves,
suggesting stronger inflow variations on the propeller
in following waves.
Figure 12 shows the cavitation inception
diagrams in bow quartering waves for PCA1 and
PCA2, controlling rpm and power respectively. It is
remarkable in these plots that the power control
algorithm appears unable to orient the operational
region more into the cavitation free bucket. As for
following sea diagrams in Figure 9, a significant
difference is also obtained between calm water
inception (indicated by the light blue line) and
dynamic inception for bow quartering waves.

Kt [-]

Figure 12 Effect of Propeller Control Algorithm on


cavitation inception for bow quartering waves, sea
state 4

Effect of Sea State


Other than the baseline case of sea state 4, two
additional sea states have been considered for bow
quartering waves; Sea states 2 and 6.
Similar to the lack of effect the PCAs have on
CIS in bow quartering waves in sea state 4, Figure 13
shows that there is hardly any difference between the
two PCAs tested for sea states 2 and 6.

13

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

quartering waves in sea state 6 is only some 36% of


the inception speed in sea state 2.
It is further demonstrated that the dynamic
characteristics of the propeller control system,
modeled here as a generic PID controller, have an
important effect on the cavitation inception speed. A
constant thrust coefficient algorithm produced a
cavitation inception speed that is some 29% higher
than a commonly used propeller rpm control algorithm
in following seas at sea state 4. For bow quartering
waves in the same sea state, no significant difference
could be found in inception speed when the controller
settings are unchanged from those used in following
seas.
It is noted that during the current phase of the
project, it became clear that the time-dependency of
the added resistance in waves is significant and is
likely to affect the outcome of the simulations. Work
has been done to derive an algorithm for estimating
the unsteady added resistance in waves, reported in
Van't Veer (2002). This algorithm is yet to be
implemented in the PEASE simulator.
A few remarks need to be made with regard
to the Propeller Control Algorithm or PCA:
- The coefficients in the PID controller have been
selected for stable operation without a proper
optimization of the coefficients. The coefficients
have been determined only for following seas in
sea state 4, and have been applied unchanged
throughout the rest of the study. Optimization of
the controller is likely to improve the results for
other conditions.
- In optimizing the Propeller Control Algorithm,
due attention needs to be given to the
characteristics of the prime mover, i.e. the diesel
engine.
The dynamic cavitation inception limits occur
significantly earlier in terms of thrust coefficient or in
terms of cavitation number than is the case for steady
state inception in calm water. The difference in
inception must be caused by variations in the wake
field and dynamics in the ambient pressure.

120

Perc. time free of cav.

100

80

60
SS4, PCA1-rpmcontrol
SS4, PCA2-Pcontrol

40

SS2, PCA1-rpmcontrol
SS2, PCA2-Pcontrol
SS6, PCA1-rpmcontrol

20

0
0.40

SS6, PCA2-Pcontrol

0.60

0.80

1.00

1.20

Vs [-]

Figure 13 Effect of sea state on cavitation inception


speed at bow quartering waves (H135 deg) and
various PCAs. All ship speeds are normalized with
Vref(4,135).

The effect of sea state on CIS appears to be


significant, however, which is a confirmation of the
conclusions found in the previous phase of the CRN
Project. The non-dimensional speeds cited in Figure
13 all refer to the calculated inception speed for sea
state 4, heading 135 deg in an earlier phase of the
Propellers in Service project, where a constant speed
and constant propeller rotation rate were imposed. The
deviation of the inception speed value from unity in
sea state 4 is a measure of the effect of the dynamic
simulation.
CONCLUSIONS AND RECOMMENDATIONS
Conclusions
This paper presents results of a simulation model for
estimating the effects of operational conditions on
propeller cavitation inception performance. The results
are obtained from a prototype analysis package for
simulating naval ship operations in a seaway. This
simulation tool has a number of aspects that are not
yet complete and therefore will require further
development and validation. The step by step
approach followed currently in developing this
complex simulator allows for a regular evaluation of
its progress. As such, this paper can be regarded as
one of a series of milestones reached in achieving the
goals of this project.
The important effect of sea state on
Cavitation Inception Speed, as found for example in
the sea trials of HM Tydeman described earlier in this
paper, is qualitatively confirmed with the current
PEASE simulator. The inception speed in bow

Recommendations
It is recommended that an improved formulation for
the time dependent added resistance in waves be
implemented in the PEASE simulator, which currently
uses a constant time averaged value.
Little attention has been given to the proper
selection and optimization of the control coefficients
in the Propeller Control Algorithm. Most of the
attention was given to the following seas condition in
sea state 4. This has likely resulted in the PCA not
showing a significant effect on Cavitation Inception
Speed in bow quartering waves. More attention needs
to be given to the optimization of the control system

14

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

for a wider range of wave frequencies. This will also


necessitate a closer look at the dynamics of the driving
engine and its control system.
A better understanding is required of the
differences obtained between results for steady
inception in calm water and dynamic inception in a
seaway. This will ultimately result in improved control
and design strategies that delay the Cavitation
Inception Speed.
A better understanding of the sensitivity and
validity of intermediate results and their effect on the
resulting dynamic behaviour of propeller cavitation is
required. Ultimately, an understanding of the
uncertainty in the simulation and its resulting statistics
will be required.
A number of specific tasks were
acknowledged at the beginning of the current project,
which could not be addressed within the scope of the
present study. The most important tasks remaining are:
To derive a simulation model for propeller
cavitation behaviour during a maneuver.
To study the potential benefits of propeller pitch
control, including continuous wave frequency
control in a seaway.
To assess the potential benefit of an operator
guidance system.
To develop and implement an improved algorithm
for tip vortex cavitation inception.
To study the sensitivity of cavitation inception for
the unsteady pressures in the propeller disk and to
improve the pressure prediction if necessary.

Stapersma, D., Interaction between propulsor and


engine, Proceedings of the 34th WEGEMT School,
Delft, June 2000.
Van Gent, W., "Unsteady lifting surface theory for
ship screws: Derivation and numerical treatment of
integral equation", Journal of Ship Research, Vol 19,
No. 4, pp 243-253
Van Gent, W. and Hoekstra, M., "Force field
approach for propeller-wake interaction", MARIN
Report No. 44303-7-SR, March 1985.
Van Terwisga, T.J.C., Assenberg, F. and Van't Veer,
R., "The PEASE Simulation Tool", MARIN Report
No. 16639-5-CP/ICT, Nov. 2002.
Van't Veer, R., "Validation of PRECAL predicted
wake velocities with model test data", MARIN
Report No. 16639-3-CPS, Dec. 2001.
Van't Veer, R., "Time varying added resistance in
waves and its implication on the forward speed",
MARIN Report No. 16639-4-CPS, Nov. 2002.
Van't Veer, R. "Theory manual for PRECAL 5.0",
MARIN Report No. 17926-2-CPS, Nov. 2003.
Van Wijngaarden, H.C.J. et al., "'Propellers-InService Effects' - CRN Project, Phase III - Report of
Desk Study", MARIN Report No. 14449-1-RD, Feb.
2000.

REFERENCES

Van Wijngaarden, H.C.J., "Study of velocity and


pressure components in the propeller plane", MARIN
Report No. 16639-2-CP, Feb. 2001.

Aalbers, A.B. and Van Gent, W.; Unsteady wake


velocities due to waves and motions measured on a
ship model in head waves, NSMB report No. 573,
1984.

Verkuyl, J.B. and Van Terwisga, P.F.; Testing a


new full scale cavitation observation system on board
of H.M. Tydeman, NCT 50 Proceedings, Newcastle
University, April 2000.

Kuiper, G., Grimm, M., McNeice, B., Noble, D.,


Krikke, M., Propeller Inflow a Full Scale During a
Manoeuvre, Twenty-Fourth Symposium on Naval
Hydrodynamics, Fukuoka, Japan, July 8-13, 2002,
Vol. II, Session 13: Wake Dynamics, pp. 234-249.

15

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics


St. Johns, Newfoundland and Labrador, Canada, August 8-13, 2004

Roll Motion in a Nonlinear Pseudo-Spectral Ship Motion Model


Ray-Qing Lin 1 and Bradley L. Campbell 2
David Taylor Model Basin (NSWC/CD), U. S. A.
9500 Macarthur Blvd, West Bethesda, MD 208175700
E-mail 1 LinRQ@nswccd.navy.mil, 2 CampbellBL@nswccd.navy.mil.

ABSTRACT
Finite amplitude ship motions are predicted using a
fully nonlinear pseudo-spectral ship motion model.
At each time step, the forces on the submerged
surfaces of the ship and ship body restoring
buoyancy forces are transfer to the ship coordinates
reference frame, the fully nonlinear ship motions
are calculated in the ship reference frame without
computing linear forcing coefficients and
parameters, such as added mass and damping (Lin,
et al., 2004a, 2004b; Lin & Kuang, 2004b). This
approach gives model results that are more
accurate, especially for a complicated hull. The
model is also computationally more efficient than
linear ship motion models because we only
calculate one velocity potential instead of seven.
As is well known, the beam of the ship is much
smaller than the length of the ship; therefore, to
simulate the roll motion accurately is much more
challenging than the simulation of pitch and heave
motions. A comparison of our new model results
with experimental data for pitch and heave motions
was presented in (Lin, et al., 2003). For pitch and
heave motions, our model results agree well with
the experimental data not only when the Froude
number is small, but also when the Froude number
is large (Lin and Kunag, 2004b). In this study, we
will compare the roll motions between our pseudospectral model results and the experimental data for
two ONR hull forms: Flared Hull and Tumblehome
Hull, in regular wave beam seas. The new model
results agree well with the experimental data. The
differences in roll motions between the new model
results and experimental data are within the
measurement errors.
In this paper, we will also present the theory
for the nonlinear pseudo-spectral model, including
the unique methods for solving the body boundary
condition for arbitrary ship shapes, the free surface

boundary condition, and the radiation boundary


conditions.
INTRODUCTION
For over a century naval architects have sought to
accurately predict the motion of ships in waves
using both model testing and mathematical
predictions. In 1861 Froude was the first to study
roll motion. Modern computations began in 1950s.
There are two major theories. The first one uses
statistical theory to study the ship responses in a
seaway, such as St. Denis & Pierson (1953). A
number of scientists, such as Peters & Stoker
(1957) and Newman (1961) used multiple small
parameters to improve the statistical theory.
Unfortunately their theory didnt compare well
with experiments. The second approach used
linear ship motion theories to predict the responses
of the ship to regular waves. Korvin-Kroukovsky
(1955) and Korvin-Kroukovsky & Jacobs (1957)
first developed slender-body strip theory to predict
the heave and pitch. Their numerical results were
comparable with experiments. In 1969, Ogilvie &
Tuck used systematic analysis to determine the
added mass and damping for heave and pitch
motions in head seas and improved strip theory to
the point that it became mathematically consistent..
Maruo (1970) was first to bridge the gap between
strip theory and slender-body theory. Strip theory
is based on a short wavelength assumption, and
slender body theory is based on a long wavelength
assumption.
Since the 1970s, computer power has rapidly
increased along the science and technology of
three-dimensional ship motions. Scientists started
to more precisely simulate ship motion by using
the Neumann-Kelvin approach where the free
surface and unsteady ship motion both are
linearized. These type models, usually employ
finite different, finite element, or boundary element
methods (Hess & Smith, 1964; Dawson, 1977;

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics


St. Johns, Newfoundland and Labrador, Canada, August 8-13, 2004

Beck & Magee, 1990, Bai, et al., 1992; Magee,


1994; Sclavounos, 1996; Scorpio & Beck, 1998;
Qui, et al., 2001).
Since nonlinear term may be equal to or
greater than linear terms (Lin, & Segel, 1988),
recently, scientists have focused on threedimensional,
fully
nonlinear
calculations,
especially when predicting unsteady ship motions.
Ship motions are affected by many physical
phenomena, such as wave-wave interactions
between the ship and incident waves, viscosity,
wave breaking, and ship waves generated by ship
motion, etc. Different physical phenomena occur
at different scales, following different physical
laws, and require different approaches. There is no
method that can include all the physical laws
precisely. The goal of a method can only to
develop the model based upon the major physical
laws which best addresses the particular
computation, a model then may be able to predict a
very complicated physical phenomena precisely
when the developer couples a number of these
individual models. There are two major types of
models used in the study of ship motions: viscous
or inviscid flow. Viscous flow models are often
used in calm water, because the viscosity is more
important when water is calm. Viscous flow
models such as time domain Reynolds Averaged
Navier-Stokes (RANS) equations (Wilson, et al.,
1998; Gentaz, et al., 1999; Xing, et al., 2001) are
often used for these calm water flow predictions
For seakeeping predictions, people often use
inviscid flow (potential flow), because the
nonlinear wave-wave interactions between ships
and incident waves are dominant. These wavewave interactions are large scale, and the viscous
flow around the ship is less important. Prediction
codes such as SWAN (Sclavounos, 1996, Yifeng,
1998, Sclavounos, et al., 1998, Kim, et al., 1999),
Large Amplitude Ship Motion Program (LAMP)
(Engle, et al., 1997; Weems, et al., 2000) are
examples of such codes. However, all these
potential flow models have numerical stability
problems due to breaking of short waves near the
ship. Furthermore, their unsteady ship motion
predictions use linear added mass and damping
coefficients from linear strip theory.
The development of a robust and full nonlinear
ship motion model for both strong and weak threedimensional wave-wave interactions between
ambient surface waves and ships in arbitrary water
depth for arbitrary ship forms is very important to
ship seakeeping, and ship and harbor designs. To
reach our goal, we use a global pseudo-spectral
method with a finite element or finite difference

ship boundary (Lin, et al., 2004a; Lin & Kuang,


2004a). The model not only can be used to study
the both strong and weak three-dimensional wavewave interactions between ambient surface waves
and arbitrary ships, but also is very efficient. The
model convergent speed is N log N instead of N2,
where N is the number of unknown variables.
Furthermore, we developed a new method that
transfers the singularity points of ship pressure to a
smooth function by calculating the ship pressure in
ship normal vector coordinate system (Lin, et al.,
2004b). This allows us to study very complicated
hull forms, including multi-hull ships, accurately
and efficiently even at high speeds.
Most recently, we upgraded the new model
from full nonlinear steady ship motion to full
nonlinear unsteady ship motion.
Instead of
calculated linear added mass, damping coefficients,
etc, and seven individual velocity potentials, we
calculated the exciting force on the exact wetted
surface and body restoring forces. This method
only requires us to calculate one total velocity
potential. This full nonlinear unsteady ship motion
model is not only accurate, especially for
complicated hull forms in large-steep incident
waves, but is also computationally efficient. In this
study, we compare the roll motions predicted using
the new model with experimental data for the ONR
flared and tumble-home hulls.
MATHEMATICAL MODEL
In this model, fully nonlinear refers to ship
motion generated waves, wave-wave interactions
between ship and incident wave, and the ship
motion itself. Previous fully nonlinear ship motion
models usually not only used some linearized terms
to replace the full nonlinear terms to avoid the
numerical stability problems due to breaking of
short waves near the ship, but also didnt include
fully nonlinear ship motion itself. Often they used
linearized strip coefficients for added mass and
damping coefficients, and used eight potential
velocities for three-dimensional transfer motions,
and three dimensional rotational motions, plus
incident wave, and ship waves generated by ship
motions separately. However, the new model
calculates only total free surface elevation , and
total velocity potential in Eulerian coordinate,
and the ship motion is calculated in ship coordinate
reference system with mass center as the origin.
The basic equations are described as following.
I. Hydrodynamics:

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics


St. Johns, Newfoundland and Labrador, Canada, August 8-13, 2004

For an incompressible fluid in an Euleriancoordinate system. The governing field equation


is:
2 2 h +

2
z

= 0,

for - H

z .

(1)

where h denotes the horizontal gradient and


is the total velocity potential, which includes the
velocity potential generated by ship motion, the
incident waves, and forced ship motion.
The dynamic and kinematic boundary
conditions on the free surface, z = and is the
total free surface elevation, which includes incident
waves and ship waves, are expressed as:

p u
1
+ ( u s ) + g + s x
t
2
t

(2)

h = 0
and

+ ( h ) ( h u s ) =
,
(3)
t
z
respectively, where is the kinematic viscosity of
the fluid, and us is total horizontal transfer velocity,
P is pressure due to the ship moving, is density
of the water, g is gravitational acceleration, x is
position vector, z is the vertical distance from the
origin.
The radiation boundary condition is:
* e
=
us,
x
x

* = e ;

* = e + u s; * = e + ;

X=b

(4)

G G
x=c

(5)

G
where b is a forward boundary, and c is a side and
aft boundary. The subscript e denotes the
incident wave. In calm water, both e and e are
equal to zero. Superscripts, *, denote values on
the boundaries at a distance from the ship.

(6)

The boundary condition on the ships hull

is
n ( t ) = n ( t ) v s( t ) ,

v s ( t ) = u s ( t ) + u sz ( t ) + s ( t ) xr ,

(7)

( t ) is the unit normal vector on the ships


where n

where s (t ) is rotational ship motion velocity


vector, u sh (t ) is the translation in the z direction
(heave motion), and r is the distance from wetted
surface to the center of mass.
The ship boundary in our model is described
in local fine grids, which are generated from a
boundary element representation of the ship.
Our model employs a pseudo-spectral
technique to solve the problem represented by
equations (1)(9). The time-dependent velocity
potential and wave elevation are each represented
by a two-dimensional Fourier series in the x- and ydirections. The nonlinear terms in the equations
(2) and (3) are computed in the physical space at
collocation points, which are then transformed
back into the spectral space (expansion
coefficients) via FFTs. At each time step, the
pressure on the ships surface is first calculated
on the very fine grid representing the ship's
surface, and then transformed back to the (regular)
collocation points via bilinear interpolation (Lin, et
al., 2004a). The free surface elevation and velocity
potential on the ship's surface are also transformed
between the fine grid boundary and the collocation
points. For the details of the model, we refer the
reader to Lin, et al. (2004a). How to obtain the
nonsingular pressure on the hull surface can be
found in Lin, et al., (2004b).
II. Equations of Motion:
The transfer motions can be described as:
m ship

du s k
dt

+ Dtrans k u sk = Fk

dt

= s ( t ) xn (t ) ,

I +D+ R

+ Fk

restore

(10)

translation motion, Dtrans k is damping coefficient,

Fk

I + D+ R

and Fk

restore

are the total forcing on wet

surface and the total body response force, and


k=1,2,3 represent in x, y, z direction respectively.
The rotational motions can be described as:

surface, and

dn ( t )

(9)

where m ship is the hull displace, u sk is the ship

On the bottom, z = -H,

+ ( h H ) ( h u s ) = 0.
z

where vs is total velocity vector, which includes the


ship's forward speed, and the forced translations
and rotations, as follows:

(8)

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics


St. Johns, Newfoundland and Labrador, Canada, August 8-13, 2004

(1) m I jk
k =4

+ Dk

6
d 2 k
+
k x[(1) m I jk k ]

dt 2
k =4

d k
= j I + D + R + j Re store ,
dt

j = 4,5,6.
(11)

m = 2 if j = k
m = 1 if j k
Roll motion (k, j = 4), pitch motion (k, j = 5), and
yaw motion (k, j = 6):

I jk = 0

if j = 4 and k = 5, or j = 5 and k = 4,

or j = 5 and k = 6, or j = 6 and k = 5, therefore,


when we calculate the roll motion in beam seas, we
must also calculate the yaw motion, as well as
transfer motions. From equation (11), the roll and
yaw motions can be described as:
d 2 4

COMPARISON BETWEEN NEW MODEL


RESULTS WITH EXPERIMENTAL DATA
I. Hull Forms
ONR Flared and Tumblehome hulls are used in
these comparisons. The below water portions of
these two hulls are the same, but above water, the
sides of the Flared hull slope +10 degree outward,
and Tumblehome hull slope 10 degree inward.
Renderings of the three-dimensional models of
both hulls are shown in Figure 1.

d 2 6

+ 4 x[( I 44 4 ) ( I 46 4 )]
dt 2
dt 2
d 4
+ D4
= 4 I + D + R + 4 restore ,
dt
d 2 6
d 2 4

+ 6 x[( I 66 6 ) ( I 64 6 )]
I
I 66
64
dt 2
dt 2
d 6
+ D6
= 6 I + D + R + 6 restore ,
dt
(12)

I 44

I 46

is descried in Lin and Kunag (2004b). Solving


equations (10) and (12), we can obtain the roll
motion 4 ( t ) in beam seas.

where I jk is mass moment of inertia, ( t ) is roll


4

d 4
) is roll angular velocity;
dt
d
6 is yaw angle , and 6 = ( 6 ) is yaw
dt
angular velocity; D4 and D6 are roll and yaw
angle , and 4 = (

damping coefficients respectively, which are the


function of Froude number and obtained by code
5500 model tests.

I +D+ R

is total force due to incident waves,

diffracted waves, radiation over the wet surface as


well as the distance between each grid point of the
wet surface and hull mass center, and j

Re store

is

body restoring force due to buoyancy force and the


distance between buoyancy center and hull mass
center. The above ship motion model should be
defined in ship coordinate, and mass center as
original, but the water line is defined on Eulerian
coordinate, where the z=0 is at free surface. The
detail of obtaining the j

I +D+ R

and j

Figure 1 ONR ship hulls, right hand side is flared hull


and left hand side is tumblehome hull.

Full scale, both hulls displace 8790 tons. Their


waterline length is 154 meters, the beam is 18.8
meters, the draft is 5.5 meters, and the freeboard is
9 meters. The center of mass of both hulls is at
2.6, 0., 2.74 meters in x, y, z directions of Eulerian
coordinates, where the geometry center is the
origin and z=0 is free surface. The x direction is
positive forward, y direction is positive, to port,
and the z direction is positive up. The model scale
ratio is 32. Figure 2 shows the body plan for both
of the models. In the following model results,
L=12261, M=192, N=192, where L is number of
grid points on the whole ship boundary, and M and
N are collocation points in the x- and y-directions
of the whole domain respectively.

Re store

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics


St. Johns, Newfoundland and Labrador, Canada, August 8-13, 2004

Body Plan
15000

N=0.5

N=1

N=1.5

N=2

N=2.5

N=3

N=3.5

N=4

N=4.5

N=4.67

N=4.71

N=5

N=5.5

N=6

N=6.5

N=7

N=7.5

N=8

N=8.5

N=9

N=9.5

N=10

N=10.5

N=11

N=11.5

N=12

N=12.5

N=13

N=13.5

N=14

N=14.5

N=15

N=15.5

N=16

N=16.5

N=17

N=17.5

N=18

N=18.5

N=19

N=19.5

-0.04

N=20

N=20.5

N=21

-0.06

13000

11000

9000

Z (mm ABL)

7000

moment and restoring moment predicted by the


new model a) for flared hull; b) for tumblehome hull.
The figure shows that the roll restoring moment is
opposite to the roll exciting moment.

N=0

Roll Exciting Force on Flairsided Hull (dr36)


Fn=0.19, /L=2.5, wh=3.41m, 1.25m span

5000

0.06

3000

-5000

-3000

-1000
-1000

1000

3000

5000

7000

9000

11000

13000

15000

-3000

0.04
E4//g/a/s/L

1000

-5000
Y Offset (mm)

Roll Exciting
Force

0.02
0
-0.02 0

N=0

N=0.5

N=1

N=1.5

N=2

N=2.5

N=3

N=3.5

N=4

N=4.5

N=4.67

N=4.71

N=5

N=5.5

N=6

9000

N=6.5

N=7

N=7.5

7000

N=8

N=8.5

N=9

N=9.5

N=10

N=10.5

N=11

N=11.5

N=12

N=12.5

N=13

N=13.5

N=14

N=14.5

N=15

N=15.5

N=16

N=16.5

N=17

N=17.5

N=18

N=18.5

N=19

N=19.5

15000

13000

Z (mm ABL)

11000

Roll Exciting Force on Tumblehome Hull (dr23)


Fn=0.19, /L=2.25, wh=2.84m, 2.5m span

3000

-1000
-1000

1000

3000

5000

7000

9000

11000

13000

15000

-3000

-5000
Y Offset (mm)

N=20

Figure 2b Body plan for tumblehome hull.

0.06
0.04
E4/ /g/a/s/L

1000

-3000

Restore Force

Figure 3a Time series of normalized rolling exciting


moment and restoring moment, calculated by the new
model for flared hull in test dr36, the Froude number =
0.1945, incident wavelength is equal to 2.5 ship length,
wave height is 3.41 meters and bilge keel span is 1.25
meter.

5000

-5000

15

Time

Figure 2a Body plan for flared hull.

Body Plan

10

Roll Exciting
Force

0.02
0
-0.02 0

10

15

Restore Force

-0.04
-0.06

To associate collocation points, we used a 64 by


128 wave number matrix. The exciting force and
moment, and the restoring force and moment, are
calculated on the fine grid on the hull surface.
II.
Roll Exciting Moments and Restore
Moments in the New Model
Unlike ship motion models that use linear strip
theory and added mass and damping coefficients to
study the unsteady ship motion, the new model
calculates the unsteady ship motions by calculating
the exciting forces and restoring forces exactly.
The translated motions are described in Equation
(10), and the rotational motions are described by
Equation (11). The ship rotational motions are
based on the rotational torque and restoring torque.
The roll exciting moment is due to the incident
wave forces, diffraction forces, and radiation forces
on each grid point of the wetted surface of ship
and the distance from the center of mass center to
those points on the wetted surface. The restoring
moment is due to buoyant response. In this study,
all incident waves represent beam seas. Figure 3
shows the time series of normalized roll exciting

Time

Figure 3b Time series of normalized rolling exciting


moment and restoring moment, calculated by new model
for Tumblehome Hull in test dr23, the Froude number is
0.1945, incident wavelength is equal to 2.25 ship lengths,
wave height is 2.84 meters and bilge keel span is 2.5
meter.

The difference of the absolute values between


restoring moment and exciting moment for both
hulls is less than 50 percent, but their phases are
different. The roll exciting moments and restoring
moments in other cases have similar patterns, but
are not shown.
III. Comparison New Model Results with
Experimental Data
Figure 4 shows the time series of normalized roll
motion for the two hulls. One can see that the
numerical model results agree very well with the
experimental data.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics


St. Johns, Newfoundland and Labrador, Canada, August 8-13, 2004

Roll Motion on Flairsided Hull (dr36)


Fn=0.19, /L=2.5, wh=3.41m, 1.25m span

Roll Motion on Flairsided Hull


Fn=0.19, 1.25m span

3
New Model
Results

2
0
-2 0

10

15

Experimental
Data

X4/Ka

X4/Ka

New Model
Results

Experimental
Data

-4

-6

Figure 4a Time series of normalized roll angles for the


flared hull in test dr36, the Froude number = 0.1945,
incident wavelength is equal to 2.5 ship length, wave
height is 3.41 meters and bilge keel is span 1.25 meter,
solid line represents the results from the new model and
the dashed line represents the experimental data.

Roll Motion on Tumblehome Hull


Fn=0.19, 2.5m span
3

4
New Model
Results

2
0
5

10

15

Experimental
Data

-4

X4/Ka

X4/Ka

Figure 5a shows the normalized roll angle for the flared


hull as a function of the incident wavelength, Froude
number = 0.1945, the bilge keel span is 1.25 meters, the
solid line represents the new model results and the
dashed line represents the experimental data.

Roll Motion on Tumblehome Hull (dr23)


Fn=0.19, /L=2.25, wh=2.84m, 2.5m span

-2 0

2
/L

Time

New Model
Results

1
0
0

Experimental
Data

/L

-6
Time

Figure 4b Time series of normalized roll angles for the


tumblehome hull in test dr23, the Froude number is
0.1945, incident wavelength is equal to 2.25 ship lengths,
wave height is 2.84 meters and bilge keel span is 2.5
meter, solid line represents the results from the new
model and the dashed line represents the experimental
data.

The difference between the numerical model


results from the new model and experimental data
is within the measurement error 15 percent for both
ships. In fact we have compared the new model
results with experimental data for more than 10
random cases. All results show that the numerical
model results agree very well with the
experimental data, and the difference between the
new model results and experimental data are all
within the measurement errors (15%), Figure 5
provides the computed and measured roll response
operators for the two hull forms. Both new model
results and experimental data show that the
normalized roll angle increases with the incident
wavelength increases, until they reach the
maximum roll angle when incident wavelength
equals 2.252.5 ship length.

Figure 5b shows the normalized roll angle for the


tumblehome hull as a function of incident wavelength,
Froude number = 0.1945, the bilge keel is 2.5 meters, the
solid line represents the new model results and the
dashed line represents the experimental data

Furthermore, Figure 5 also shows that in the most


cases, the normalized roll motion angles of the
Tumblehome hull is slightly greater than that of
the Flared hull, but the maximum normalized ship
roll motion angles are not significantly different
between two hulls. However, it should be noted
that the bilge keels on the Tumblehome Hull have a
2.5 m span, while the bilge keels on the Flared
Hull have a 1.25 m span. This indicates the roll
angles of Tumblehome hull would greater than
those in Flared hull when the bilge keels are the
same.
The Figure 6 shows time series of normalized
roll motion angles for Tumblehome hull, a)
Fn=0.19, and b) Fn=0.33 (dr25).

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics


St. Johns, Newfoundland and Labrador, Canada, August 8-13, 2004

Roll Motion on Tumblehome Hull


Fn=0.19, /L=1.0, wh=1.22m, 2.5m span
1.5
1
New Model Results

X4/Ka

0.5
0
-0.5

10

15

Experimental Data
(Fn=0.33)

-1
-1.5
Time

Figure 6a Time series of normalized roll angle for the


tumblehome hull, incident wavelength is equal to ship
length, wave height is 1.22 meters and bilge keel span is
2.5 meter, solid line represent the results from the new
model in test dr25, Froude number = 0.19, and the
dashed line represents the experimental data in test dr25.

Roll Motion on Tumblehome Hull (dr25)


Fn=0.33, /L=1., wh=1.22m, 2.5m span
1.5
1
New Model
Results

X4/Ka

0.5
0
-0.5 0

10

15

Experimental
Data

-1
-1.5
Time

Figure 6b Time series of normalized roll angle for the


tumblehome hull in test dr25 incident wavelength is
equal to ship length, wave height is 1.22 meters and bilge
keel span is 2.5 meter, solid line represents the results
from the new model, and the dashed line represents the
experimental data,.

Both cases are in the same environmental


conditions, such that incident wavelength is equal
to ship length, wave height is 1.22 meters, and
bilge keel span is 2.5 meters. Unlike the head sea
cases, the pitch and heave depend highly on the
ship forward speed (Froude Number), in the beam
sea cases, the normalized roll angles are not
significantly dependent on the ship forward speed,
at least when Froude number is not so large. One
can see that the roll angles between Figures 6a and
6b are not significantly different. This is because
the frequency of incident wave in ship coordinate
system is:

s = e

e 2us
g

cos ,

(13)

where g is gravity accelerate velocity,

is the

frequency of incident wave in earth coordinate, u s

is ship forward speed, and is the angle between


the incident wave propagation direction and the
ship forward direction.
In the head sea
case, = , and in the beam sea case = /2.
Equation (13) shows the frequency of incident
waves in ship coordinate, svaries with the ship
forward speed, u s in head sea case, but the
frequency of the incident wave in ship coordinate
doesnt vary with ship forward speed u s in the
beam sea case.

DISCUSSIONS
Unlike linear ship motion models., the new model
(Lin, et al., 2004a, b; Lin and Kuang, 2004a, b)
follows the first principle by calculating the exact
rolling torque and restore torque. The new fully
nonlinear ship motion model results agree well
with the experimental data for both Flared and
Tumblehome hulls, and the difference between
numerical model results and experimental data are
within
the
measurement
errors
(15%).
Furthermore, the new model results can identify
the different roll motion angles between the two
hulls, which a linear theory cannot. The roll of
Tumblehome hull are greater than those of Flared
hull.
The model results also shows that the Froude
number (ship forward speed) doesnt significantly
impact the roll motions in beam seas, at least when
Froude number is not large. The impact of Froude
number on roll when in oblique seas needs to be
examined.
Finally we should point out that, the new
model is much more efficient computationally than
previous ship motion models (Lin, et al., 2004a and
b, Lin and Kuang 2004a).

ACKNOWLEDGEMENTS
This work was partly supported by a grant from
Office of Naval Research managed by Dr. Pat
Purtell. R. Lin is supported by the grants from the
David Taylor Model Basin, Carderock Division,
Naval Surface Warfare Center Independent
Laboratory In-House Research (ILIR) program
administered by Dr. John Barkyoumb. We want to
thank Bruce Webster and Mike Davis of the David
Taylor Model Basin, Hydromechanics Department,
without their help, this work could not have been
done. We also want to thank Richard C. Bishop,
and Beverly S. Simon of the Seakeeping Division
of the DTMB Hydromechanics Department who

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics


St. Johns, Newfoundland and Labrador, Canada, August 8-13, 2004

provided the experimental data. Finally we want to


thank Dr. Reed for his comments and suggestions.

REFERENCES
Bai, K,. J. Kim & H. Lee (1992) A Localized
Finite-Element Method for Nonlinear FreeSurface Wave Problems. Proc. 19th Symp.
Naval Hydro., Seoul., South Korea, pp. 90
114.
Beck, R. F. & A. R. Magee (1990) Time-Domain
Analysis For Predicting Ship Motions, in
Dynamics of Marine vehicles and Structures in
Waves, W. G. Price, P. Temarel & A. J.
Keane, Eds. Elsevier Science Publishers B. V.,
pp. 4964.
Dawson, C. W. (1977 A Practical Computer
Method For Solving Ship-Wave Problems.
Proc. 2nd Intl. Conf. Num. Ship Hydro.
Berkeley, CA, pp. 3038.
Engle, A., Lin, W. M, N. Salvesen, & Y. S. Shin
(1997) Application of 3-D Nonlinear Wave
Load and Structure Reponse Simulations in
Naval Ship Design. ASNE Naval Engineer J.
Froude, W. (1861) On the Rolling of Ships. Trans.
Inst. Naval, Arch., 2:180229.
Gentaz, L., P. E. Guillermo, B. Alessandrini & G.
Delhommeau (1999) Three-dimensional FreeSurface Viscous Flow Around A Ship in
Forced Motion. Proc. 7th Intl. Conf. Num.
Ship Hydro., Paris France, 12 p.
Hess, J. L. & A. M. O. Smith (1964) Calculation of
Nonlifting Potential Flow About Arbitrary
Three-Dimensional Bodies.. J. Ship Res.,
8(2):2244.
Kim, Y., S. Kim, D. Renick & P. D. Sclavounos
(1999) Linear and Nonlinear Flows and
Responses of Ships by a Rankine Panel
Method. Proc. 7th Intl. Conf. Num. Ship
Hydro., Paris, France.
Korvin-Kroukovsky, B. V. (1955) Investigation of
Ship Motions in Regular Waves. Trans.
SNAME., 63:386435.
Korvin-Kroukovsky, B. V. & W. R. Jacobs (1957)
Pitching and Heave Motions of a Ship in
Regular Waves. Trans. SNAME., 65:590632.
Lin, C. C. & L. A. Segel (1988) Mathematics
Applied to Deterministic Problems in the
Natural Sciences.
Classics in Applied
Mathematics, SIAM, Macmillan, New York.

Lin, R.-Q. & W. Kuang (2004a) Nonlinear Waves


by a Steadily Moving Ship in Environmental
Waves. J. Marine Technology, Vol. 8: 109116.
Lin, R.-Q. and W. Kuang (2004b) Solid-body
Motion in a Fully Nonlinear Ship Motion
Model. Submitted to the J. of Ship Res.
Lin, R.-Q,, W. Kuang, & A. M. Reed (2004a) Fully
Nonlinear Wave-Wave Interaction Between
Ship and Surface Wave, Part I: Ship Wave in
Calm Water Condition. J. Ship Res. (in press).
Lin, R.-Q., W. Kuang, & A. M. Reed (2004b) Full
Nonlinear Wave-Wave Interaction Between
Ship and Surface Wave, Part II:
Ship
Boundary Condition. J. Ship Res., (in press).
Lin, R.-Q., A. M. Reed, & W. Belknap (2003)
Fully Nonlinear Wave-wave Interaction
Between a High-Speed vessel and Incident
Waves. FAST2003, Vol. 1, pp. A3-18.
Magee, A. (1994) Seakeeping Applications Using
A Time-Domain Method, Proc.20th Symp.
Naval Hydro., Santa Barbara., CA, 19 p.
Maruo, H. (1970) An Improvement of The Slender
Body Theory For Oscillating Ships With Zero
Forward Speed. Bull. Fac. Eng. Yokohama
Natl Univ., 19:4556.
Newman, J. N. (1961) A Linearized Theory for the
Motion of a Thin Ship in Regular Waves. J.
Ship Res., 3(1):119.
Ogilvie, T. F. & E. O. Tuck (1969) A Rational
Strip Theory of Ship Motion, Part I. Report
013, Dept. Nav. Arch. Mar. Eng. Univ.
Michigan, Ann Arbor, ix+92p.
Peters, A. S. & J. J. Stoker (1957) The Motion of A
Ship, As a Floating Rigid Body, in a Seaway.
Comm. Pure Appl. Math., 10:399490.
Qiu, W., S. Ando & C. C. Hsiung (2001) A PanelFree method For The Time-Domain Radiation
Problem. Proc. 16th Intl. Workshop Water
Waves Float. Bodies, Hiroshima, 2001, Japan,
4p.
Sclavounos, P. D. (1996) Computations of wave
Ship Interactions, in Advances in Marine
Hydrodynamics.
M.
Ohkusu,
Ed.,
Computation
Mechanics
Publications,
Southampton, pp. 23378.
Sclavounos, P. D., S. Kim & Y. Kim (1998)
Nonlinear Load and responses of ships and

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics


St. Johns, Newfoundland and Labrador, Canada, August 8-13, 2004

Offshore Platforms in Deterministic and


Random Waves.
Wave98 Ocean Wave
Kinematics, Dynamics and Load On Structure,
Houston, USA.
Scorpio, S. M. & R. F. Beck (1998) A Multiple
Accelerated Desingularized method for
Computing Nonlinear Wave Forces On
Bodies.
J. Offshore Mech. Arctic Eng.,
120(2):7106.
St. Denis, M. & W. J. Pierson (1953) On The
Motion of ships in Confused Seas. Trans.
SNAME, 61;280354.
Weems, K., W. M. Lin, S. Zhang (2000) Time
Domain predictions For Motions and Loads of
Ships and Marine Structure in Large Seas
Using a Mixed-Singularity Formulation. Proc.
4th Osaka Colloquium on Seakeeping
Performance of Ships, Osaka, Japan.
Wilson, R., E. Paterson & F. Stern (1998)
Unsteady RANS CFD Method For Naval
Combatants in Waves. Proc. 22nd Symp.
Naval Hydro., Washington, D. C., pp. 53249.
Xing, Y., I. Hadzic, S. Muzaferija, & M. Peric
(2001) CFD Simulation of Flow-Induced
Floating-Body Motions.
Proc. 16th Intl.
Workshop Water Waves Float. Bodies,
Hiroshima, Japan, 4 p.
Yifeng, H. (1998) Nonlinear Ship Motions by a
Rankine Panel Method. PhD Thesis, MIT.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
John F. ODea
Naval Surface Warfare Center, Carderock Division,
USA
Although ship motions have often been
estimated with linear methods, it has long been
recognized that prediction of rolling motion must
take into account significant nonlinear effects. Since
ship roll motion is typically lightly damped, large
resonance response can occur in waves. As a result,
nonlinear effects on both the dynamics (viscous
damping) and statics (GZ curve) are recognized to be
important.
For these reasons, a fully nonlinear approach
to the roll problem, including viscous effects as well
as potential flow, would be welcomed. The authors
recognize in their introduction that others (Wilson,
Gentaz, Xing) are working with RANS codes to solve
the damping problem, although the authors imply
these others are only applying their work to calm
water flow predictions (the titles of the referenced
papers indicate otherwise). The mathematics in the
present paper appears to be strictly potential flow. In
the rigid body equations of motion (eq. 10-12), the
damping forces involving the coefficients Dk are
strictly linear since the coefficients are multiplied
only by linear velocity terms. The authors state that
D4 (for roll) and D6 (for yaw) are taken from model
tests. Thus we have the rather surprising result that a
fully nonlinear ship motion model resorts to using
empirically obtained damping coefficients, and these
coefficients are linearized!
The paper then goes on to make
comparisons with a set of hull forms in which the
geometry is varied above the static waterline (flared
and tumblehome hulls). This variation should reveal
a difference in at least the restoring moment (a
difference in GZ). The roll exciting moment and
restoring moment are compared in Figs. 3a (Flare)
and 3b (Tumblehome). However, the comparison is
clouded by the fact that so many other parameters are
varied between these two figures wave amplitude,
bilge keel size and wavelength (the caption under
Fig. 3a contradicts the label on the figure in this
case). The authors also state that the phase
relationship for exciting and restoring moment differ
for the two hulls, but that is not obvious from these
figures.
The authors then make comparisons to
model test results for the two hulls. Figs 4a, b
compare the predictions to test results, indicating
good agreement. However, the results are not for
large motions. Roll motion appears to be only twice
wave slope, and wave steepness appears to be

/H~100. The labels and captions are confusing and


contradictory as to wavelength, so it is not obvious
whether these results are at roll resonance. Similarly,
in Fig. 6a, b the correlation is for /L=1.0 (far from
roll resonance) and the roll motion is only as large as
wave slope (which again appears to be /H~100. It is
likely that any linear roll motion prediction code,
when calibrated with measured damping coefficients
as was done here, would correlate equally well.
In light of the title of the paper, it is
unfortunate that the authors did not make any
comparisons of their model to data where
nonlinearities might be demonstrated. The conditions
presented (low wave steepness) just do not show roll
motion large enough to reveal any obvious
differences in roll response that can be attributed to
the Flare vs. Tumblehome above-water hull form. A
series of varying wave steepnesses, with wavelength
and period chosen specifically to match the roll
resonant period, might have been able to demonstrate
such differences. Even without such systematic
model tests, a series of predictions in these varying
conditions might have gone a long way toward
demonstrating
the
different
roll
response
characteristics. As it stands, I cannot agree with the
authors conclusions that the new model results can
identify the different roll motion angles between the
two hulls, which a linear theory cannot.
AUTHORS REPLY
Reading ODeas comments, it seems that I
have not clearly explained the differences between
this new model and traditional models.
Before
answering his questions I would like to try to further
explain the differences, and the mathematics and
physics involved, as I think this will help in the
understanding of my replies to ODeas comments.
The new ship motion model is entirely different from
all the previous ship motion models. Except for
viscosity dissipation, it is based on first principles,
especially in solving solid body motion and
predicting the ship motion induced waves.
The newly developed Nonlinear PseudoSpectral Ship Motion Model includes two coupled
models: (A) A Nonlinear Pseudo-Spectral Steady
Ship motion model, which is based on the Equations
(1) to (9) in the paper; (B) A Solid-body motion
model, which includes two types of ship motions: (a)
rotational motion in Equation (10) of the paper;
where s=4 is roll motion, s=5 is pitch motion, and
s=6 is yaw motion, and (b) the transfer motions as
described by:

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

d 2 s
d
I + D+ R
Re store
m 2 + D ( t ) s s = Fs
+ Fs
dt
dt
(1r)
where m is total mass of the ship, s is
transfer displacement, Fs I + D + R is total force, and
Fs Re store is restoring force, D (t ) s is the transfer
dissipation term, where s=1 is the surge motion, s=2
is the sway motion, and s=3 is the heave motion.
As one can see the kinematics and
dynamic equations (2) and (3) in the paper are both
nonlinear equations. Therefore, the ship motion
induced waves and wave-wave interactions between
incident waves and ship bodies are all calculated
nonlinearly. They include the resonant phenomena
when the ship frequency and the frequencies of
incident waves or the frequencies of ship motion
induced waves are similar.
It is well known that a linear
model can be strictly potential flow, but a nonlinear
model can not be strictly potential flow. The

dissipation term in equation (2) in the paper


shows that the new ship motion model is not strictly
potential flow, but bi-potential flow. This dissipation
term is higher order wave number dependent. The
more short wave length waves present, the more
dissipation.
In the solid body motion model, we solve
the solid body motions in Equation (10) of paper and
Equation (1r) exactly, by calculating the forces and
restoring forces exactly, except viscosity dissipation.
The forces and restoring forces are based on the
actual ship displacement (segment by segment) as
defined by the wave water line intersections with the
hull, and the moment arms of those segments. Unlike
previous ship motion models, we dont use linear
parameterization, such as a11, b11, , b44 , etc. We
2

also dont use eight velocity potentials to study the


ship motions, we use one total velocity potential and
one free surface (Lin and Kuang, 2004a and b).
The ship motions, and the ships
attitude, which are calculated by model (B) are
transferred through the ship boundary condition in
Equations (7)-(9) in the paper into model (A). Model
(A) calculates the ship motion generated waves, and
their dissipation.
Model (A) also calculates the nonlinear ship
pressure, which is directly used in calculating the
forces in model (B).
ODea in his first comment said that the
nonlinear effects are very important in roll motion,
the new model is strictly linear potential flow, and

D4 and D6 are taken from the model tests, so the


new model is not a nonlinear model.
I agree with ODea that the nonlinear
effects are very important to roll motion, in fact not
only to roll motion, but in all six degrees of ship
motion. This is the major reason why we are
developing this new nonlinear ship motion model. As
I described above the new ship motion model solves
the solid body motion Equations (10) in the paper,
and equation (1r) using first principles, except
viscosity for a bi-potential flow. Unlike previous
linear ship motion models which linearize the
velocity potential to eight velocity potentials for each
degree of freedom, ship generated wave, and incident
wave; the new model uses only one total velocity
potential and one free surface elevation. Furthermore,
unlike previous ship motion models which use
parameters to calculate each motion and restoring
force, we calculated the forcing and restoring forces
exactly, with no parameters involved. The ship
motion damping, which is due primarily to the ship
motion induced waves and their dissipation, is not
calculated by D4 and D6 . In fact they are not
calculated in Equation (10-12) of model (B) as
ODea comments. They are calculated in the
Equation (2) and (3) of model (A). D4 and D6 , as
described in equation (10) in the paper, are solid
body motion viscosity dissipation, which does not
include any ship motion induced waves or their
dissipation. D4 and D6 are very small effects, less
than 1% of the total restoring force. If there is a bilge
keel, ship motion will induce more short wave length
waves. The greater the bilge keel size, the more
induced short wave length waves and dissipation
because in Equation (3) is function of wave
number. As we described above, this is calculated in
model (A), not in model (B), and certainly is not
related to D4 or D6 .
2

ODea in his second comment said that the


comparison of roll exciting moment and restoring
moment in Figure 3 (a) Flare, (b) Tumblehome
included too many parameters: wave amplitude, bilge
keel size, and wavelength. He also noted that the
authors stated that the phase relationship for exciting
and restoring moments differ for two hulls. ODea
commented that this is not obvious from these
figures.
ODea has misinterpreted Figure (3). The
goal of Figure 3 was simply to show the exciting
moment and restoring moment as calculated from
first principles by our model. The parameters listed

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

are simply those for which the ONR Forms were


tested. The comments on phase made in the paper
are comparing the phase of exciting moment and
restoring moments. Differences in phasing between
the two hulls are not mentioned. Please see page 5 in
the paper. However, there is a typographical error in
the figure caption of Figure 3a and as well as Figure
4a, it should be the incident wavelength is equal 2.5
ship lengths as indicated in the title parameter. The
number, 2.5, is missing in the original paper.
ODea in his third comment said that
authors make comparison to the model tests for the
two hulls in Figure 4, indicating good agreement, but
the results are not for large motions.
I agree with ODea saying the roll motion is
not a large roll motion. The maximum roll motion in

used pitch and heave motions predictions to


demonstrate the difference between extreme motion
predictions from the new model and traditional
models. Lin and Kuang (2004c) compared the
predicted heave and pitch motions of the Delft
Catamaran No. 372 at various wave-length/shiplength ratios for different Froude numbers for: the
new model, the VERS numerical model (MarinTek),
and observational data from MARIN and Delft
Model tests. The comparison is shown in Figure 1,
(a) for heave motion, and (b) for pitch motion.

Heave Motion On Delft Catamaran No. 372 (Fr=0.75)


4
3.5

30 for the ONR Form ships.


ODea in his fourth comment said that
any linear motion prediction code, when calibrated
with measured damping coefficients as was done
here, would correlate equally well.
In the new model that part of roll damping
which is due to ship motion induced waves as
described above, is calculated in model (A), not
model (B), and not included in D4 and D6 .

D4 and D6 are a function of wetted ship surface,


ship geometry, mass characteristics, etc.
D4 and D6 are not based on first principles.

3
X3/ a

the paper is 30 . This is because the experimental


data didnt have the roll motion data greater than

New Model

2.5

VERS

Delft Model Test

1.5

MARIN Model Test

1
0.5
0
0

D6 are very small, less than 1% of the restoring

force. in model (A) does not change for ship


motion calculations, it is the kinematic viscosity of
the fluid as described in the paper. It is not a constant
and it is determined by wave theory. The New
Model is based on first principles, and does not
require the condition by condition tuning.
ODea in his fifth comment said that authors
did not compare their new model with other existing
models.
There is very little experimental data for
extreme roll motion published.
While the
experimental data used for comparison in this paper
does not have the large amplitude data that ODea
and the authors would like, in another publication I

1.5

2.5

Figure 1a Normalized heave motions vary


with incident wave-length/ship-length. Solid line
represents the new model results, dashed line
represents the numerical model results from VERS,
square dots represent the Delft Model test, and
triangles represent the MARIN Model test.

Pitch Motion On Delft Catamaran No. 372 (Fr=0.75)


3.5
3
2.5

X5 /ka

fixed for all the calculations, and for all different


conditions. As described above, the values of D4 and

/L

D4 and D6 have constant coefficients. These


coefficients are obtained from calm water data, not
from data for every test. When the coefficient is
determined, the formulations of D4 and D6 are

0.5

New Model
VERES

Delft Model Test


MARIN Model Test

1.5
1
0.5
0
0

0.5

1.5

2.5

/L

Figure 1b Normalized pitch motions vary


with incident wave-length/ship-length. Solid line
represents the new model results, dashed line
represents the numerical model results from VERS,
square dots represent the Delft Model test, and
triangles represent the MARIN Model test.
The new model results agree well with the
observational data even when the Froude number is
equal to 0.75 in the resonant situation (maximum
motion), but for the same Froude number, the results
from VERS diverge significantly from observational

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

data in the resonant situation. The VERS results


agree with the observational data only when Froude
number is very small (Lin and Kuang, 2004c). If the
previous models cant predict the pitch and heave
motions for resonant conditions of large Froude
number (strong nonlinear effects) correctly, how can
they predict the roll motion correctly? Nonlinear
effects usually are stronger in roll motions then in
pitch and heave motions.
ODea in his sixth comment said that since
the roll motion angle is small, I cannot agree with
authors conclusions that the new model results can
identify the different roll motion angles between the
two hulls, which a linear theory cannot.
The Flair and Tumblehome hulls are exactly
same below the water-line, but are very different
above the water-line, any roll motion will result in
different motions from the two hulls, especially when

Lin, R.-Q. and W. Kuang, 2004b. SolidBody Motion in Full Nonlinear Ship Motion Model,
submitted to J. of Ship Research.
Lin, R.-Q. and W. Kuang, 2004c. SolidBody Motion in Full Nonlinear Ship Motion Model,
submitted to J. of Ship Research.

the roll motions are 30 . Experimental data show


that the roll motions between two hulls are different,
as did the new model. The previous ship motion
models, which are based on the linear theory, cant
accurately predict such differences because those
numerical models calculate solid body motion, by
using eight linearized velocity potentials. In addition
they are generally parameterizations using GM, etc,
to estimate the restoring moment. They will not
accurately predict the different restoring moments
between two hulls, even when the roll motions are
not large.
The authors look forward to comparison of
the new model with existing ship motion models for
data sets with extreme motions, both now and when
future planned enhancements have been made to the
model.
In conclusion, ODeas comments are based
on not knowing the differences in the mathematics
and physics between the new model and previous
models, especially in the solid body motions in model
(B). It is hoped this reply has explained the
differences.

REFERENCES
Lin, W. M. and D. K. P. Yue, 1990:
Numerical solution for large-amplitude ship motions
in time-domain, Proc. 18th Symp. Naval Hydrod., U.
Michigan, Ann Arbor, Michigan.
Lin, R.-Q. and W. Kuang, 2004a. Nonlinear
Waves by a Steadily Moving Ship in Environmental
Waves, J. of Marine Technology, Vol..8, 109-116.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Krish Thiagarajan
University of Western Australia, Australia
Interesting Paper! I am interested in a case
of a moored ship in bidirectional waves (waves
approaching from two different directions at the same
time). Here, nonlinear interaction between the
incident waves is important. Also, the restoring
forces on the vessel are nonlinear due to mooring.
Can the authors method be applied to this situation?
AUTHORS REPLY
Thank you for your comment! Yes, the new
model can predict such a case. The new model solves
the six degrees of freedom ship motion equations
exactly, except for viscosity which currently uses
parameterization. The new model uses one potential
velocity to model six degrees of freedom of ship
motion, incident wave, and ship generated waves.
The restoring forces on the vessel due to the mooring
are nonlinear, but you have to add the mooring forces
to simulate your particular scenario. Since we only
used one total potential velocity for all forcing,
motions, ship motion generated wave, the new model
can handle two incident waves coming from different
directions at the same time. In fact the new model
can handle numerous waves coming from numerous
directions at the same time. No matter how many
waves, they can be decomposed into two directions.
We have compared the new model results and
experimental data in bow seas, which included roll,
pitch, and heave motions at the same time (Lin et al.,
2004). From a mathematical point of view, this is
similar to the case you suggested. The fact that
wavelengths of incident waves are different is not a
problem.
REFERENCE
Lin, R-Q., Bruce Webster and Michael Davis and
Bradley L. Campbell, Richard C. Bishop, and Terry
Applebee, 2004. A Comparison Between A Fully
Nonlinear Pseudo-Spectral Ship Motion Model and
Model Tests in the NSWCCD Seakeeping Basin for
A Study of Roll and Pitch Motion NSWCCD Tech.
Report.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Solomon C. Yim
Oregon State University, USA
Thank you for a very nice presentation. The
title of your talk is Roll Motion Is your model
sufficiently general to analyze all 6 degrees of
freedom? If not, what extensions are needed?
AUTHORS REPLY
Thank you for your comment! The model is
sufficiently general to analyze all degrees of freedom.
Unlike the previous ship motion models which used
eight velocity potentials for each degree of freedom
of ship motion, plus the ship motion generated wave
and incident wave. Therefore, the six degrees of
freedom ship motions are linearized. The new model
solves the six degrees of freedom ship motion
equations exactly, except for the viscosity where it
currently uses parameterization. The new model only
uses one total velocity potential for six degrees of
freedom motion, incident waves, and ship motion
generated waves. Therefore, the new model not only
can predict the six degrees of ship motion, but also
can predict the interactions between the six degrees
of ship motions. In this paper, we compared the
model results with experimental data in roll motions,
in other papers, such as Lin and Kuang, (2004), we
compare the new model and experimental data for
pitch and heave motions. Of course, I am not saying
the new model can predict arbitrary motion in
arbitrary environment for arbitrary ship body at this
moment. The new model is still under development.
REFERENCE
Lin, R.-Q. and W. Kuang, 2004. Solid-Body Motion
in Full Nonlinear Ship Motion Model, submitted to J.
of Ship Research.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics


St. Johns, Newfoundland and Labrador, CANADA, 8-13 August 2004

Validation of Ship Motion Predictions with Sea Trials


Data for a Naval Destroyer in Multidirectional Seas
Kevin McTaggart and Dave Stredulinsky
(Defence R&D Canada - Atlantic, Canada)
ABSTRACT

where ship motions must be duely considered. Ship


motions also influence the effectiveness of crew, weapon systems, and sensor systems. The recent trend
toward large-scale simulations of military systems
is creating increased demand for modelling of ship
motions and their influence on ship systems. Depending on simulation requirements, appropriate
trade-offs must be made between computational requirements and prediction fidelity.

This paper presents comparisons of ship motion


predictions with measured motions for the Canadian naval destroyer HMCS Nipigon in seaways with
significant wave heights ranging from 3.7 m to 5.8 m.
The numerical predictions include modelling of encountered directional seaways, which were measured
by a directional wave buoy. Motions were predicted
using a strip theory code and with DRDC Atlantics
new ShipMo3D library, which uses a panel method
based on the frequency domain Green function for
zero forward speed. To investigate the influence of
appendages on radiation forces, frequency domain
computations included treatment of the combined
panelled hull and panelled appendages. Time domain computations were performed for both linear
and nonlinear evaluation of hull forces due to buoyancy and incident waves. Agreement between predicted and observed RMS motions and zero-crossing
periods is generally good. Differences between results from the five different sets of predictions are
relatively minor.

To address ongoing requirements for ship


motion predictions, DRDC Atlantic is developing
the ShipMo3D library for simulation of ship motions and sea loads in waves. The library includes
modules for computations in both the frequency
and time domains. For time domain computations,
nonlinear computations of buoyancy and incident
wave forces are available. To meet requirements
of military simulation applications, it is intended
that computations be available that can run faster
than real-time without significantly compromising
prediction fidelity.
Validation with experimental data is an essential part of development of any ship motion prediction code. When considering the relative merits of model tank tests and full-scale trials, model
tank tests typically have the advantages of lower
costs, greater control over encountered conditions,
and completeness of data. However, full-scale trials
are considered essential because they include phenomena that can be absent in model tests due to
oversight or scaling effects. Consequently, both fullscale trials and model tests are recommended for
validation of numerical ship motion predictions. To
support ongoing validation efforts, DRDC typically
conducts one or two sea trials per year measuring

INTRODUCTION
Predictions of ship motions in waves are being used
by navies for a widening number of applications. In
the past, ship motion predictions have been used
primarily for engineering design applications. In
recent years, there has been increasing interest in
ship motion predictions by naval operators due to
the influence of ship motions on safety and operational effectiveness. Search and rescue missions
and ship-borne helicopter operations are examples

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

ship motions and sea loads in waves. The trials are


usually held off the coast of Nova Scotia in the winter months, when rougher sea conditions are more
likely to occur. This paper describes the validation
of ship motion predictions with sea trial data for
the destroyer HMCS Nipigon.

mined from previously computed frequency domain


values, similar to the approach used by Ballard et
al. (2003). Comparisons of hydrodynamic forces
and motions for the frequency and time domains
have verified correct prediction of time domain coefficients from frequency domain coefficients.
Motions are currently computed for a ship
travelling at nominally steady speed and heading.
Motion computations are performed using a translating earth axis system. Nonlinear buoyancy and
incident wave forces can be evaluated based on the
instantaneous wetted surface. Nonlinear buoyancy
and incident wave forces are initially computed in
a ship-fixed axis system and then rotated to the
translating earth axis system. Other hydrodynamic
forces (i.e., added mass, retardation, and wave diffraction) are assumed to maintain their translating
earth axis system values regardless of large amplitude ship motions.

OVERVIEW OF SHIPMO3D
LIBRARY FOR PREDICTING
SHIP MOTIONS IN WAVES
DRDC Atlantic has been an active developer, user,
and sponsor of codes for predicting ship motions
and sea loads since the 1970s. Several different
codes are currently used, with each code having
its own particular strengths and weaknesses. The
in-house strip theory code SHIPMO7 (McTaggart
et al. 1997, McTaggart 2000) was originally developed by Schmitke (1978) and has a large user base.
Other codes for which DRDC Atlantic has been a
co-sponsor are used when time domain predictions
are required or when three-dimensional hydrodynamic effects are important.

SHIPMO3D APPENDAGE FORCE


PREDICTIONS
Sway, roll, and yaw motions are significantly influenced by appendage forces. In addition, viscous
forces acting on the ship hull make a noticeable contribution to roll damping. These forces are modelled in ShipMo3D based on approaches presented
by Schmitke (1978) and Himeno (1981). Unfortunately, roll motions continue to present a significant
challenge for ship motion codes due to difficulties
in predicting roll damping components.

The ShipMo3D library is an effort to consolidate ship motion prediction capabilities into a
unified entity that will satisfy a range of application
requirements. ShipMo3D uses a panel method to
provide enhanced accuracy relative to strip theory,
and to ensure applicability to the range of vessels
operated by the Canadian Navy. Hydrodynamic
coefficients are computed in the frequency domain
based on the Green function for zero forward speed,
similar to approaches presented by Beck and Loken (1989) and Papanikolaou and Schellin (1992).
The decision to use this method was based partly
on the relatively slow speeds of present and anticipated Canadian naval vessels, which operate at
Froude numbers less than 0.4. Other considerations included computational efficiency and robustness relative to methods using non-zero forward
speed Green functions in either the time or frequency domains. Comparisons with experiments
presented by McTaggart et al. (1997) and Schellin
et al. (2002) indicate that the zero forward speed
Green function can lead to very good motion and
sea load predictions at moderate ship speeds.

When examining methods for predicting


viscous roll damping components, there is great
variability in predicted values for bilge keel drag
forces. The viscous roll moment acting on a bilge
keel can be expressed as:
Z xf ore
1
3
Cd (x) s(x) reBK
(x) dx (1)
F4bkv = 42
2
xaf t
where is water density, 4 is roll velocity, xaf t
and xf ore are the aft and forward longitudinal coordinates of the bilge keel, Cd (x) is the local drag
coefficient, and s(x) is the local span. The effective
cube of the local viscous roll moment arm, rev3 (x), is
given by:

The ShipMo3D library provides motion predictions in both the frequency and time domains.
For time domain predictions, hydrodynamic coefficients, including retardation functions, are deter-

3
reBK
(x)

3
= rBK
(x) (y cos + z sin )

(2)

where rBK (x) is the radius from the local bilge keel
center to the ship center of gravity (CG), y is the
2

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

lateral coordinate (+ port) relative to the ship CG,


is the dihedral angle of the bilge keel relative to
horizontal (+ upward), and z is the vertical coordinate (+ upward) relative to the ship CG. Neglecting hull interaction and free surface effects, a long
flat plate normal to flow will have a drag coefficient
of approximately 1.2. In contrast, bilge keel drag
coefficients given by Lloyd (1989) vary between 2
and 12, with the upper value occurring for small
amplitude roll motions.

radiation and diffraction computations by assuming that appendages are thin and modelling them
using dipole panels (Chakrabarti, 1987). Due to
the significant variation of dipole strengths across
a flat plate (Meyerhoff, 1970), each appendage typically has several panels in each of the span-wise
and chord-wise directions.

NAVAL DESTROYER HMCS


NIPIGON

DRDC Atlantics strip theory code SHIPMO7 models bilge keel viscous drag using the approach described by Schmitke (1978) based on the
work of Kato (1966). Katos formulation is quite
complicated, but can be simplified to the following
for a representative destroyer:
!0.6
rBK (x) b4 e
p
Cd (x) 5
g s(x)



11Rbilge (x)
1.0 + 3.5 exp
(3)
rBK (x)

The naval destroyer HMCS Nipigon (Figure 1) was


the subject of a comprehensive sea trial measuring ship motions and sea loads in December 1997.
HMCS Nipigon was the last steam-driven destroyer
in the Canadian Fleet. Figure 2 gives a body plan
for Nipigon and Table 1 gives particulars. Nipigon
appendages include 2 rudders, 2 outer propeller
shaft brackets, 2 inner propeller shaft brackets, 2
bilge keels, and a skeg, with dimensions as given in
Table 2.

where Rbilge (x) is the local bilge radius of the hull,


b4 is roll amplitude, and e is wave encounter frequency. Perhaps the most noteworthy feature of
Equation (3) is the large degree of variation of the
drag coefficient with roll velocity amplitude b4 e .
When considering Equations (1) and (3) together,
the proportionality of viscous roll drag to (b
4 e )1.4
combined with the presence of gravitational acceleration in Equation (3) possibly arise from trying to
model radiation and viscous forces simultaneously.
The ShipMo3D library currently models the
bilge keel drag coefficients as follows:
!
b

4
Cd = Cdref (b
4ref )
(4)
b
4ref

Figure 1: HMCS NIPIGON

where Cdref is an input reference drag coefficient


for roll velocity amplitude b
4ref and is an input
decay exponent. This approach provides flexibility in the selection of bilge keel roll damping coefficients, including the possibilities of using input
values from experiments or CFD computations.

SEA TRIAL ON HMCS NIPIGON


During the December 1997 sea trial, a series of 71
trial runs of 20 to 30 minute duration was undertaken at two nominal ship speeds in head, bow,
beam, quartering and following seas. Data were
collected in higher sea states (4, 5 and 6) for three
days over December 2, 3 and 4th and for four days
at lower sea states (2 and 3) over December 8 to
11th. The low ship speed was about 8 knots and

Appendage force predictions by Schmitke


neglect free-surface effects. The influence of the
hull on appendage potential flow forces is modelled by assuming that the local hull is a planar
boundary. The influence of the appendages on hull
forces is neglected. The ShipMo3D library introduces an option for including appendages in hull
3

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

...
...
.
...
.
...
.
...
.
...
.
...
.
...
.
...
.
...
.
...
.
....... ....... ....... ....... ....... ....... ....... ....... ....... ....... ......... ....... ....... ....... ....... ....... ....... ....... ....... ....... .....
.
...
..
..
..
...
..
..
..
...
..
..
..
...
..
..
..
...
..

..
... ..........
... ...............
.
.
... ... . .....
... .... ... ... ...........
.
.. . .. . . . .....
........................................
.. ... .. .. .. ... ..............
...... ... ... ... .. ..
.
.
.
.
. . . . .. ......
.. . .
...... ... ... .. .. ...
.. ... ... ... ... ..... ...............
......... ..... ..... .... .... ....
.
. ..
.
.. .. .. .. .. .. .... ........
......... ... ... ... ... ...
... ... ... ... ... ... ... .........
............... ..... ..... .... .... ....
..
............ ... ... .... .... ....
.. .. .. .. .. .. .. .......
........... ... ... ... .. ..
........... ... .. ... .... ...
... ... ... ... ... ... ... .......
........... .. ... ..... ...... ......
.............. .... ..... .......... ............................................... ... ... ... ... ... ... ... ........
.
............... .... ........ .............................
.
............... .... ...........
................ ... ... ... ... ..... ...... .... ..........
.............................. ................
.
.
.
.
.
.
.
.
.
................................................ ......................... .. .. .. .... ...... ...... ...........................
...
..
..........................................................
.
................................................ ................ .. .. .... ...........................................................................
. ................................................ ........ .. .. ..............................................
.......................................... ........................................ ..
.. ....................................

Table 2: Appendages for HMCS Nipigon


Rudders (2)
Station
19.3
Root lateral offset
1.981 m
Root above baseline
3.197 m
Span
3.023 m
Root chord
2.286 m
Tip chord
1.886 m
Dihedral angle (port)
-90 deg
Area (port+starboard)
12.6 m2
Outer shaft brackets (2)
Station
18.3
Root lateral offset
3.109 m
Root above baseline
3.200 m
Span
2.256 m
Root chord
0.800 m
Tip chord
0.800 m
Dihedral angle (port)
-99.5 deg
Area (port+starboard)
3.6 m2
Inner shaft brackets (2)
Station
18.3
Root lateral offset
1.158 m
Root above baseline
2.957 m
Span
2.240 m
Root chord
0.800 m
Tip chord
0.800 m
Dihedral angle (port)
-64.2 deg
Area (port+starboard)
3.6 m2
Bilge Keels (2)
Fore station
8.82
Aft station
13.79
Span
0.610 m
Area (port+starboard)
32.9 m2
Skeg at Centerline
Fore station
14.0
Aft station
16.5
Aft span
1.236 m
Area
6.8 m2
Station 20 is aft perpendicular

Figure 2: Body Plan for HMCS Nipigon

Table 1: Particulars for HMCS Nipigon


Displacement
Length between perpendiculars
Beam
Draft
Trim by stern
Height of CG above keel
CG from forward perpendicular
Block coefficient
Prismatic coefficient
Wetted hull area
Metacentric height
Roll radius of gyration
Pitch radius of gyration
Natural roll period

3027 tonnes
108.4 m
12.8 m
4.3 m
0.5 m
5.1 m
57.5 m
0.503
0.625
1446 m2
1.2 m
5.6 m
27.1 m
10.6 s

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

high ship speed was between 14 and 18 knots depending on what speed the ship could maintain in
the given sea state. Speed was limited due to the
loss of one of the two ship boilers on the first day
of the trial.
Trial instrumentation included a ship motions package, a wave buoy, a TSK over-the-bow
wave height meter, an array of 24 pressure transducers outfitted in the hull below the waterline, 15
single strain gauges, and 4 rosette strain gauges.
Wave data were collected for 20 minutes out of
every hour from the wave buoy. An Endeco type
1156 directional wave buoy was used for the first
3 days of the trial, and a type 956 buoy was used
for the last 4 days. Figure 3 shows an Endeco wave
buoy being deployed from Nipigon. The main data
acquisition system was a PC-based LabVIEW system. All sixty instrumentation channels were digitally sampled at 20 Hz. Time histories, statistical
distributions and minimum and maximum values
were determined and recorded.
For the present study only the ship motion
and wave data have been used. A limited number
of runs from the first three days of the trial was
selected to satisfy the following 3 criteria:
the significant wave height Hs was greater
than 3 m,
there was a clear dominant wave direction and
minimal directional wave spreading,
the nominal relative wave direction was oblique
(bow quartering, beam, or stern quartering
seas).
Table 3 gives a summary of the runs selected for
the present study, with Tz denoting zero-crossing
wave period. Figure 4 shows an example measured
wave spectrum.

Figure 3: Endeco Wave Buoy Being Deployed from


HMCS Nipigon

Rudder deflections were not measured during the sea trial, and are assumed to have had no
effect on ship roll motions. In reality, the rudder
motions could have significantly influenced the ship
roll motions, particularly if rudder motions were
at a frequency similar to the ship natural roll frequency.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

NUMERICAL PREDICTIONS OF
SEA TRIAL MOTIONS

Table 3: NIPIGON Trial Runs Used in Present


Study
RunSpeed Hs Tz Wave dir.
Ship
Rel.
(kt) (m) (s) (from, deg) head. wave
Mean Dev. (to, deg) head.
203 8
3.73 7.73 218
39
265
BQ
204 8
3.67 7.21 225
39
85
SQ
206 16 3.89 7.39 228
25
355
SQ
209 13 4.75 8.07 224
33
265
BQ
210 15 4.75 8.07 224
33
85
SQ
303 8
5.82 9.45 274
41
310
BQ
304 8
5.57 8.81 272
39
120
SQ
305 8
5.57 8.81 272
39
220
BQ
306 8
5.16 8.66 266
45
50
SQ
309 14 5.39 8.95 279
32
315
BQ
310 14 5.44 8.73 269
43
135
SQ
403 8
5.01 9.34 244
40
290
BQ
404 8
4.90 9.60 238
47
110
SQ
409 16 4.52 8.37 238
50
285
BQ
410 16 4.52 8.37 238
50
105
SQ
413 8
4.98 8.86 245
40
330
Bm
Relative wave headings
BQ - Bow quartering
SQ - Stern quartering
Bm - Beam

Numerical predictions of ship motions during sea


trial runs were made using the strip theory program
SHIPMO7 and with the ShipMo3D library.
All predictions considered the ship motions
in the measured short-crested seaways. Four different types of predictions were made using the
ShipMo3D library:
frequency domain predictions, appendages not
included in hull radiation and diffraction computations,
frequency domain predictions, both hull and
appendages included in radiation and diffraction computations,
time domain predictions with linear buoyancy
and incident wave forces,
time domain predictions with nonlinear buoyancy and incident wave forces.
The time domain predictions do not include the appendages in the hull radiation and diffraction computations.
For three-dimensional radiation and diffraction computations, the wet hull was represented by
814 flat panels (407 panels on the port side), with
most panels being quadrilaterals and the remainder
being triangles. The dry hull, required for computing nonlinear buoyancy and incident wave forces,
was represented by a total of 1,276 panels. Figure 5 shows the panelled hull, with the wet portion
represented by white panels and the dry portion
represented by grey panels. For radiation and diffraction computations including panelling of the appendages, the port and starboard appendages were
modelled using 308 panels (154 panels on the port
side), and the centerline skeg was modelled using
18 panels. Figure 6 shows the port half of the hull
in red, and appendages on the port side and centerline in yellow. Figure 7 shows expanded views of
the panelled aft appendages.

0
0.2 Hz
0.15
0.1
0.05
270

90

10

10

180

Contours at 10, 20, 40, 60, and


80 % of peak: 0.72 m2/(Hz-deg)

Hydrodynamic coefficients in the frequency


domain were computed at the zero and infinite frequency limits and at intermediate encounter frequencies of 0.1, 0.2, . . ., 6.0 rad/s. For each encounter frequency, computed terms include the zero

Figure 4: Measured Directional Wave Spectrum for


Run 203

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

speed coefficients and terms proportional to ship


speed U and U 2 that can be used to evaluate hydrodynamic coefficients at arbitrary forward speed.
Wave diffraction terms were evaluated for speeds of
0, 5, . . ., 30 knots, relative wave directions of 0, 10,
. . ., 180 degrees, and wave frequencies of 0.2, 0.3,
. . ., 2.0 rad/s. The computed hydrodynamic coefficients and diffraction forces form a database from
which hydrodynamic forces can be obtained for any
seaway encountered by the ship.
Figure 5: Nipigon Wet (White) and Dry (Grey)
Panelled Hull

For time domain simulations, retardation


functions were computed using convolution integrals applied to previously computed frequency domain coefficients. The retardation functions were
computed for delay times of 0.0, 0.1, . . ., 20 s. Retardation forces are negligible beyond delay times
of 20 s. Time domain simulations were conducted
using both linear and nonlinear forces from buoyancy and incident waves. For nonlinear incident
wave forces, Wheeler stretching (Wheeler, 1970)
was used to model the variation of velocity potential with distance below the instantaneous free surface. The time domain simulations used a time step
size of 0.2 s and total duration of 30 minutes.
The bilge keel roll damping was modelled
using Equation (4) with a reference drag coefficient
of 16.5 for the ship rolling with an amplitude of 5
degrees at its natural roll period of 10.6 s. The decay coefficient has a value of 0.60. Both the reference drag coefficient and decay coefficient were
obtained from calculations using Katos method.
Figure 8 shows the variation of bilge keel drag coefficient with roll amplitude for oscillations at the
ship natural roll frequency. During time domain
computations, it was necessary to select a nominal roll velocity amplitude for computing the bilge
keel drag coefficient. At each time step, the nominal roll velocity amplitude was set to 1.25 times
the RMS roll velocity during the most recent 20 s
of motion. All other appendages were assumed to
have constant drag coefficient values of 1.17.

Figure 6: Aft Portion of Nipigon Wet Panelled Hull


(Red) with Appendages (Yellow)

Bilge keel
...
.................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................................
. ...............................................................................................................................................................................

Skeg
......
...................... .
................................. .. .....................
.....................................................................................................................................................................................................................................

Inner ..bracket
......
...........
......................
.................
...............................

Outer ...bracket
.....
..........
.
.......................
.........................
.....................

Time domain simulations with the ShipMo3D library currently are limited to a ship with
nominally steady speed and heading. Due to the
absence of rudder motion data during the sea trial,
rudder motions were assumed to be small. This
assumption will have little influence on predicted
roll motions if the rudder motions were limited to
low frequency. To ensure course-keeping, additional
stiffness and damping terms given in Table 4 were
used for surge, sway, and yaw. The stiffness and

Rudder
....................
...... ... ... .. .. ....
...... ... ... ... ... ......
..................................
...... ... ... ... ... ......
.................................
............................
.................

Figure 7: Profiles of Nipigon Panelled Appendages

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

50

...
...
...
...
...
...
...
...
...
...
...
...
...
...
...
...
...
...
...
...
...
...
....
....
....
.....
.....
......
.......
........
.........
...........
.............
...............
....................
.........................
..................................
................

Drag coefficient Cd

40
30
20
10
0

10

the instantaneous wetted hull, computational times


were approximately one third as fast as real-time.

COMPARISONS OF SEA TRIAL


DATA AND PREDICTIONS
Tables 5 to 10 give observed and predicted values of RMS displacements and zero-crossing periods for heave, roll, and pitch. Agreement between
observed and predicted values is generally good,
with the best agreement occurring for heave, followed by pitch and roll, respectively. The five different prediction methods give quite similar results.
Figures 9 to 14 show scatter plots of predicted versus observed RMS motions and zero-crossing periods, where the predictions are from time domain
simulations with nonlinear buoyancy and FroudeKrylov forces.

20

Roll amplitude 4 (deg)

Figure 8: Bilge Keel Drag Coefficient Versus Roll


Amplitude at Ship Natural Roll Period of 10.6 s
damping values were selected such that they would
have only a minor influence on predicted motions,
and are based on natural periods of approximately
100 s and damping of approximately half of critical
damping.

INFLUENCE OF APPENDAGE
FORCE PREDICTIONS ON ROLL
HYDRODYNAMICS
The frequency domain predictions presented in Tables 5 to 10 indicate that including the panelled
appendages in radiation and diffraction computations has very little influence on predicted hydrodynamic forces for Nipigon. Figures 15 and 16 show
roll added mass and damping versus ship speed for
motions at Nipigons natural roll period of 10.6 s.
Added mass is non-dimensionalized by ship roll inertia I44 , and damping is non-dimensionalized by
cr
critical roll damping at zero speed B44
. Results are
given for zero amplitude roll motions (i.e., viscous
roll forces are negligible). Figure 15 indicates that
treating the appendages separately from the hull
gives slightly higher added mass than when using
panelled appendages integrated with the hull radiation computations. As expected, the bare hull
has lower added mass than the appended hull. Figure 16 shows that including the panelled appendages
in the radiation computations has negligible effect
on predicted roll damping, with the relevant plotted lines being indistinguishable.

Table 4: Additional Stiffness and Damping for Ensuring Course-Keeping with ShipMo3D Computations
Surge
Sway
Yaw

Stiffness
1 104 N/m
2 104 N/m
2 108 Nm/rad

Damping
2 105 N/(m/s)
4 105 N/(m/s)
2 108 Nm/(rad/s)

Due to strong roll resonance behavior, it


is essential that sufficiently fine increments be used
for wave frequency and wave direction. The computations in both the frequency and time domains
used incident seaway components with wave frequencies of 0.2, 0.25, . . ., 2.0 rad/s and absolute
wave directions of 0, 10, . . ., 360 degrees. For measured directional sea spectra, the number of nonzero wave spectral components within the specified
combinations of wave frequency and wave direction
ranged from 158 to 323.
For the time domain computations using
linear buoyancy and incident wave forces, computational times were approximately three times faster
than real-time on an 800 MHz Pentium III desktop
computer. When nonlinear buoyancy and incident
wave forces were evaluated at each time step on
8

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Table 5: Observed and Predicted


Run Trial Strip
3D freq.
theory Appendages
Sep. Integ.
203 0.70 0.65 0.66 0.66
204 0.55 0.48 0.47 0.47
206 0.76 0.60 0.58 0.58
209 0.84 0.90 0.96 0.96
210 0.89 0.71 0.67 0.67
303 1.38 1.10 1.11 1.11
304 1.02 0.93 0.90 0.90
305 0.98 1.19 1.19 1.19
306 0.81 0.97 0.95 0.95
309 1.33 1.07 1.12 1.11
310 0.97 0.92 0.89 0.89
403 1.15 1.00 1.01 1.00
404 0.88 0.93 0.92 0.92
409 1.13 0.88 0.94 0.94
410 0.78 0.74 0.72 0.72
413 1.09 1.14 1.15 1.15
Predicted/observed
Mean
0.94 0.94 0.94
Deviation 0.14 0.14 0.14

Table 7: Observed and Predicted


Run Trial Strip
3D freq.
theory Appendages
Sep. Integ.
203 2.83 3.37 4.03 4.12
204 3.97 4.28 4.76 4.80
206 4.25 5.15 5.93 6.09
209 2.64 2.54 3.11 3.23
210 4.98 4.97 5.85 6.00
303 4.95 5.24 5.40 5.49
304 6.16 4.91 5.30 5.36
305 3.18 5.35 5.58 5.67
306 4.46 6.17 6.32 6.40
309 3.93 3.45 4.12 4.28
310 5.41 3.91 4.92 5.02
403 3.95 4.60 5.09 5.17
404 4.26 4.75 5.19 5.26
409 2.63 2.68 3.58 3.72
410 3.75 3.46 4.11 4.22
413 4.24 6.15 6.24 6.33
Predicted/observed
Mean
1.10 1.24 1.27
Deviation 0.24 0.22 0.22

RMS Heave (m)


3D time
Hull forces
Linear Nonlin.
0.68
0.68
0.46
0.46
0.59
0.59
0.99
1.00
0.68
0.68
1.01
1.00
0.80
0.80
1.17
1.17
1.00
1.00
1.12
1.13
0.88
0.88
1.05
1.05
0.89
0.89
0.96
0.97
0.72
0.72
1.13
1.13
0.93
0.15

0.94
0.16

Table 6: Observed and Predicted Heave ZeroCrossing Period (s)


Run Trial Strip
3D freq.
3D time
theory Appendages
Hull forces
Sep. Integ. Linear Nonlin.
203 8.1
8.4
8.2
8.3
8.2
8.2
204 9.3
10.3 10.2 10.2
10.3
10.2
206 11.8 10.5 10.2 10.2
10.1
10.1
209 7.3
7.0
6.8
6.9
6.8
6.8
210 11.7 12.0 11.7 11.7
11.5
11.5
303 9.7
9.8
9.7
9.7
9.7
9.6
304 12.2 13.3 13.3 13.3
12.9
12.7
305 8.6
8.7
8.5
8.5
8.3
8.3
306 11.1 10.8 10.6 10.7
10.6
10.6
309 9.6
8.2
7.9
7.9
7.8
7.8
310 11.7 13.3 13.1 13.1
13.3
13.2
403 10.0
9.5
9.3
9.4
9.4
9.5
404 11.2 12.3 12.3 12.3
12.1
12.1
409 8.3
7.5
7.2
7.2
7.0
7.0
410 11.3 12.2 11.8 11.8
11.5
11.5
413 10.2
8.8
8.7
8.7
8.7
8.7
Predicted/observed
Mean
1.00 0.98 0.98
0.97
0.97
Deviation 0.09 0.09 0.09
0.09
0.09

RMS Roll (deg)


3D time
Hull forces
Linear Nonlin.
3.67
3.73
4.22
4.27
5.47
5.59
3.02
3.21
5.99
6.19
4.90
4.89
4.66
4.73
4.75
4.89
5.88
5.93
3.75
4.06
4.64
4.78
4.68
4.75
4.49
4.56
3.47
3.70
4.42
4.56
5.72
5.77
1.15
0.19

1.19
0.20

Table 8: Observed and Predicted Roll ZeroCrossing Period (s)


Run Trial Strip
3D freq.
3D time
theory Appendages
Hull forces
Sep. Integ. Linear Nonlin.
203 9.2
9.7
9.8
9.7
9.8
9.9
204 10.2 10.5 10.5 10.5
10.3
10.5
206 12.3 11.0 11.0 11.0
10.8
10.9
209 8.4
8.5
8.8
8.7
8.8
8.9
210 12.3 11.8 11.7 11.7
11.5
11.5
303 9.5
10.6 10.5 10.5
10.7
10.5
304 11.0 10.9 10.8 10.8
10.7
10.7
305 8.8
10.0 10.0
9.9
10.0
9.9
306 10.7 10.4 10.4 10.4
10.2
10.3
309 9.1
9.6
9.8
9.7
9.7
9.5
310 11.5 12.1 11.8 11.8
12.4
12.4
403 10.0 10.3 10.3 10.3
10.2
10.3
404 10.7 10.7 10.7 10.7
10.5
10.4
409 9.5
9.7 10.1 10.0
9.8
9.8
410 12.1 12.6 12.4 12.3
11.8
12.0
413 10.0 10.2 10.1 10.1
9.9
9.8
Predicted/observed
Mean
1.02 1.02 1.02
1.02
1.02
Deviation 0.06 0.06 0.05
0.06
0.06

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Table 9: Observed and Predicted RMS Pitch (deg)


Run Trial Strip
3D freq.
3D time
theory Appendages
Hull forces
Sep. Integ. Linear Nonlin.
203 1.03 1.14 1.18 1.18
1.19
1.21
204 0.79 0.93 0.88 0.88
0.89
0.90
206 1.00 1.01 0.88 0.88
0.93
0.94
209 1.45 1.83 1.96 1.96
2.02
2.04
210 1.08 1.23 1.10 1.10
1.19
1.21
303 1.66 1.98 2.05 2.05
1.80
1.84
304 1.51 1.39 1.29 1.29
1.26
1.28
305 1.89 1.72 1.79 1.79
1.83
1.85
306 1.44 1.34 1.29 1.29
1.32
1.34
309 1.59 1.84 1.98 1.98
1.99
2.00
310 1.21 1.31 1.22 1.22
1.24
1.25
403 1.37 1.64 1.71 1.71
1.83
1.86
404 1.08 1.29 1.24 1.24
1.31
1.33
409 1.59 1.62 1.75 1.74
1.74
1.75
410 1.04 1.13 1.07 1.07
1.11
1.13
413 1.10 1.21 1.26 1.26
1.25
1.28
Predicted/observed
Mean
1.09 1.09 1.09
1.10
1.12
Deviation 0.10 0.14 0.14
0.15
0.15

Predicted RMS Heave (m)

1.5

....
.....
.....
.....
.....
.
.
.
....
.....
.....
.....
.....
.
.
.
.
.....
....
.....
.....
.....
.
.
.
.
.....
.....
.....
.....
....
.
.
.
.
.....
.....
.....
.....
.....
.
.
.
....
.....
.....
.....
.....
.
.
.
.
.....
....
.....
.....
.....
.
.
.
.
....
.....
.....
.....
.....
.
.
.
.
.....
.....
.....
.....
....
.
.
.
.
.....
.....
.....
.....
.....
.
.
.
....
.....
.....
....

1.0

+
+

0.5

0.0
0.0

0.5

Relative heading
+ Bow quartering
Beam
Stern quartering
1.0

1.5

Observed RMS Heave (m)

Figure 9: Predicted Versus Observed RMS Heave,


Time Domain Predictions with Nonlinear Buoyancy and Incident Wave Forces

15

Predicted Tz Heave (s)

Table 10: Observed and Predicted Pitch ZeroCrossing Period (s)


Run Trial Strip
3D freq.
3D time
theory Appendages
Hull forces
Sep. Integ. Linear Nonlin.
203 6.3
7.3
7.2
7.2
7.1
7.1
204 7.9
10.0 9.8
9.8
10.1
10.1
206 12.1 11.6 11.4 11.4
11.1
11.0
209 5.8
6.2
6.2
6.2
6.2
6.2
210 10.1 13.6 13.4 13.4
14.5
14.6
303 7.3
7.8
7.8
7.8
7.5
7.5
304 10.8 12.5 12.6 12.6
12.8
12.8
305 7.3
7.0
6.9
6.9
6.9
6.8
306 11.4 10.8 10.4 10.4
10.3
10.5
309 6.7
6.3
6.3
6.3
6.1
6.2
310 10.3 14.1 13.7 13.7
13.9
13.9
403 7.7
7.5
7.5
7.5
7.9
7.9
404 9.2
11.6 11.4 11.4
11.6
11.7
409 6.5
6.0
5.9
5.9
5.9
5.9
410 10.2 12.6 11.9 11.9
12.2
12.2
413 8.0
7.3
7.1
7.1
7.1
7.1
Predicted/observed
Mean
1.10 1.08 1.08
1.09
1.09
Deviation 0.16 0.15 0.15
0.17
0.17

10

....
.....
.....
....
.
.
.
.
.....
.....
.....
.....
.....
.
.
.
....
.....
.....
.....
.....
.
.
.
.
.....
....
.....
.....
.....
.
.
.
.
....
.....
.....
.....
.....
.
.
.
.
.....
.....
.....
.....
....
.
.
.
.
.....
.....
.....
.....
.....
.
.
.
....
.....
.....
.....
.....
.
.
.
.
....
.....
.....
.....
.....
.
.
.
.
.....
.....
.....
.....
....
.
.
.
.
....
.....
.....
.....
....

++

++ +
+ +

Relative heading
+ Bow quartering
Beam
Stern quartering

10

15

Observed Tz Heave (s)

Figure 10: Predicted Versus Observed Heave ZeroCrossing Period, Time Domain Predictions with
Nonlinear Buoyancy and Incident Wave Forces

10

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

....
.....
.....
.....
.....
.
.
.
....
.....
.....
.....
.....
.
.
.
.
.....
....
.....
.....
.....
.
.
.
.
.....
.....
.....
.....
....
.
.
.
.
.....
.....
.....
.....
.....
.
.
.
....
.....
.....
.....
.....
.
.
.
.
.....
....
.....
.....
.....
.
.
.
.
....
.....
.....
.....
.....
.
.
.
.
.....
.....
.....
.....
....
.
.
.
.
.....
.....
.....
.....
.....
.
.
.
....
.....
.....
....

++
+

Relative heading
+ Bow quartering
Beam
Stern quartering

+
+

+ +
....
.....
.....
.....
.
.
.
.
+
+ ..........+
+ ...............

Predicted RMS Pitch (deg)

Predicted RMS Roll (deg)

....
.....
.....
....
.
.
.
.
.....
.....
.....
.....
.....
.
.
.
....
.....
.....
.....
.....
.
.
.
.
.....
....
.....
.....
.....
.
.
.
.
....
.....
.....
.....
.....
.
.
.
.
.....
.....
.....
.....
....
.
.
.
.
.....
.....
.....
.....
.....
.
.
.
....
.....
.....
.....
.....
.
.
.
.
....
.....
.....
.....
.....
.
.
.
.
.....
.....
.....
.....
....
.
.
.
.
....
.....
.....
.....
....

15

++
+++
+
+

Relative heading
+ Bow quartering
Beam
Stern quartering

10

...
.....
.....
.....
....
.
.
.
.
.....
.....
.....
.....
.....
.
.
.
....
.....
.....
.....
.....
.
.
.
.
.....
....
.....
.....
.....
.
.
.
.
.....
.....
.....
.....
....
.
.
.
.
.....
.....
.....
.....
.....
.
.
.
....
.....
.....
.....
.....
.
.
.
.
.....
....
.....
.....
.....
.
.
.
.
....
.....
.....
.....
.....
.
.
.
.
.....
.....
.....
.....
....
.
.
.
.
....
.....
.....
.....

Predicted Tz Pitch (s)

Predicted Tz Roll (s)

Figure 13: Predicted Versus Observed RMS Pitch,


Time Domain Predictions with Nonlinear Buoyancy and Incident Wave Forces

Observed RMS Pitch (m)

Figure 11: Predicted Versus Observed RMS Roll,


Time Domain Predictions with Nonlinear Buoyancy and Incident Wave Forces

10

Relative heading
+ Bow quartering
Beam
Stern quartering

Observed RMS Roll (m)

15

.
.....
.....
.....
.....
.
.
.
....
.....
.....
.....
.....
.
.
.
.
.....
....
.....
.....
.....
.
.
.
.
....
.....
.....
.....
.....
.
.
.
.
.....
.....
.....
.....
....
.
.
.
.
.....
.....
.....
.....
.....
.
.
.
....
.....
.....
.....
.....
.
.
.
.
....
.....
.....
.....
.....
.
.
.
.
....
.....
.....
.....
....
.....

10

Observed Tz Roll (s)

+
+ +
+
+
++

15

Relative heading
+ Bow quartering
Beam
Stern quartering

10

15

Observed Tz Pitch (s)

Figure 12: Predicted Versus Observed Roll ZeroCrossing Period, Time Domain Predictions with
Nonlinear Buoyancy and Incident Wave Forces

Figure 14: Predicted Versus Observed Pitch ZeroCrossing Period, Time Domain Predictions with
Nonlinear Buoyancy and Incident Wave Forces

11

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
The Nipigon sea trial included simultaneous measurements of ship motions and directional wave spectra. Measurements of rudder motions would have
enhanced the value of the data for validation of
lateral plane motion predictions. In the absence of
measured rudder motions, coursekeeping has been
modelled using low frequency stiffness terms, and
associated damping terms, for surge, sway and yaw.

Roll Added Mass A44 /I44

0.12

The numerical predictions give generally


good agreement with the RMS motions and zerocrossing periods during the Nipigon sea trial. As
expected, agreement is better for heave and pitch
than for roll. Among the numerical prediction methods, the predicted motions are very similar. The
good results from strip theory are likely due to
Nipigons slender geometry. Time domain predictions using linear and nonlinear forces from buoyancy and incident waves give very similar results,
likely due to Nipigon having minimal flare near the
waterline at most stations.

0.10
0.08

................................................................................................................................................................
...................................................................................... ....... ....... ....... ....... ..... ..... ...... ....... ....... ....... ....... ....... .

...................................

0.06

.....

0.04

....... ....... ....


.............................

Bare hull
Panelled appendages with hull
Appendages separate from hull

0.02
0.00

10

20

30

Ship Speed U (knots)

Figure 15: Nipigon Roll Added Mass Versus Ship


Speed at Roll Natural Period of 10.6 s, Zero Roll
Amplitude

cr
Roll Damping B44 /B44

The negligible influence of panelled appendages on the roll radiation damping is likely due to
the small appendage sizes relative to the hull and
to the degree of submergence of the appendages.
Larger appendages located closer to the free surface would have a larger influence on roll radiation
damping. For Nipigon and other ships with small
appendages of large submergence, it appears unnecessary to include panelled appendages in radiation
and diffraction computations.
The scattergrams in Figures 9 to 14 provide further insight into trends. The overprediction
of RMS roll appears to be similar for bow quartering and stern quartering seas. The underprediction
of viscous appendage forces is a possible cause for
overprediction of RMS roll. For RMS pitch, there
is a trend toward overprediction in bow quartering
seas, which could be due to higher amplitude motions. Pitch zero-crossing periods exhibit a trend
toward overprediction in stern quartering seas, possibly due to the assumption of high encounter frequency when considering forward speed effects.

0.12

.....

0.10

....... ....... ....


.............................

Bare hull
Panelled appendages with hull
Appendages separate from hull

..
........
........
.
.
.
.
.
.
......
.........
........
0.06
.
.
.
.
.
.
.
.......
........
........
.
.
.
.
.
.
.....
0.04
........
........
.
.
.
.
.
.
.
.....
.........
0.02
.........
.
.
.
.
.
.
.
.
.
......
............
........ . . . . . . . . . . . . . . . . .
.
.
.
.
.
.
.
..... . . .

0.08

0.00

10

20

30

Ship Speed U (knots)

Figure 16: Nipigon Roll Damping Versus Ship


Speed at Roll Natural Period of 10.6 s, Zero Roll
Amplitude

Accurate prediction of viscous roll damping forces is still considered to be a major challenge
for obtaining accurate ship motion predictions. The
current motion predictions using Katos method
(1966) for bilge keel roll damping give reasonable
roll results. However, the roll motions presented
12

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

from the Nipigon trial are limited to relatively narrow ranges of RMS values (2.63 to 6.16 degrees) and
zero-crossing periods (8.4 to 13.2 s). It is essential
that roll motions be validated over wide ranges of
motion amplitudes and frequencies.

Lloyd, A.R.J.M., Seakeeping: Ship Behaviour in


Rough Weather, Ellis Horwood, Chichester, England, 1989.
McTaggart, K., Lateral Ship Motions and Sea Loads
in Waves Including Appendage and Viscous Forces,
International Shipbuilding Progress, Vol. 47, 2001,
pp. 141160.

CONCLUSIONS

McTaggart, K., Datta, I., Stirling, A., Gibson, S.,


and Glen, I., Motions and Loads of a Hydroelastic
Frigate Model in Severe Seas, Transactions, Society of Naval Architects and Marine Engineers, Vol.
105, 1997.

Numerical predictions have been compared with


measured heave, roll, and pitch for a naval destroyer
in multidirectional seas. The numerical predictions
give generally good agreement with measured RMS
motions and zero-crossing periods. Strip theory
and three-dimensional predictions give very similar results, likely because of the slender hull form
of the destroyer HMCS Nipigon. The influence of
appendage radiation forces has been investigated
by including panelled appendages in radiation computations. For Nipigon, inclusion of appendages in
radiation computations has negligible effect on predicted motions in comparison with the commonly
used approach of treating appendages separately
from the hull and assuming deep submergence. Evaluation of nonlinear buoyancy and incident wave
forces based on the instantaneous wetted hull surface has only a minor influence on predicted motions relative to those assuming linear hull forces.

Meyerhoff, W.K., Added Mass of Thin Rectangular Plates Calculated from Potential Theory,
Journal of Ship Research, Vol. 4, 1970, pp. 100
111.
Papanikolaou, A.D. and Schellin, T.E., A Three
Dimensional Panel Method for Motions and Loads
of Ships with Forward Speed, Schiffstechnik (Ship
Technology Research), Vol. 39, 1992, pp. 147156.
Schellin, T.E., Chen, X.-B., Beiersdorf, C. and Maron,
A., Comparative Frequency Domain Seakeeping
Analysis of a Fast Monohull in Regular Head Waves,
21st International Conference on Offshore Mechanics and Arctic Engineering, Oslo, 2002.
Schmitke, R.T., Ship Sway, Roll, and Yaw Motions in Oblique Seas, Transactions, Society of
Naval Architects and Marine Engineers, Vol. 86,
1978, pp. 2646.

REFERENCES
Ballard, E.J., Hudson, D.A., Price, W.G. and Temarel, P., Time Domain Simulation of Symmetric
Ship Motions in Waves, International Journal of
Maritime Engineering, Vol. 143, No. A3, 2003, pp.
120.

Wheeler, D.J, Method for Prediction of Forces


Produced by Irregular Waves, Journal of Petroleum Technology, 1970, pp. 359367.

Beck, R.F. and Loken, A.E., ThreeDimensional


Effects in Ship RelativeMotion Problems, Journal of Ship Research, Vol. 33, 1989, pp. 261268.
Chakrabarti, S.K., Hydrodynamics of Offshore Structures, Springer-Verlag, 1987.
Himeno, Y., Prediction of Ship Roll Damping State of the Art, Report 239, 1981, Department of
Naval Architecture and Marine Engineering, University of Michigan.
Kato, H., Effects of Bilge Keels on the Roll Damping of Ships, Memories of the Defence Academy,
Japan, Vol. 4, 1966.

13

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Larry J. Doctors
The University of New South Wales, Australia
I would like to express my gratitude for this
very interesting paper and the presentation.
Could you kindly clarify the justification for
considering the inclusion of the nonlinear buoyancy
forces, while still assuming that the hydrodynamic
forces are linear? One could argue that this approach
is inconsistent.
AUTHORS REPLY
We note that when including nonlinear
buoyancy forces we simultaneously consider
nonlinear incident wave forces. The buoyancy and
incident wave forces are typically the largest force
components acting on the ship, and are also relatively
easy to compute when including nonlinearities. In
contrast, the added mass, retardation, and diffraction
wave force components are smaller in magnitude and
much more difficult to compute when including
nonlinearities.
Our validation work to date has been limited
to moderate sea conditions, for which nonlinear
forces will have a small influence on predicted
motions. We plan to do future validation work using
model test data for severe conditions, which will
hopefully indicate whether our current treatment of
nonlinear forces is adequate.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Arthur M. Reed
Naval Surface Warfare Center, Carderock Division,
USA
How are forward speed corrections
incorporated with the zero speed Green function to
obtain the hydrodynamic coefficients at forward
speed?
AUTHORS REPLY
Our treatment of forward speed effects is
similar to that used by Beck and Loken (1989) and
Papanikolaou and Schellin (1992). The influence of
forward speed on the body boundary condition is
treated in an approximate manner. The influence of
forward speed on the free surface boundary condition
is assumed to be negligible. We are pleased with the
accuracy and robustness of the approach for moderate
ship speeds. For higher ship speeds, we recognize
the requirement for a more accurate treatment of the
influence of ship speed on boundary conditions.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Allen Engle
Naval Surface Warfare Center, Carderock Division,
USA
The authors should be congratulated for
their efforts. The rigorous collection of full scale
trials data and correlation with numerical simulations
is an arduous task that is not often undertaken.
However, as the authors point out, such data is
extremely valuable, particularly when conflicting
scaling laws (i.e., Reynolds and Froude).come into
play.
Im very glad to see that the authors were
able to use an Endeco wave buoy as part of the wave
measurement instrumentation package. It is crucial
that the effects of wave spreading be established as
part of any correlation effort.
The comparisons with simulations are quite
acceptable and its good to see that even simplified
methods such as strip theory can provide an analyst
with a fair degree of insight to a ships seakeeping.
That the best results are obtained for heave and pitch
are not surprising, as vertical plane motions are
governed by potential flow wave damping. However,
this is not the case with roll motion as viscous effects
will dominate. Given the high degree uncertainty
related to predicting roll motions, have the authors
considered strain gaging the ships bilge keels and
other control surfaces? Such a force decomposition
would greatly assist in developing better roll damping
algorithms.

AUTHORS RESPONSE
We thank Mr. Engle for his complimentary
remarks, and for his recognition of the value of fullscale trials.
Prior to Mr. Engles remarks we had not
considered strain-gauging bilge keels and other
control surfaces. We concur that measuring forces on
appendages could provide very useful insight,
justifying the effort required to install the required
instrumentation. When preparing for trials measuring
sea loads, such as that conducted on HMCS Nipigon,
we install strain gages and pressure transducers on
the hull while the ship is in dry dock. Under such
circumstances, instrumentation of appendages could
be performed with an acceptable degree of
incremental effort.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Jinzhu Xia
Australian Maritime College, Australia
I would like to comment on the authors
discussion on the effects of nonlinearity in
hydrostatic and slamming forces on vertical motions.
This would depend on hull shapes. For S175
Containerships, for example, nonlinearity in
hydrostatic force may have a significant influence on
heave and pitch motions. The influence of the
slamming force on motions could also be notable, but
would have a much more significant role in
determining the sagging and hogging bending
moment. (Wang, Xia and Jensen, 2000, ONR; Xia
and Wang, 1997, Journal of Ship Research)
AUTHORS REPLY
We concur that nonlinear hydrostatic and
slamming forces will depend on hull geometry, and
will increase as the geometry in the vicinity of the
instantaneous waterline deviates from being wallsided. We also agree that bending moments tend to
be more sensitive than motions to nonlinear forces.
In the next stages of development for the ShipMo3D
library, we plan to incorporate nonlinear sea loads
and validate predictions using available model test
data for S175 container ships

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics


St. John's, Newfoundland and Labrador, Canada, 8-13 August 2004

A Single-Phase Level Set Method with Application to


Breaking Waves and Forward Speed Diffraction Problem
Robert Wilson1, Pablo Carrica1, Mark Hyman2, and Fred Stern1
(1University of Iowa, USA,
2
Naval Surface Warfare Center, Coastal Systems Station, USA)

ABSTRACT
A steady and unsteady single-phase level set
method is developed for the RANS code CFDSHIPIOWA to simulate free surface flows around complex
geometry with large amplitude motions and
maneuvering. A structured overset grid approach is
used to allow flexibility in grid generation, local mesh
refinement, and to efficiently resolve incident waves.
The method is demonstrated by simulating the flow
around a surface combatant in calm water at three
Froude numbers using embedded overset grids for local
refinement of the overturning bow wave. Comparison
of numerical results with measurements shows that the
method is able to accurately simulate the overturning
bow wave, resulting free-surface vortices, and
boundary layer. In addition, the simulation is used to
fill-in the sparse experimental dataset and to help
explain the flow pattern induced by the overturning
bow wave. Application to a bluff geometry with
overturning waves is provided by simulating a landing
craft in calm water. The method is also applied to the
simulation of a surface combatant advancing in incident
waves. Comparisons of the unsteady wave field and
axial velocity at the nominal wake plane with
experimental measurements show the ability to
accurately and efficiently predict unsteady free surface
flows around practical geometries. The results show
the capability of the method for resolving complex free
surface topologies associated with high Froude number,
bluff geometry, and incident waves and the promise for
application to large amplitude motions and
maneuvering. Detailed verification, validation, and
application to other test cases are needed to fully
establish the capability of the method.
1 INTRODUCTION
Ship hydrodynamics presents many unique
challenges due to complex geometry, environment, and
operating conditions, which results in many complex
physics and modeling issues. Operating conditions

with high Froude number, bluff geometry, and/or


large amplitude ship motions and maneuvers can lead
to steep, overturning, spilling, and breaking waves
around the ship which can produce spray, foam, and
bubbles; affecting the performance of the hull and
propulsion system and increasing air and water
signatures.
Of the many modeling issues encountered, the
air/water interface continues to be one of the most
difficult areas to deal with computationally, due to
complex free surface topology (e.g., breaking waves,
spray), unsteadiness due to ship motions or incident
waves, large differences in fluid properties between
water and air, requirement for sufficient resolution of
both the turbulent boundary layer and waves, and
stability considerations. In addition, problems with
complex free surface topologies can be modeled and
resolved at a range of scales (e.g., modeling of the
effect of the breaking wave on a smoothly resolved
free surface or resolution of a spilling breaking wave
and modeling of the two-phase flow).
Two approaches for modeling of the air/water
interface have traditionally been followed: (i) surface
tracking, where a 2D kinematic condition is enforced
at the free surface and the grid is dynamically
conformed to the free surface and ship geometry (i.e.,
the domain covers the water region only) and (ii)
surface capturing where the grid is fixed throughout
the simulation and the transport of a scalar function is
used to capture the interface. The first approach has
the advantage that the free surface interface remains
sharp and the grid can be easily clustered around the
interface to resolve large gradients. However, the
surface tracking approach is limited to flows with
small to medium wave slopes and cannot resolve
steep, overturning, or breaking waves due to
difficulties in the grid conforming process.
The second approach for free surface
modeling does not suffer these limitations since the
computational domain is fixed and the interface is

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

captured between the air and water regions. An early


example of a free surface capturing approach was
provided by Harlow and Welch (1965) using massless
particles to mark and follow the interface region. Later,
the volume of fluid (VOF) method was proposed by
Hirt and Nichols (1981), where the fluid in both the
liquid and gas region is marked and followed by the
convection of a step function (i.e., the function is unity
in water and zero in air), which represents the volume
fraction of liquid within each cell. The main advantage
of this method is robustness, though the implementation
of the interface detection algorithm is difficult and is
still a matter of research, see for instance Kim and Lee
(2003).

trimaran).
Unstructured grids offer increased
flexibility and reduced user input for discretizing
complex ship geometries. Unstructured flow solvers
have been used to compute free surface flows around
ship models. Hino et al. (2003) simulated the steady
flow around a cargo ship with several configurations
of podded propulsors, although the free surface was
not taken into account; Lohner and Yang (1998) used
a surface tracking approach for the Wigley hull; and
Burg et al., (2002) used a surface tracking approach
to simulate a fully appended surface combatant. The
above mentioned unstructured simulations were all at
medium Froude number so that the free surface was
smooth without steep or breaking waves.

More recently, level set methods were proposed


by Osher and Sethian (1988) and subsequently applied
to a wide range of interfacial problems including
gas/liquid fluid flows, solid/liquid phase changes in
materials processing, and flame front propagation in
combustion (for a review see Sethian and Smereka,
2003). With this method, the zero contour of a 3D
level set function is used to define the location of the
interface, while the function itself represents a signed
distance to the interface (positive in liquid and negative
in gas). With proper discretization, the level set
function has a constant gradient unit magnitude across
the interface and is therefore less prone to numerical
diffusion during transport.

An alternative to unstructured grids is the use


of structured overset grids. Much of the work in this
area was originally developed by the aerospace field
and has been used to simulate flow around bodies
with relative motion, re-entry vehicles, subsonic
vehicles, rotorcraft vehicles, and turbomachinery
(Chan et al., 2002 and Suhs et al., 2002). With this
method, complex configurations can be discretized
using a combination of body-fitted structured grids
embedded inside topologically simple background
grids, which extend to the far-field. Grid points
inside the surface geometry are rendered inactive,
while interpolation is used to transfer information
between overlapping grids using pre-processing
software. Application of overset grid techniques to
the computation of free surface flows around ships
was demonstrated by Orihara and Miyata (2003)
where the pitch and heave motions for two practical
hull forms were predicted using a single-phase
density function surface capturing approach.
Computation of the fluid flow around the stern region
of three practical ship models was also performed by
Regnstrom et al. (2000) using overset grids, although
the free surface was not computed but replaced with
a symmetry plane.

There have been many applications of surface


tracking and capturing approaches to ship
hydrodynamics. The surface tracking approach has
been successfully applied to predict resistance and
propulsion where ships advance with constant speed in
calm water and are restrained from motions (e.g.,
Wilson et al., 2000); the forward speed diffraction
problem where ships advance with constant speed, are
restrained from motions, and encounter regular head
waves (Rhee and Stern, 2001); prediction of roll decay
motion for a surface combatant (Wilson and Stern,
2002); and pitch and heave motions for the Wigley hull
(Weymouth et al., 2003). In Di Mascio et al. (2003),
the single-phase level set method was used to simulate
the steady flow around a cargo ship and surface
combatant using structured multi-block grids. The twophase level set method was used with structured grids
to compute the steady flow around a container ship
(Cura Hochbaum and Vogt 2000), and to simulate
unsteady flow for a container ship with incident waves
(Cura Hochbaum and Vogt 2002).
Surface ships typically have complex geometry
due to appendages (e.g., shafts, propellers, rudders,
bilge keels, water pumps) and use of novel hull design
to achieve stealth requirements (e.g., tumblehome,

The present paper presents the formulation


and implementation of a steady and unsteady level
set method into the RANS solver CFDSHIP-IOWA
to manage both complex ship geometry and complex
interfacial topology. Overset grids are used to
provide flexibility in grid generation and local grid
refinement for free surface flows.
The paper
provides an overview of the mathematical model and
numerical method with a focus on the free surface
model and modifications necessary for the use of
overset grids. The capability of the method is
demonstrated by application to practical ship hulls in
calm water. Analysis and validation of the results are
focused on the simulation of the surface combatant at
Fr=0.35 with overturning bow wave since a large

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

experimental dataset is available for comparison


(Olivieri et al., 2002; Olivieri et al., 2004). Although
the structure of non-linear free-surface waves has been
studied previously (e.g., Miyata and Inui, 1984; Dong
et al., 1997) a complete understanding of breaking bow
waves and the resulting vortical flow is not yet
available.
Capability of the unsteady formulation is
demonstrated by simulating a surface combatant with
incident waves. A summary of those results are
provided here. A complete description of the unsteady
method is provided by Carrica et al., (2004), including
the verification and validation for analytical and
numerical benchmark problems.
2 MATHEMATICAL MODELING
The general-purpose RANS solver, CFDSHIPIOWA, has been developed at the University of Iowa
for ship hydrodynamics and high-performance
computing platforms and has been used for thesis and
project research as well as transitioned to academic,
government, and industry researchers. Documentation
of the basic method (version 3.03) is provided in
Paterson et al. (2003), while extensions to the
mathematical and numerical formulations for the
single-phase level set method, advanced iterative
solver, conservative formulations, and modification of
the overset grid approach for free surface flows
(version 4.0) are presented in this paper.
We consider the flow of two incompressible
fluids (air and water) that can adopt an arbitrary
interfacial topology. We start with the equations for the
standard level set method for incompressible fluids
(Sussman et al. 1994), and appropriate simplifications
will be introduced and justified where necessary to
derive the single-phase level set method.
2.1 Equations of motion
The local, instantaneous equations of motion
of a two-phase incompressible flow can be written as
(Drew and Passman, 1998):
uk ,i
t

+ uk , j

uk ,i
x j

1 pk
+
( 2 k Dk ,ij ) + gi (1)
k xi x j

uk , j
x j

=0

(2)

where the subscript k = w or a indicates the phase


present at a given point in space, either water or air. ui ,
p and Dij are the velocity, pressure and rate of
deformation tensor, k and k are the density and

viscosity of fluid k, and gi is the gravity


acceleration. At the interface, the interfacial
boundary or jump conditions apply. Neglecting
surface tension, we can write the following for
immiscible fluids:

p ij + 2 Dij n j = 0

(3)

where the bracket means w a and the normal n j is


taken from the liquid to the gas. Although neglected
here, surface tension can be included in the model
(e.g., Di Mascio et al., 2003).
The jump conditions (3) can be integrated
into the equations of motion (1) and (2), resulting in
a body force concentrated in the interface of a single
incompressible fluid with variable properties
(Brackbill et al. 1992):

ui
u
1 p

+uj i =
+
( 2 Dij ) + gi
t
x j
xi x j

(4)

The fluid properties are now a function of the spatial


location, in general constant within each fluid.
The location of the interface is given by the
zero level set of the function , known as the level
set function that is positive in water and negative in
air. Since the free surface is a material interface (in
absence of interfacial mass transfer such as
evaporation or condensation), then the equation for
the level set function is:

+ uj
=0
t
x j

(5)

Since is a distance function, the gradient


of the level set function points normal to the interface
into the water. The water-to-air normal can then be
computed as:
ni =

xi
xi

(6)

In the standard level set method, the fluid


properties, which are discontinuous across the
air/water interface, are made continuous through
artificial smoothing (Sussman et al. 1994). Then the
resulting field equations (continuous for the whole
domain) are solved with an appropriate numerical
method. A second approach is to maintain the
discontinuity across the interface and enforce the
jump conditions, Eq. (3), directly. This is known as
the Ghost Fluid method (Fedkiw et al. 1999). For
ship hydrodynamics applications, the large density
and viscosity differences in air and water allow us to

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

make important simplifying assumptions that lead to


the single-phase level set method, as discussed below
and on the following sections.
After proper averaging of the momentum and
mass conservation equations for the water phase, Eqs.
(1) and (2) with k = w, we obtain the Reynolds
Averaged Navier-Stokes equations:
u j

=0

(7)

ui
u

1 p
+ uj i =
+
t
x j
xi x j

ui

ui u j + si (8)

x j

x j

where ui u j is the Reynolds stress tensor and si is


any volume source other than the gravity acceleration.
The turbulence quantities are computed using a
blended k / k model of turbulence (Menter
1994). In this two-equation model, the Reynolds stress
tensor is modeled in the usual fashion as:
u u j 2
ui u j = t i +
k ij
x
j xi 3

(9)

where k = ui2 / 2 is the turbulent kinetic energy and t


is the turbulent kinetic energy. Substituting Eq. (9) into
the momentum conservation equation, Eq. (8), and
using dimensionless variables we obtain:
ui
u
p

+ uj i =
+
t
x j
xi x j

1 ui

+ si (10)
Reeff x j

where for convenience we maintain the same


nomenclature as in Eq. (8) though from this point on all
the variables are dimensionless unless explicitly stated.
The effective Reynolds number is defined as
Re = U 0 L / , with U0 a characteristic velocity (freestream velocity) and L a characteristic length (ship
length). The dimensionless piezometric pressure is
defined as p = pabs / U 02 + z / Fr 2 + 2k / 3 with pabs the
absolute pressure, and the Froude number,
Fr = U o / g L . The Reynolds stresses are related to
the mean rate of strain through an isotropic eddy
viscosity t, which is calculated using Menters
blended k-/k- model without the optional shear-stress
transport (Mentor, 1994) and without the use of wall
functions. The turbulent kinetic energy k and the
turbulence specific dissipation rate are computed
from transport equations.
2.2 Free surface

For level set methods, the free surface is


identified as the zero level set of the distance
function , obtained after solving Eq. (5). The twophase level set method naturally accounts for the
jump conditions at the free surface since these are
embedded on the formulation. In the single-phase
level set method the jump conditions at the free
surface must be explicitly enforced since we solve
the equations of motion only in water. The jump
condition in any direction tangential to the free
surface (given by the tangent vector ti ) is (Kang et
al. 2000):
ui

u
ti n j + i ni t j = 0

x j
x j

(11)

which, after neglecting shear stress in the air leads to:


ui
nj
x j

=0

(12)

int

The jump condition in the direction normal to


the interface can be written in dimensional form as:

ui
ni n j = 0
pabs 2
x j

(13)

As a good approximation for air/water interfaces, the


pressure on the air is constant and equal to the
atmospheric pressure. Under the stated hypotheses,
the dimensionless piezometric pressure at the
interface is given by:

pint =

zint
Fr 2

(14)

where we neglected the contribution of the turbulent


kinetic energy at the free surface. In order to
transport the level set function, Eq. (5), and also the
velocities and turbulent quantities, we need a velocity
in air near the interface. Eq. (12) provides a suitable
velocity that at the same time satisfies the jump
conditions. Eq. (12) can be solved over the whole
domain in air providing a velocity field in air. We
notice that this velocity does not satisfy the equations
of motion in air but Eq. (12). The velocity so
obtained is called an extension velocity. In practice,
it is necessary to add a diffusion term to the RHS of
Eq. (12) to eliminate shadows, i.e., regions where
the ship hull is curved such that the extension
velocity is not well defined when convecting the
velocity from the interface to the air with the level set
normals. All results are presented using a coefficient
of diffusion of 1x10-4. Through testing of several

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

cases, it was found that this level of diffusion was


sufficient to eliminate shadow regions.
2.3 Reinitialization
We notice that the solution of Eq. (5) should
maintain the location of the interface (the zero level set)
but not necessarily will move the other levels of in a
way that will keep it a distance function. A critical step
in the solution of free surface problems with level set
methods is the reinitialization step. This step is
necessary to maintain as a distance function. This is
very important on single-phase level set methods since
the computation of the normal, Eq. (6), assumes that
is a distance function. The normal is then used
extensively to extend dependent variables in air and to
enforce jump conditions.
In order to keep a distance function we
reinitialize it periodically everywhere except at the
interface by solving:

nj

= sign (0 )
x j

(15)

where 0 is the level set function before reinitializing.


n j points in this case into the fluid to be reinitialized,
and is given by Eq. (6) in air and by the negative of Eq.
(6) in water. Thus (15) is an eikonal equation and
propagates information from the interface outwards.
Notice that Eq. (15) is nonlinear since n j depends on

. The numerical solution of Eq. (15) requires a


special treatment and is discussed in the section on
numerical methods.
3 NUMERICAL DETAILS
We use a structured, body-fitted, nonorthogonal multiblock approach with overset grids that
can cross the interface and ghost cells. The overall
scheme is parallelized using MPI. In the following
sections we discuss details of the numerical methods.
3.1 Coordinate transformation
The continuous governing equations are
transformed from the physical domain in Cartesian
(x,y,z,t) coordinates into the computational domain in
non-orthogonal curvilinear coordinates (, , , ). A
partial transformation is used in which only the
independent variables are transformed, leaving the
velocity components ui in the base coordinates. With
standard transformations, the continuity and momentum
equations are given by:

(b u ) = 0
j

xj
ui 1 ) k
+
b j u j
ui =
k
J

)j k
1 k p 1 bl bl ui

+ si
bi
+
J
k J j J Reeff k

(16)

(17)

notice that the convective term has been written in


fully conservative form, and has contributions from
the transport velocity and the grid velocity (for
moving grids).
3.2 Discretization scheme

We use second-order Euler backward


difference for the time derivatives of all variables,

such that = 1 1.5 n 2 n 1 + 0.5 n 2

for the general variable . The convective terms are


discretized using a second-order upwind method.
Taking a control volume of unit side on the
computational domain, the convective terms for an
arbitrary variable can be written as:
xj
1 )k
b j u j
=

J k
1
( Cd Cu ) + ( Ce Cw ) + ( Cn Cs )
J

(18)

where d, u, e, w, n and s stand for the down (i+1/2),


up (i-1/2), east (j+1/2), west (j-1/2), north (k+1/2)
and south (k-1/2) faces of the control volume. For
example, at the down face we have:
Cd = U% d [ d ( a i +1 + b i + c i 1 )
+ (1 d )( d i + e i +1 + f i + 2 )]

(19)

where the effective contravariant velocity on the


down face is:

)
x j
U% d = b1j u j



i +1/ 2

(20)

and d is defined according to:


d = 1 if

d = 0 if

U% d > 0
U% d < 0

(21)

The convective terms in the other directions


are computed similarly. For example, a second-order
upwind scheme has coefficients a=0, b=1.5, c=-0.5,
d=0, e=1.5, f=-0.5. Notice that in Eq. (20) the

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

contravariant velocity has to be evaluated at a half-cell


location. The transport and grid velocities are evaluated
by linear interpolation between the adjacent grid points.
The metrics are computed directly at the half-cell
location in such a way so as to avoid spurious sources
on the evaluation of the conservative divergence. The
resulting discretization has a 13-point stencil. The
described convective discretization is applied for all the
equations with convective terms, including the
momentum conservation, turbulence, level set and
extension equations.

3.3 Pressure Poisson equation

The viscous terms in Eq. (17), and also on the


turbulence equations are computed with similar
considerations using a second-order central difference
scheme. The diffusive term can be written as:
)j k
1 bl bl ui

=
(22)
J j J Reeff k
Dd Du + De Dw + Dn Ds

The velocity field ui* does not satisfy


continuity, since a pressure obtained from iteration
m-1 was used. Eq. (25) can be rewritten as:

where the diffusive flux in the down direction is:


)
1 bl1blk ui
Dd =
(23)

J J Reeff k
i +1/ 2

where again the metrics and effective Reynolds number


are computed at the half-cell location. Other directions
are computed similarly. Notice that the implementation
of Eq. (23) leads to centered derivatives in the three
directions ( , , ), which results in a 19-point stencil
comprising all the first neighbors to the local grid point
with the exception of the eight corners (i1,j1,k1).
Of the 19 grid points, 7 are aligned with the coordinate
lines, meaning second derivatives with respect to each
of the computational coordinates, but 12 result in
second derivatives with respect to two different
computational coordinates. These cross terms cannot
be built into a set of three penta-diagonal matrices from
an ADI method (i.e., one for each coordinate line) and
thus are sent to the source term.
After discretization as described, the
momentum equations are written in algebraic form as:
Aijk uin

+
nb

Anb uin,nb

1 p n
= Sui bik k
J

The Pressure Implicit Split Operator (PISO)


algorithm (Issa, 1985) is used to couple the
momentum and continuity equations. All the
variables are defined on the grid points. In the first
step, the momentum equation is solved using the last
available pressure to compute the pressure gradient:
Aijk ui* + Anb ui*,nb = Sui
nb

ui = ui

1 k p m 1
bi
J
k

1 k p
bi
JAijk
k

where Aijk and Anb denote the central and neighboring


coefficients of the discretized momentum equations,
respectively. The source term Si contains velocities
from the previous two time steps (n-1) and (n-2) and
the mixed derivative terms.

(26)

where the velocity field ui does satisfy continuity for


the given pressure field p. The pseudo-velocity ui is
defined as:
ui =

1
Aijk

*
Sui Anb ui ,nb

nb

(27)

Taking the divergence of Eq. (26) we obtain:


)j
bi bik p
)
(28)

= j bi j ui
j JA
k
ijk
which is a Poisson equation for the pressure. Eq. (28)
can be expressed as:
jk p
E
=d
j
k

(29)

where E jk and d are defined below. To couple the


pressure and velocity fields, Eq. (28) is evaluated on
a control volume with unit side on the computational
domain, as done with the viscous terms on the
momentum or turbulence equations. The source term
is:

d = U d U u + Ve Vw + Wn Ws
(24)

(25)

(30)

where U i are the contravariant velocities evaluated


at the cell faces. For instance:
)
U d = bl1

ul ,i + ul ,i +1
i +1/ 2

(31)

Similarly, the contravariant pseudo-pressure


gradient is:

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

p = E jk p

j
k

(32)

with

jk

)
bi j bik
=
JAijk

For instance, the contravariant


pressure gradient on the down face is:

(33)
pseudo-

p = E11 p

d
i +1, j , k pi , j , k +
E12
pi , j +1,k pi , j 1,k + pi +1, j +1,k pi +1, j 1,k + (34)
4

13

E
4

At solid walls a special boundary condition


is used to avoid the singularity caused by the no-slip
condition. The interface interaction with the solid
wall of a ship is dominated by surface tension, which
causes the generation of drops and bubbles, allowing
films and drops to slip down the hull, etc.
Simulations of these scales on a ship hydrodynamics
problem would be prohibitive. Since we are not
interested at this point in these phenomena, to avoid
instabilities we blank out the level set function on the
normal direction to the hull, applying to all grid
points that lay within a certain distance to the hull an
extended zero gradient boundary condition as:

( rhull + a n ) = ( rhull + d n )

(a < d )

(36)

( pi, j,k +1 pi, j,k 1 + pi +1, j,k +1 pi +1, j,k 1 )

where rhull is the location of any grid point at the

Similar expressions apply for the other faces.

hull, n is the normal to the hull and a is the distance


from the hull, which has to be smaller than the
blanking out distance d.

p
p +
p
p +
p p = d (35)

d
u
e
w
n
s

3.5 Single-phase level set specifics

Then

is the final expression for the pressure Poisson


equation.
3.4 Initial and boundary conditions

Initial conditions must be provided in order to


start a transient computation. These are problem
dependent and can be input through a restart file. A
simple free-stream condition can be imposed in which
the velocity is (1,0,0), the pressure is zero, the
turbulence quantities are set to the free-stream values
k = 1.107 , = 9.0 , and the free surface is set at z=0.
This results in an impulsive acceleration of the ship,
which in some cases can cause extreme transients that
can take a long time to dissipate. For the cases in which
a steady-state is sought, a slow acceleration from rest
leads to faster convergence.

The implemented boundary conditions are


standard for most of the equations and boundaries,
including solid, domain truncation, and computational
boundaries. These result in Dirichlet or Neumann
conditions for the velocities, pressure, k, and level
set function. Details on most boundary conditions can
be found in Paterson et al. (2003). We will discuss here
some specific boundary conditions that arise from the
single-phase level set method.

The level set function Eq. (5) is a


convection equation that is solved using the secondorder upwind method discussed above. Issues
particular to the single-phase level set method
implementation are discussed in the following
sections.
Reinitialization. The reinitialization procedure is
split into two steps. The first step is a close point
reinitialization for those grid points that are
neighbors to the interface as in the Fast Marching
Method (Sethian 1996). In the second step, we solve
Eq. (15) for the rest of the grid points, using the nearboundary points as Dirichlet boundary conditions.
This is easy to implement in parallel environments
and reasonably inexpensive. We extend the method
of Adalsteinsson & Sethian (1999) to three
dimensional curvilinear grids to obtain a good signed
distance for the first neighbors to an interface (the
beginning set of their close points). In curvilinear
grids with very large aspect ratio (in boundary layer
grids can be as large as 105) we cannot use the
distance to the first neighbor to define the
geometrical distance to the interface. Details of the
close point initialization algorithm are given in
Carrica et al. (2004), and will be briefly discussed
here.

Three types of boundary conditions are used


for the level set function. At Dirchlet boundaries, like
an inlet, the level set function is prescribed, which is
equivalent to prescribing the location of the air/water
interface. Everywhere else, except at solid walls, a zero
gradient boundary condition is used.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

=0

Figure 1 First neighbors to the free surface.


Figure 1 shows a grid with large aspect ratio
where the first neighbors to the interface are marked
with black circles. The algorithm searches for the
closest point along the three coordinate lines in each
direction, including neighbor blocks. Let these
locations be r + , r + , r + , r , r + , r , where we must
note that in some directions there might be no
interfaces. The distance from the point to be
reinitialized to the interface is given by the distance to
the plane formed by the three points:
r = min ( r + , r ) , r = min ( r + , r ) ,
r = min ( r + , r )

(37)

and results in:


( r rp ) ( r rp ) ( r rp )

d=
r

( ) ( r )

(38)

The second step of the reinitialization, the


solution of Eq. (15), is computed using the numerical
methods described before. Notice that the equation is
nonlinear because the normal depends on , thus
requiring a few nonlinear iterations.
Pressure condition. The pressure at the free surface
was derived from the jump conditions and is given by
Eq. (14). Because we are using a surface capturing
approach, the free surface is not located at the grid
points. This means that we need to set the pressure on
the neighbor points to the free surface in such a way
that the interpolated pressure at the free surface is that
determined by Eq. (14). The free surface itself can be
easily identified by locating the change in sign of
between two contiguous grid points along any
coordinate line.
We classify the grid points in three categories.
The grid points in water with all the first neighbors in
water are standard points that can be computed without
any special treatment and need no additional
consideration. The second set of grid points are those

that are in air, and since the pressure in air is not used
in the method it can have any value. We choose to
enforce Eq. (14) for the points in air, where z is now
the vertical coordinate at the grid point. The third set
of points is comprised of grid points in water in
which at least one of the neighbors is in air, and are
the points that we turn our attention to.
For any grid point p in water that has a
neighbor in air na, the interfacial pressure condition
of Eq. (14) is enforced locally. As shown in Figure 2,
interpolating along the gridline we find that the
relative distance between the grid point in water and
the interface is:

p
p na

(39)

Using also linear interpolation along the line


joining the points p and na and using Eq. (14) we
obtain for the interfacial pressure:

pint =

(1 ) z p + zna
Fr 2

(40)

In order to find the pressure at the neighbor


in air that will enforce the correct pressure at the
interface, we have to extrapolate from the known
values of pressure and the pressure at the interface
itself. We set an additional condition to the
extrapolation algorithm, which is that the resulting
pressure stencil will remain the same as that for the
points that are completely in liquid, i.e. a 19-point
stencil. This condition results in keeping the structure
of the pressure matrix simple. Our extrapolation
scheme is a modification of that of Di Mascio et al.
(2003), and has been shown to be stable under all
conditions tested.
The pressure at the interface can be found
from extrapolation of the pressure values at the
points h and at the interface, where h is located
halfway between the local point p and the opposite
neighbor in water to na, shown as nw in Figure 2,

rh = r p + rnw

2 . Thus:

pna = ( pint ph )

dist (rna ,int)


+ pint (41)
dist (rp ,int) + dist (rp , rh )

where the pressure at the mid-point h is computed


from the following average:
ph =

p p + pnw
2

(42)

As discussed before, this scheme does not


modify the connectivity and therefore maintains the

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

original stencil, since all the necessary neighbors to


enforce the interfacial pressure condition are the same
necessary to build the pressure matrix (for typical 19
point stencils). Thus the method can be easily
implemented in existing multi-block codes with no
modifications on the existing inter-block transfer
scheme. This is especially attractive in parallel
implementations.
When computing the pressure gradient,
necessary only in grid points in liquid where Eq. (17) is
solved, neighbors to the free surface will need to use
the interfacial pressure instead of the neighbor pressure,
since the neighbor is located in air and its pressure is
not defined.

Figure 2 Pressure extrapolation into the air.


Total time derivative. The computation of the total
time derivative terms near the interface requires a
special treatment for grid points that change from air to
water during a time step. The available velocity for the
previous time step for the grid point in air is the normal
extension, which does not satisfy the equations of
motion. We use here the method of Carrica et al.
(2004) to compute the total time derivative and readers
are referred to this work for details.
3.6 Overset grid implementation
The CFDSHIP-IOWA flow solver has been
modified to utilize overset grids, giving the code the
capability to simulate flow over complex geometry and
to perform local mesh refinement as documented for
version 3.03 in Paterson et al. (2003). However, since
the interpolation coefficients for overset interblock
communication are computed in a pre-processing step,
blocks with both overset and free surface boundaries or
overset blocks that communicate with free surface
blocks cannot be used with the free surface tracking
approach in v3.03, since the grid is dynamically
conformed at each time step, which changes the
overlapping region. As a result, the interpolation
coefficients are no longer valid. Use of a fixed grid
approach (i.e., the single-phase level set method
proposed here), allows the interpolation coefficients to
be computed once in a pre-processing step and to
remain valid throughout the simulation. Thus, the

limitations on the use of overset grids in the previous


method are removed.
Hole cutting and connectivity information
between overlapping blocks is created in a preprocessing step by Pegasus 5.1: an automated preprocessor code for overset grids, originally
developed at NASA Ames Research Center for
aerospace applications. Once the basic grid system
has been generated (i.e., before the grid is
decomposed into sub-blocks for parallel processing),
the grid and boundary condition file is postprocessed with Pegasus to yield connectivity
information. Three basic steps are performed with
this software: (i) hole cutting where points are
identified as being inside or outside the solid body,
(ii) identification of interpolation points that are
either on the outer boundary of an overset block or
are hole-fringe points, and (iii) identification of
donor cells and associated interpolation stencils used
to update interpolation points.
After the hole cutting and connectivity
information is generated for the basic grid system,
the grid is decomposed into sub-blocks of roughly
equal size to be used in a coarse-grain MPI-based
parallel approach. With information from both the
basic and decomposed grid systems, the file
containing the connectivity information for the basic
grid system is post-processed to generate
connectivity information for the decomposed grid.
This guarantees that both the basic and decomposed
grid systems have identical connectivity information,
which is not always the case if the Pegasus code is
run separately on each grid system. The grid and
connectivity file for the decomposed grid are then
used with other files as inputs to the flow solver.
The interface between Pegasus and the flow
solver was designed to take advantage of many
available features in Pegasus such as a double fringe
layer, which allows the flow solver to retain the same
higher-order five-point 1D stencil at near hole and
near outer boundary points. Level-2 interpolation is
also used so that minimum hole sizes are enlarged in
an effort to match grid size and quality between
overlapping meshes, which increases the accuracy of
intergrid interpolations. Level-2 interpolation also
facilitates the use of local mesh refinement, where
the finer mesh is always selected as the donor mesh
when two or more meshes are available (i.e., the flow
solution is always computed on the mesh with the
finest resolution). Figure 3 shows a detailed view of
the connectivity information between body-fitted
boundary layer and orthogonal background grids at
the midship section. Level-2 interpolation has been

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

used here so that the outer region of the expanding


boundary layer grid is rendered inactive and these
fringe points receive information from the background
grid, which has finer resolution.
Values at interpolation (fringe) points f I are
updated from donor values f D after each ADI iteration
sweep of the transport equations for the velocity [Eqs.
(10) and (12)], turbulence quantities, and level set
function [Eqs. (5) and (15)]
f I IB , JB , KB =

id =1

id

f Did

(43)

where id denotes the interpolation coefficient and


id=1,8 sweeps over the eight donor points which are
associated with the interpolant at the grid point
(IB,JB,KB). Since we have valid physical equations for
these quantities in both air and water, no special
treatment is required when updating interpolation
points from donor points that are either fully in air,
fully in water, or around the air/water interface. The
pressure is computed by solving a pressure Poisson
equation, where a global matrix is assembled for all
blocks, so that intergrid communication between
overlapping blocks can be built directly into the matrix.
At all interpolant points in water far from the interface
(i.e., all donor points are also in water), Eq. (43) is used
to set the matrix entries on the left-hand-side of the
global matrix with the right-hand side vector set to
zero.

are partially in air. Figure 4 shows a 2D example of


a near-interface interpolant point that has two donor
points in air and two in water. Since we do not have
valid pressures for the donor points in air, the link to
each donor point in air is broken in the assembly of
the global matrix and replaced by the interfacial
pressure pint = zint / Fr 2 from Eq. (14) located at the
intersection of the donor grid and interface location
(indicated in Figure 4).
Since the original
interpolation coefficients are based on the donor
points in air, which are not used, the interpolation
coefficients are recomputed based on the donor
points in water and the points at the intersection of
the donor grid and interface location. Equation (43)
is modified to compute the interpolant pressure when
some of the donor parts are in air
pI IB , JB , KB

ndw

*idw pDidw =

idw=1

8 ndw

ida =1

*ida

zint ida
Fr 2

(44)

where ndw and nda indicate the number of donors in


water and air, respectively and * denotes the
updated interpolation coefficients. The left-handside of Eq. (44) is used to set the matrix entries for
the pressure Poisson equation, while the right hand
side is included in the source term since this term is
known before the solution of the matrix.

Figure 4 Computation of the interpolant value


(circle) from near-interface donor points (square).
3.7 Solution Strategy
Figure 3 Intergrid transfer between body-fitted
boundary layer (open symbols) and Cartesian
background (closed symbols) grids. Hole (circle),
fringe (square), and active (no symbol) points.
As discussed in Section 3.3, a matrix is
assembled and the pressure Poisson equation is solved
in the water region only. For near-interface points, the
matrix is modified and appropriate boundary conditions
are set so that the dynamic free surface boundary
condition for pressure, Eq. (14), is enforced. This
procedure must be modified when near-interface
interpolant points are computed from donor points that

The discretization methods discussed before


yield coupled, nonlinear algebraic systems. All the
systems with the exception of the pressure equation
are solved with an ADI approach along the
coordinate lines using pentadiagonal solvers.
Multiblock coupling is achieved transferring
interblock information on each ADI iteration. Since
each grid block belongs to a separate processor in a
parallel implementation, multiblock transfer implies
exchange of information between processors. The
parallel implementation in CFDSHIP-IOWA is
accomplished using MPI.
In addition to the
information exchange between blocks at block-block

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

interfaces, overset grid implementations need exchange


of information between donor and interpolant blocks.
For all equations, again with the exception of the
pressure equation, Overset interpolation information is
exchanged between processors on each ADI iteration.
The pressure equation is built into a sparse pressure
matrix and then solved with the PETSc toolkit (Balay et
al. 2002), a parallel solver of large linear and nonlinear algebraic systems. All the interblock and Overset
information is built into the pressure matrix and thus no
iterations are necessary to enforce the multiblock
conditions.
The overall solution strategy is described below.
At the beginning of the computation the interblock
information is loaded in arrays to speed up the transfer
between processors. Variables are initialized and initial
conditions are set or read. The variables are solved one
at the time and then a global iteration loop couples
them. The turbulence equations are solved first. Then
the level set transport equation, which upon
convergence is reinitialized in two steps. The close
points are reinitialized geometrically first and then all
the other points are reinitialized solving Eq. (15). Since
the equation is nonlinear, a few iterations are needed
here to converge. Then the PISO algorithm is solved. It
comprises the implicit solution of Eq. (25) for the
velocities, and the solution of the pressure equation, Eq.
(35), with the explicit update of the velocity, Eq. (26).
This last step is repeated a few times with no need to
rebuild the matrix (only d changes). The enforcement
of the continuity condition, Eq. (7), which comprises
the solution of the pressure matrix and the explicit
update of the velocity field, is the most expensive
procedure during each global iteration.
All the
extensions in air are built into the system and solved
simultaneously with the equations for water, which
means that every grid point has an equation, either in
air or water, either as active, boundary, or hole point.
If the time step is converged, the convective extension
is solved. When previous time step solutions are loaded
to start the next time step, the solution is loaded in air.
4 EXAMPLES
The single-phase level set method is applied to
three test cases: (i) a surface combatant (also referred to
as Model 5415) in calm water at three Froude numbers,
(ii) a surface combatant encountering incident waves,
and (iii) a landing craft in calm water. The surface
combatant in calm water was one of three test cases at
the Gothenburg 2000 Workshop on Computational
Ship Hydrodynamics, where results were presented
using the surface tracking approach at Fr=0.28 (Wilson
et al., 2000).

Recently, the steady and unsteady singlephase level set approach with overset grids was
implemented into CFDSHIP-IOWA and results were
obtained for the surface combatant for Fr=0.28, 0.35,
and 0.41. Comparisons between surface capturing
and previously obtained surface tracking results show
much improved free-surface modeling, particularly at
the two highest Froude numbers where the surface
tracking method cannot resolve the overturning bow
wave.
Unsteady simulations of the surface
combatant with incident waves were performed due
to the existence of a large experimental database
measured at IIHR. This test case will also be
included in the upcoming Tokyo 2005 Workshop on
Computational Ship Hydrodynamics. Results were
also compared with previously unpublished results
(Wilson and Stern, 2000) using the free surface
tracking approach. The landing craft with bluff bow
geometry is being studied at the Naval Surface
Warfare Center, Coastal Systems Station (CSS) using
the current single-phase method as part of a code
transition effort between IIHR and CSS.
4.1 Surface Combatant in Calm Water
Conditions and Data. Simulations of the surface
combatant in calm water are performed for three
Froude numbers, Fr=0.28, 0.35, and 0.41, which
correspond to Reynolds numbers of Re=1.26x107,
1.57x107, and 1.845x107. The focus of this paper
will be on the Fr=0.35 case since a relatively large
experimental dataset exists for this case, including a
photographic study, free-surface mean and rootmean-square
measurements,
and
velocity
measurements at four cross-sections, x/L=0.15, 0.20,
0.40, and 0.50 (Olivieri et al., 2002; Olivieri et al.,
2004). Although extensive, the experimental dataset
is considered sparse in comparison to information
available from the CFD simulations. Limited results
are presented for the Fr=0.28 simulation since this
case has been studied extensively and does not
contain an overturning bow wave.
Grids. Overlapping grids are used to allow for
flexibility in grid generation and local mesh
refinement for free surface waves. An overlapping
grid is used for Fr=0.28 with an O-O topology grid
from the hull surface to far-field boundary and
overset refinement blocks in the far-field and transom
stern for improved refinement of the Kelvin waves as
shown in Figure 5. Also shown are grids for Fr=0.35
and Fr=0.41, where one and two levels of overset
refinement blocks are used to resolve the overturning
bow waves, respectively. A summary of grid sizes is

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

provided in Table 1. The total sizes of the three grids


are 1.99M, 2.43M, and 3.84M for Fr=0.28, 0.35, and
0.41 simulations, respectively. To resolve the turbulent
boundary layer at the hull surface, grid points are
clustered so that the normalized near wall spacing is
y + < 1 for the highest speed, Fr=0.41. In addition, grid
points are clustered around the calm water plane in the
vertical direction to provide resolution at the freesurface interface.

bow level 1
bow level 2

215x53x91
138x91x121
3.84M total

Resistance. Comparison of CFD predictions and


experimental data for frictional CF, pressure CP, and
total CT resistance is shown in Table 2, where very
good agreement is obtained with differences less than
2.1%.
Table 2. Resistance predictions for surface
combatant in calm water simulations.
Fr

Force

CFD

Data

% diff

0.35

CF
CP
CT
CF
CP
CT

4.79
1.93
2.86
6.53
3.50
3.03

4.84
CR=2.04
ITTC 2.80
6.67
CR=3.94
ITTC 2.73

1.1

0.41

2.1

Free-surface wave field and flow.

Figure 5 Background and chimera refinement grids


near the free-surface plane for Fr=0.28 (top), Fr=0.35
(middle), and Fr=0.41 (bottom).
Table 1. Summary of grids for surface combatant in
calm water simulations.
Fr

Number
Base
Blocks

0.28

Block
Name

Grid
Dimension

background
far-field refine.
transom refine.

201x101x71
191x69x87
138x71x71
1.99M total
211x81x71
209x77x76
2.43M total
201x90x71

0.35

background
bow refine.

0.41

background

Number
Decomp.
Blocks
32

32

38

Figure 6 shows the CFD wave field around the


surface combatant at Fr=0.35 and 0.41, which shows
the wave pattern and the effect of forward speed on
the overturning bow wave. Figure 7 shows a
comparison of free surface contours for the surface
combatant at Fr=0.28, both with and without the farfield and transom refinement blocks. It can be seen
that the addition of the refinement blocks results in
much improved resolution of the Kelvin wave pattern
and closer agreement with experimentally measured
values. Since the background grid rapidly expands
from the hull surface to the far-field, large transverse
grid spacing leads to numerical dissipation of the farfield wave pattern for the grid system without the
refinement blocks. A comparison of wave elevation
contours with measurements at Fr=0.35 is provided
in Figure 8, where the CFD prediction is able to
capture the height of the overturning bow and
shoulder waves. The main differences are that the
bow wave height of the CFD prediction is underpredicted (i.e., the breaking is not as strong as in the
data) and the wavelength is shorter in comparison
with the data. Figure 9 shows the CFD wave
elevation contours for the Fr=0.41 case where it can
be seen that the height and wavelength of the
breaking bow wave is increased. Experimental
measurements are not available for comparison at
this ship speed.
Figure 10 provides a comparison of a
photograph of the bow wave around the ship model
and free-surface perturbation streamlines from the
CFD prediction (i.e., streamlines in an earth-fixed
coordinate system where the forward speed of the
ship is subtracted from the axial velocity). As

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

indicated in the photograph, two scars on the freesurface interface can be observed (i.e., small-scale
depressions in the free-surface, which originate
downstream of the breaking bow wave and are aligned
with the axial direction). The location of the scars
correlate with the location of the converging and
diverging perturbation streamline pattern in the CFD
simulation. The perturbation streamline pattern for the
entire wave field is shown in Figure 11 for the three
ship speeds, where it can be seen that convergence and
divergence of the streamlines is due to the interaction
of the shape of the hull and free-surface topology.
Very close to the hull, the forebody of the ship pushes
fluid laterally towards the far-field, while the afterbody
pulls fluid towards the ship. The figures also show a
direct correlation between the crests and troughs of the
Kelvin wave system and the perturbation streamlines.
Figure 12 shows that the effect of the trough in between
the bow and shoulder waves (0.2<x/L<0.5) is to
generate a favorable pressure gradient normal to the
trough line, which accelerates the axial flow and turns
the flow towards the hull giving the observed
streamline pattern (i.e., axial velocity greater than the
freestream value and negative transverse velocity). It is
this sharp transition region, between the outward flow
caused by the forebody and the inward accelerating
flow at the trough, which generates the diverging freesurface streamlines denoted as scar 1 in Figure 10.
Downstream of the trough line, the flow experiences an
adverse pressure gradient and decelerates, eventually
reversing sign as seen at the shoulder wave if Figure 12
(i.e., the axial and transverse perturbation velocity
become negative and positive, respectively). The
reversal of the perturbation flow in between the trough
and the shoulder wave results in a set of converging
free-surface streamlines (denoted as scar 2 in Figure
10). Similar comparisons at Fr=0.41 (not shown here)
also indicate a correlation between experimentally
observed scars and converging/diverging streamlines in
the CFD simulation. At Fr=0.28, there may be an
indication of scars in the photographs, but they are
more difficult to detect since they are presumably
weaker at this speed.
Even though there is a correlation between the
experimentally observed scars and the perturbation
streamlines, small scale depressions in the free-surface
elevation are not observed in the CFD simulations.
Two possible explanations are that (i) the CFD grid is
very coarse in comparison to the scale of the scars
and/or (ii) there is no surface tension in the CFD
calculations so we do not resolve the capillary waves
on the free-surface. The scars do not show in the mean
free-surface measurements (Figure 8), presumably
because the measurement grid was too coarse.
Additional research is required to determine the

mechanism for the generation of the scars. Future


measurements and simulations may be able to resolve
the scars if the measurement and computational grids
are extremely fine.

Figure 6 Free-surface pattern for the surface


combatant: Fr=0.35 (top) and Fr=0.41 (bottom).

Figure 7 Comparison of free surface contours for


surface combatant in calm water, Fr=0.28.
CFD (top) and data (bottom).

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Figure 8 Free surface contours for the surface


combatant in calm water, Fr=0.35.

Figure 9 Free surface contours for the surface


combatant in calm water, Fr=0.41.

scars
c

Figure 11 Perturbation streamlines for Fr=0.28 (top),


0.35 (middle), and 0.41 (bottom).
Figure 10 Perspective of the bow wave for experiment
(top) and CFD (bottom), Fr=0.35.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Figure 12 Trough detail of free-surface contours and


perturbation streamlines for Fr=0.35.
Boundary Layer and Free-Surface Vortices. Freesurface elevation and axial vorticity contours are shown
in Figure 13 from the Fr=0.35 simulation, where five
free-surface vortices are identified and labeled as V1V5. V1 and V4 are associated with the overturning bow
wave, while V2, V3, and V4 are associated with the
shoulder wave. The location of the vortex cores was
found by tracing the center of closed contours of axial
vorticity from 0.1 < x/L < 0.6.
A comparison of axial velocity and vorticity
contours from experiment and simulation are shown in
Figure 14 and 15 for Fr=0.35. Also shown are CFD
results from the Fr=0.41 case for comparison. The
single-phase level-set method is able to accurately
predict the wake of the low speed fluid at the freesurface downstream of the overturning bow wave
(x/L=0.15 and 0.20). Also, the acceleration of the axial
velocity by the trough in between the bow and shoulder
wave is correctly predicted at x/L=0.40 and 0.50.
Comparisons of the axial vorticity show good
agreement with experimental measurements with
respect to the magnitude, location, and content of the
free-surface vorticies. Analysis of axial vorticity for
the Fr=0.28 non-breaking case shows weaker levels of
axial vorticity and does not show persistence of V1 into
the shoulder wave.
Based on the good agreement of axial velocity
and vorticity, the CFD results can be used to fill-in the
sparse experimental dataset of to understand and
explain the effects of the overturning bow wave.

Figure 13 Under free-surface perspective of wave


contours (background flood) and axial vorticity
(slices) for Fr=0.35.
4.2 Surface Combatant in Regular Waves
The current unsteady level set method was
presented in detail and the accuracy was
demonstrated by Carrica et al. (2004) using three
unsteady test cases: a plane progressive wave,
viscous sloshing in a two-dimensional tank, and the
forward speed diffraction problem for the surface
combatant. The results were compared to analytical
and experimental results and the method was shown
to be both stable and accurate. Results from the
forward speed diffraction problem are summarized
here to demonstrate the capability of the method.
Conditions and Grids. The unsteady simulation for
the surface combatant in waves was performed for
the experimental base case: medium speed (Fr=0.28,
Re=4.65x106), long wavelength (=1.5L), and low
wave steepness (Ak=0.025). An overlapping grid
system is also used for the forward speed diffraction
case to accurately resolve the incident wave and
turbulent boundary layer with a body-fitted grid
around the ship hull and two orthogonal background
grids, (i.e., similar to the grid shown in Figure 5a but
without refinement blocks). The axial grid spacing
for the background blocks is uniform and selected to
resolve the incident wave with 70 points per wave.
The basic grid system is decomposed into 32 blocks
with a total of roughly 2M total grid points.
Approximately 80 time steps per wave period are
used to resolve unsteady flow, which results in a time
step of t =0.00683.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Figure 14 Axial velocity contours for Fr=0.35 at x/L =0.15, 0.20, 0.40, and 0.50:
data (first column) and CFD (second column). CFD for Fr=0.41 (third column).

Figure 15 Axial vorticity contours for Fr=0.35 at x/L =0.15, 0.20, 0.40, and 0.50:
data (first column) and CFD (second column). CFD for Fr=0.41 (third column).

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Results. The unsteady simulation was impulsively


started from an initial condition, which was a
superposition of a uniform free stream and a
perturbation due to the incident wave. As a result, an
initial transient is present in the simulation for the
first three wave periods; followed by an essentially
periodic solution for the next five wave periods.
CFD results are compared to measurements at
quarter-phase intervals for one cycle during the
periodic response. Comparison of free surface
contours in Figure 16 shows that the single-phase
level set method is able to capture the unsteady
Kelvin wave pattern as the incident wave interacts
with the free surface around the ship hull.

Figure 16 Free surface contours. Experimental data


(top) and single-phase level set (bottom).

Figure 17 shows a comparison of CFD


predictions and PIV data of Longo et al. (2000) for a
quarter-phase time sequence of axial velocity
contours at the nominal wake plane. The overall
agreement between the CFD and EFD is quite good,
although the boundary layer thickness is slightly
overpredicted and the cross plane velocities (not
given here) show that the sonar dome vortex is
weaker than the one measured experimentally.

Figure 17 Axial velocity contours at the nominal


wake plane for t/T = 0, 1/4, 1/2 and 3/4.
Experimental data (port) versus single-phase level set
(starboard).

4.3 Landing Craft in Calm Water


The second geometry is a prototype for a
landing craft ship, which has a transom stern and flat
bow that is inclined with respect to the calm water
(see Figure 18). In comparison to the surface

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

combatant, the landing craft has much larger block


coefficient and beam to length ratio. Simulations of
this geometry are being used to investigate the effect
of added wave resistance and the performance of the
hull with various cargo loading configurations. The
current simulations are conducted for the barehull
geometry, but future simulations will investigate the
performance of the hull with water jet propulsors.
Conditions and Grids. Simulations for the landing
craft geometry are performed for Froude and
Reynolds numbers of Fr=0.324 and Re=20x106.
Figure 18 shows the landing craft geometry and
surface grid. The computational grid was generated
by extruding the hull surface grid outward
approximately 10-15%L using a hyperbolic solver.
A far-field grid was generated to join the outer
surface of the boundary layer grid and far-field
boundary using transfinite interpolation. The farfield boundary is hemispherical in shape and is
located 1.5L from the hull surface. The grid system
is decomposed into 12 blocks with a total of 2.56M
grid points.

condition Eq. (12) and do not conserve mass as


discussed in Section 2.2. Physically, at this speed,
we might expect the tip of the overturning wave to
continue to fall and touch the interface in front of the
bow in an unsteady manner. For this simulation, the
grid on the center plane is relatively coarse and the
tip is resolved with only 2-3 grid points so that the
wave does not fully overturn. Figure 21 shows a
detail of the flow and free surface at the transom
center plane. At this speed, the free surface nearly
detaches cleanly from the transom stern and does not
show recirculate.

Figure 19 Landing craft geometry and predicted free


surface, Fr=0.324.

Figure 18 Surface grid for landing craft geometry.


Results. Figure 19 shows a 3D perspective of the
predicted free surface for Fr=0.324. This geometry
results in a stagnation streamline along the flat
inclined bow, which leads to an overturning wave
from the ship centerline to the side of the landing
craft. Figure 20 shows a detail of the overturning
wave at the center plane. Velocity vectors in the
water region implicitly show the separating
streamline, i.e., flow above this streamline moves up
the bow of the ship into the overturning wave and in
the transverse direction towards the side of the ship.
Also shown are limiting streamlines and the free
surface near the hull, which clearly shows the
separating streamline and transverse flow towards the
side of the ship. In addition, there is a strong
downward flow at the side of the ship due to the
termination of the overturning bow wave, which
causes the formation of a pair of counter-rotating
vortices as shown by cross-plane vectors at x/L=0.6.
It should be noted that velocity vectors in air are used
to satisfy the dynamic free surface boundary

Figure 20 Level set contours and velocity vectors at


the bow center plane (top) and at x/L=0.6, view from
bow (bottom) for landing craft geometry.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

other test cases. Future code development efforts


include the addition of motions and maneuvering
capability to the code, dynamic overset interpolation,
and detached eddy simulation models as studied by
Xing et al. (2004). Also, a two-phase level set
approach is concurrently being developed to solve
problems where the air phase becomes pressurized
during the simulation (e.g., formation of bubbles or
entrainment of air) or the stresses caused by the air
are important (e.g., wind-induced stresses),
ACKNOWLEDGEMENTS
Figure 21 Level set contours and velocity vectors at
the stern center plane for landing craft geometry.
5 CONCLUSIONS
A steady and unsteady single-phase level set
method is developed for the RANS code CFDSHIPIOWA to handle both complex ship geometry and
complex interfacial topology due to higher Froude
number, bluff geometry, and/or large amplitude
motions and maneuvering. Overset grids are used to
provide flexibility in grid generation and local grid
refinement for free surface flows.
The method is demonstrated by solving for
the flow around a surface combatant in calm water at
three Froude numbers. Comparison of numerical
results with measurements shows that the method is
able to accurately simulate the free-surface pattern
including the overturning bow wave, resulting freesurface vortices, and modification of the turbulent
boundary layer. In addition, the simulation is used to
fill-in the sparse experimental dataset and to help
explain the flow pattern induced by the overturning
bow wave. Additional focus for this case will be
placed on the free-surface and flow field at the
transom, where an overset refinement block will be
added and an analysis performed. Unfortunately,
experimental measurements of the transom flow are
not available for validation.
Capability to resolve unsteady free surface
flows is demonstrated by simulating a surface
combatant with incident waves with quarter-phase
comparison of the free-surface and boundary layer
with experimental measurements.
Future work will be required to fully
establish the ability of the method including detailed
verification and validation for the surface combatant
in calm water and with incident waves for conditions
that produce non-linear responses, and application to

This research was sponsored by Office of


Naval Research grant N00014-01-1-0073 under the
administration of Dr. Patrick Purtell. The authors
would like to acknowledge the DoD High
Performance Computing Modernization Office.
Simulations were performed at the Army Research
Lab Major Shared Resource Center using both the
SGI Origin 2000 and 3000 machines.
References
Adalsteinsson, D. and Sethian, J. A., The Fast Construction of
Extension Velocities in Level Set Methods, Journal of
Computational Physics, Vol. 148, 1999, pp. 2-22.
Balay, S., Buschelman, K., Gropp, W., Kaushik, D., Knepley, M.,
Curfman, L., Smith, B. and Zhang, H., PETSc User Manual, ANL95/11-Revision 2.1.5, Argonne National Laboratory (2002).
Brackbill, J. U., Kothe, D. B. and Zemach C. A., A Continuum
Method for Modeling Surface Tension, Journal of Computational
Physics, Vol. 100, 1992, pp. 335-354.
Burg C.O.E., Sreenivas, K., Hyams, D., and Mitchell, B.,
Unstructured Nonlinear Free Surface Solutions: Verification and
Validation, Proceedings of the 24th Symposium on Naval
Hydrodynamics, Fukuoka, Japan, July 2002.
Carrica, P. M., Wilson, R. V. and Stern, F., An Unsteady SinglePhase Level Set Method for Viscous Free Surface Flows, submitted,
March 2004.
Chan, W.M., Gomez, R.J., Rogers, S.E., and Buning, P.G., Best
Practices in Overset Grid Generation, AIAA Paper 2002-3191,
Proceedings 32nd AIAA Fluid Dynamics Conference, St. Louis, MO,
2002
Cura Hochbaum, A. and Vogt, M., Flow and Resistance Prediction
for a Container Ship, Proceedings of Gothenburg 2000: A
Workshop on Numerical Ship Hydrodynamics, Chalmers University
of Technology, Gothenburg Sweden, Sept. 2000.
Cura Hochbaum, A. and Vogt, M., Towards the Simulation of
Seakeeping and Maneuvering Based on the Computation of the Free
Surface Viscous Ship Flow, 24th ONR Symposium on Naval
Hydrodynamics, Fukuoka, Japan, 2002.
Di Mascio, A., Broglia, R. and Muscari, R., A Single-Phase Level
set Method for Solving Viscous Free Surface Flows, submitted to
International Journal for Numerical Methods in Fluids, 2003.
Dong, R., Katz, J., and Huang, T., On the Structure of Bow Waves
on a Ship Model, J. Fluid Mech., Vol. 346, 1997, pp. 77-115.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Drew, D. A. and Passman, S. L., Theory of Multicomponent Fluids,


Springer-Verlag, New York 1998; p. 89.
Fedkiw, R. P., Aslam, T., Merriman, B. and Osher, S. J., A nonOscillatory Eulerian Approach to Interfaces in Multimaterial Flows
(the Ghost Fluid Method), Journal of Computational Physics, Vol.
154, 1999, pp. 393-427.
Harlow, F. H. and Welch, J., Numerical Calculation of TimeDependent Viscous Incompressible Flow of Fluid with Free
Surface, The Physics of Fluids, Vol. 8, 1965, pp. 2182-2189.

Paterson, E. G., Wilson, R. V. and Stern, F., General-Purpose


Parallel Unsteady RANS Ship Hydrodynamics Code: CFDShipIowa, IIHR report 432, Iowa Institute of Hydraulic Research, The
University of Iowa, 2003.
Regnstrom, Broberg, and Larsson, L. Ship Stern Flow Calculations
on Overlapping Composite Grids, 23rd Symposium on Naval
Hydrodynamics, Val de Reuil, France, September 2000
Rhee, S. H. and Stern, F., Unsteady RANS Method for Surface Ship
Boundary Layer and Wake and Wave Field, International Journal
for Numerical Methods in Fluids, Vol. 37, 2001, pp. 445-478.

Hino, T., Ohashi, K., and Ukon, Y., Flow Computations Around a
Ship with Appendages by an Unstructured Grid Based NS Solver,
8th Int. Conf. Numerical Ship Hydrodynamics, Busan, S. Korea, Sept.
2003.

Sethian, J. A. and Smereka, P., Level set Methods for Fluid


Interfaces, Annual Review of Fluid Mechanics, Vol. 35, 2003, pp.
341-372.

Hirt, C. W. and Nichols, B. D., Volume of Fluid (VOF) Method for


Dynamics of Free Boundaries, Journal of Computational Physics,
Vol. 39, 1981, pp. 201-221.

Sethian, J. A., Level set Methods: Evolving Interfaces in Geometry,


Fluid Mechanics, Computer Vision and Material Science, Cambridge
University Press, Cambridge, 1996.

Issa, R. I., Solution of the Implicitly Discretized Fluid Flow


Equations by Operator Splitting, Journal of Computational Physics,
Vol. 62, 1985, pp. 40-65.

Suhs, N. E. Rogers, S. E., and Dietz, W. E., PEGASUS 5: An


Automated Pre-processor for Overset-Grid CFD, AIAA Paper 20023186, AIAA Fluid Dynamics Conference, June 2002, St. Louis, MO.

Kang, M., Fedkiw, R. P. and Liu, L. D., A Boundary Condition


Capturing Method for Multiphase Incompressible Flow, Journal of
Scientific Computation, Vol. 15, 2000, pp. 323-360.

Sussman, M., Smereka, P. and Osher, S. J., A Level set Approach to


Computing Solutions to Incompressible Two-Phase Flow, Journal of
Computational Physics, Vol. 114, 1994, pp. 146-159.

Kim, M. S. and Lee, W. I., A New VOF-Based Numerical Scheme


for the Simulation of Fluid Flow with Free Surface. Part I: New Free
Surface-Tracking Algorithm and its Verification, International
Journal for Numerical Methods in Fluids, Vol. 42, 2003, pp 765790.

Weymouth, G., Wilson, R. V. and Stern, F., RANS CFD Predictions


of Pitch and Heave Ship Motions in Head Seas, accepted for
publication in Journal of Ship Research , 2004.

Larsson, L., Stern, F., Bertram, V., editors, Proceedings of the


Gothenburg 2000: A Workshop on Numerical Ship Hydrodynamics,
Chalmers University of Technology, Gothenburg Sweden, Sept.
2000.
Lohner, R. and Yang, C. Viscous Free Surface Hydrodynamics
Using Unstructured Grids, 22nd ONR Symp. on Naval
Hydrodynamics, Washington DC, USA 1998, pp.476-490
Longo, J., Shao, J., Irvine, M., and F. Stern. Phase-Averaged PIV
for Surface Combatant in Regular Head Waves, Proceedings of the
24th Symposium on Naval Hydrodynamics, Fukuoka, Japan, July
2002.
Menter, F. R., Two-Equation Eddy Viscosity Turbulence Models for
Engineering Applications, AIAA Journal, Vol. 32, 1994, pp. 15981605.

Wilson, R., Paterson, E., and Stern, F., 2000, "Verification and
Validation for RANS Simulation of a Naval Combatant,
Proceedings of Gothenburg 2000: A Workshop on Numerical Ship
Hydrodynamics, Chalmers University of Technology, Gothenburg
Sweden, Sept. 2000.
Wilson, R. and Stern. F., Unsteady Viscous Ship Hydrodynamics,
ONR 2000 Workshop on Free Surface Turbulence and Bubbly Flows,
California Institute of Technology, Pasadena, California, March 2000.
Wilson, R. and Stern, F., Unsteady RANS Simulation of a Surface
Combatant with Roll Motion, Proceedings of the 24th Symposium
on Naval Hydrodynamics, Fukuoka, Japan, July 2002.
Xing, T., Kandasamy, M., Wilson, R. and Stern, F., DES and RANS
of Unsteady Free-surface Flows, 42nd AIAA Aerospace Sciences
Meeting, Reno, Nevada, 5-8 Jan 2004, Division for Fluid Dynamics.

Miyata, H. and Inui, T., Nonlinear Ship Waves, Adv. Appl. Mech.,
Vol. 24, 1984, pp. 215-288.
Olivieri, A., Pistani, F., Di Mascio, A., and Penna, R., Breaking
Waves Generated by a Fast Displacement Ship Model, Proceedings
of the 24th Symposium on Naval Hydrodynamics, Fukuoka, Japan,
July 2002.
Olivieri, A., Pistani, F., Campana, E., Benedetti, R., LaGala, F.,
Wilson, R. and Stern, F., IIHR Hydroscience and Engineering
report in preparation, The University of Iowa, 2004.
Orihara, H. and Myata, H., A Numerical Method for Arbitrary Ship
Motions in Arbitrary Wave Conditions using Overlapping Grid
Systems, 8th Int. Conf. Numerical Ship Hydrodynamics, Busan, S.
Korea, Sept. 2003.
Osher, S. J. and Sethian, J. A., Front Propagating with Curvature
Dependent Speed: Algorithms Based on Hamilton-Jacobi
Formulations, Journal of Computational Physics, Vol. 79, 1988, pp.
12-49.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics


St. Johns, Newfoundland and Labrador, CANADA, 8-13 August 2004

A Numerical Study of Nonlinear Diffraction Loads on Floating


Bodies due to Extreme Transient Waves
J. W. Kim1, J. H. Kyoung2, K.J. Bai3 and R. C. Ertekin4
(1American Bureau of Shipping, 2Korea Research Institute of Ships & Ocean
Engineering, 3Seoul National University, 4University of Hawaii at Manoa)
ABSTRACT

extreme waves has been blamed for many losses of


ships and floating structures in the deep ocean (see e.g.,
Faulkner, 2003).
Recently, there have been extensive studies to
identify the generating mechanism of these extreme
waves. There are two different approaches: one is the
focusing of linear wave components to build up wave
crests at one location; extensive experimental and
numerical work has been done along this approach (see,
e.g., Zou & Kim, 1995; Chaplin, 1996, Clauss, 1999;
Smith & Swan, 2002), and the second approach is to
explain the generation of extreme waves by the selfmodulation of the wave train due to nonlinear
interaction. The type-1 instability of Stokes wave
(McClean, 1982), which also is known as BenjaminFeir instability, plays an important role in triggering
the self-modulation. Weakly-nonlinear wave models,
such as the cubic nonlinear Schrdinger equation, has
been applied to describe the nonlinear interaction
involved in the generation and evolution of freak
waves (see e.g., Trulsen & Dysthe, 1997; Osborne,
2001; Grue, 2002). Recently, more sophisticated
theoretical models have been applied to describe the
phenomena more accurately as in Kim & Ertekin
(2000).
In addition to the generating mechanism and
kinematics of the freak waves, one need to consider the
wave loads due to the interaction between floating
structures and freak waves when designing such
structures. The main objective of this study is to
develop an efficient numerical tool to calculate wave
loads due to freak waves impinging on a floating
structure with simple geometry.
Recently, Kim et al. (2004) developed a finiteelement method to solve the diffraction of steep
transient waves by vertical cylinders. The finiteelement method is based on Hamiltons principle and
has been successfully applied to some two-dimensional
problems (Bai & Kim, 1995; Kim & Bai, 1999; Kim et

Diffraction of highly-nonlinear transient waves of


extreme height by floating bodies is numerically
simulated within the scope of potential theory. A finiteelement method based on Hamiltons principle is used
to solve the initial-boundary-value problem. The
nonlinearities related to the free-surface condition and
the fluid domain are fully considered without any
approximation. The specified nonlinear waves are
obtained numerically from the two-dimensional wave
tank simulated by the same finite-element method. A
numerical matching scheme is developed to match the
long-crested waves at the far field and the threedimensional diffracted waves in the near field.
Numerical results for bottom mounted vertical
cylinders are presented.
INTRODUCTION
For the structural design of ships and offshore
structures, various environmental factors during their
service life are considered. Among them, the waveinduced loads, including diffraction pressure and
inertial loads due to motions, are the most important
environmental factors to be considered. The existing
design practice of floating structures have heavily
relied on statistical analysis of the loads under the
assumption that the ocean surface can be represented
by Gaussian random variables, which can be described
by the synthesis of sufficient number of linear wave
components with uniformly-distributed random phases.
However, there have been many observed occasions
where the waves with extreme heights and long crests
arise from the sea conditions that such waves should
not occur if the wave statistics follows the Gaussian
assumptions. With their extreme heights and slopes,
they have been called freak waves, rogue waves,
episodic waves or simply wall of water (see e.g.,
Bascom, 1980; Doneal, 1991). This rare event of

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

al., 2003). In three dimensions, the problem has been


applied to nonlinear wave-resistance problems (see,
Bai et al., 1992, 2003). The same method has been
extended to solve the three-dimensional diffraction
problems by introducing a new radiation condition. A
new numerical scheme was developed to match the
Stokes wave solution at the far field and the diffracted
nonlinear wave in the near field. In this new scheme, a
buffer subdomain was introduced to obtain a gradual
and smooth transition from the near field solution to
the Stokes wave at the far field. A similar technique to
match the nonlinear near-field solution to the linear
far-field solution had been used successfully in the
nonlinear wave resistance problem (see, Bai & Kim,
1992; Bai et al, 1992; Bai & Han, 1994).
In this study, the same numerical scheme has been
applied to solve the diffraction of transient longcrested waves by floating bodies. The transient wave is
generated in a two-dimensional numerical wave flume
either by pulsating a pressure patch or the selfmodulation of a Stokes wave train. Emphasis has been
made to the diffraction of steep waves with
wavelengths much longer than the diameter of the
vertical cylinder. Although the linear theory predicts
negligible wave diffraction, experiments and field
observations on offshore structures have shown
significant design issues such as wave run up and
ringing phenomena at such a set up (e.g., Zou & Kim,
1995; Grue & Husbey, 2002; Chaplin et al., 1997). The
extreme waves can easily evolve into breaking waves
and moreover cause impact loads due to breaking (e.g.,
Chan & Melville, 1989; Zhang et al., 1996). In this
paper, the issue of breaking wave loads is excluded and
we consider the diffraction of non-breaking extreme
waves due to a floating body of simple geometry.
Specifically, the diffraction of steep Stokes waves and
focused waves by vertically-wall-sided floating bodies
and pontoons are presented after an introduction of the
theoretical and numerical aspects of the finite-element
method used.

truncated at the numerical radiation boundary, SR,


which will be specified later when we discuss some
numerical examples.
Fig. 1 shows a typical
subdivision of the computational domain that is used in
the present study.
SR
y

SR

S fR

SR
D

SD
SR

(a) Plan view


z
y

SF

SB

SD

z = (x,y,t)

z=-h

(b) Perspective view


Figure 1 Definition sketches of the computational
domain
Hereafter, we nondimensionalize all physical
variables such that the mass density of the fluid, ,
gravitational acceleration, g, and the characteristic
length can be taken as unity. The characteristic length
is the water depth, h, or the wavelength, , depending
on the particular problem, and this will be indicated
accordingly.

MATHEMATICAL FORMULATION
We consider the free-surface flow of an inviscid
and incompressible fluid, in water of uniform depth, h.
The coordinate system is chosen such that the z-axis
directs against the gravity and the Oxy-plane is the
still-water level unless otherwise stated. The location
of the free surface is denoted by z = (x, y, t ) and the
bottom as z = h . The fluid domain is denoted by D,
and the boundaries of the fluid domain, the free surface,
sea bottom and vertical cylinder, are denoted by SF, SB
and S0, respectively. The lateral boundary of the fluid
domain should be at infinity. However, for
computational purposes, the far-field domain must be

We assume that the fluid motion is irrotational;


therefore the velocity potential, (x, y, z , t ) , can be
defined to describe the fluid motion. The velocity
potential is governed by the Laplace equation in the
fluid domain and is subject to appropriate boundary
conditions on the boundaries. The governing equation
in the fluid domain is therefore given by

2 = 0

in D.

(1)

The normal velocity of the fluid vanishes on the sea


floor:

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

= 0 on z = -h,
z

(x, y )( w ) on SfR (8)


t z x x y y

(2)

and on the cylinder wall, S0, i.e.,

= 0 on S0 ,
n

(3)

where (x, y, , t ) and w w (x, y, w , t ) . The


damping function (x, y ) varies from zero at the inner
boundary of SfR to a constant value at the outer area of
SfR. A more specific shape of the damping function will
be provided later in discussing the numerical examples.

where n is the outward vector normal to the cylinder


surface.
On the free surface, Sf, the kinematic and dynamic
conditions can be written as

FINITE ELEMENT METHOD


Following Bai & Kim (1995), the initialboundary-value problem, given by Eqs. (1) through (9),
is solved by a finite-element method based on
Hamiltons principle, where the Lagranian density of
the mechanical system is given by


on Sf , (4)
=

t z x x y y

1
+ + g ( R0 ) = 0 on Sf . (5)
t 2

1
L = t
2

where z = ( x, y, t ) is the free surface and R0 is the


Bernoulli constant, which can be taken as zero if the
Oxy plane is defined as the still-water level.

dz

1 2
p + q .
2

(10)

(11)

respectively. It can be shown that the stationary


condition of the time integration of the Lagrangian
with respect to the variations of , and lead to the
governing equation (1) and boundary conditions (2),
(3), (8) and (9).
We restrict the geometry of the floating structure
to consist of vertically-wall-sided cylinders. But we
allow the discontinuity of geometry along the vertical
direction, as in the case of a pontoon type structure
shown in Figure 2(a). Since the fluid domain varies
because of the surface elevation, (x, y, t), it is more
convenient to adopt a transformed coordinate system (x,
y, , t) for each layer L, defined in Fig. 2(b), such that

where r x 2 + y 2 . It should be noted that a different


radiation condition should be used if the fluid domain
is partially confined, such as in a wave tank. Since we
cannot solve the problem numerically in an infinite
fluid domain, the radiation condition at infinity is
replaced by numerical matching conditions at the finite
truncated boundary, SR , i.e.,

(x, y, z, t ) = (x, y, , t ), z L , = 1,2,3, (12)


where

(7)

which is valid only when the minimum radial distance,


r, from the cylinder to the radiation boundary, SR, is
sufficiently large because the far-field propagating
wave decays as O 1 / r . To accelerate the decay and
minimize the required distance of the radiation
boundary, artificial damping terms are added to the
free-surface conditions, Eqs. (4) and (5), in a transition
buffer domain, SfR:

p (x, y ) w , q (x, y )( w ) ,

as r , (6)

= w , = w on S R ,

The additional damping terms in the free-surface


conditions are treated as a pressure and mass flux
distribution on the free surface, which are written as

The above problem can be completed by specifying an


initial condition and a radiation condition at infinity on
the horizontal plane. We assume that the free-surface
flow is caused by a known incoming wave solution
pair w and w that satisfies the initial-boundaryvalue problem described by Eqs. (1), (2), (4) and (5)
without in the absence of a structure. Then, at infinity,
the following radiation condition should be satisfied:
1
1
, = w + O
= w + O

r
r

1
= g ( R0 ) (x, y ) w on SfR, (9)
t
2

z+h
, h z < h1 ,
h h1

z + h1
, h1 z < h2 ,
h1 h2

3 .

z + h2
, h2 z ,
+ h2

Copyright National Academy of Sciences. All rights reserved.

(13)

Twenty-Fifth Symposium on Naval Hydrodynamics

f
y

f
x

SF

SD

SB

z=-h

(a)

L3

L1

z=-h2
z=-h1

(b)

2
cos

x ( x ) = 0,

2
cos

(14)

m =1

m (x, y, t ) =

Ne

mi

(t )N i (x, y ),

where

{f

( ), f 2 ( ), f 3 ( ),...,

f M

( )}

is the set of

interpolation functions in the vertical direction, and the


set of the horizontal interpolation functions,
N 1 ( x, y ), N 2 (x, y ),..., N N e (x, y ) , is the two-

Ne

(t)N (x, y)
i

(19)

(20)

The far-field wave solution w can represent any


nonlinear transient wave solution. Three different
incoming wave solutions have been considered here:
the fully-nonlinear Stokes wave, a self-modulated freak
wave, and a freak wave generated by phase focusing.

dimensional finite-element shape functions defined in


D, and M and Ne denote the number of interpolation
functions in the vertical direction and on the horizontal
plane, respectively. The free-surface elevation is also
expanded by the same interpolation function:

(x, y, t ) =

(x x1 )
, x1 < x x 2 ;
2( x 2 x1 )
x 2 < x x3 ;
(x x 4 )
, x3 < x x 4 .
2(x 3 x 4 )

2 ( y y1 )
cos 2( y y ) , y1 < y y 2 ;
2
1

y 2 < y y3 ;
y ( y ) = 0,
2 ( y y 4 )
, y3 < y y 4 .
cos
2( y3 y 4 )

(15)

i =1

where

In the transformed domain, the height of the


domain is unity. We can then separate the depth-wise
and horizontal variations of the velocity potential as

f m ( )m (x, y, t ) ,

(17)

(x, y ) = 0 x (x ) + y ( y ) x ( x ) y ( y ) , (18)

Figure 2 Layering of the computational domain

= sin{(m 2 )}, m 3.

For the damping function (x, y) that appears in


Eqs. (8) and (9), the following expression is used:

z=-h

( x, y , , t ) =

= ; f 2 = 1 ,

For the horizontal interpolation function, we use a


linear interpolation function for a three-node triangular
element. A typical finite-element mesh around the
vertical cylinder is shown in Fig. 3. Once the spatial
discretization has been made by the finite-element
method, the Laplace equation, Eq. (1), reduces to a
system of algebraic equations for the velocity potential
at the internal nodes. An iterative scheme (Jacobi
conjugate gradient method) is used to solve the system
of equations. The matrix involved in the algebraic
equations is highly sparse. Only the non-zero entries of
the sparse matrix is stored. The free-surface conditions
given by Eqs. (4) and (5) lead to time-evolution
equations for the surface elevation and the velocity
potential on the free surface. They are first-order
differential equations and are integrated by the 4th
order Runge-Kutta method.

z = (x,y,t)

L2

(16)

i=1

The interpolation functions for each layer are given by

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

0.3

kA

0.4
0.5

0.6

0.2

0.7
0.8

0.1

0.9

0.0

0
-10

-5

10

0.0

(a)
y

0.3

M=3
M=4
Coke let (1977)

12
10

e-kd = 0.2

F.E.M.
M=2

0.4

14

0.4

kC2/g

0.8

1.2

Figure 4 Celerity-amplitude curve of the Stokes waves


in the Cokelets reference frame.

8
6

A Stokes wave can be defined by the wave height,


H, water depth, h, and wavelength, . Following
Cokelet (1977), the length scale is nondimensionalized
such that the wavelength is 2 in the
nondimensionalized coordinate. The water depth
usually used for the definition of the Stokes wave is
different from the still-water depth h and is defined by

4
2
0
-10

-5

10

(b)
Figure 3 Triangulation of the fluid domain for (a) full
ship and (b) half domain for head sea case. Coordinates
are nondimensionalized by the ship length L.

d=

Q
,
C

(21)

where Q is the mass flux and C the celerity of the


Stokes wave. The celerity, C, of the Stokes wave is
defined in the reference frame where there is no
circulation, or the increase in the velocity potential
over the length of the wave:

DIFFRACTION BY STOKES WAVES


For a floating structure with multiple surface
piercing columns, waves can be trapped between the
columns and cause impact loads on the top structures.
Since the trapping modes take several wave periods to
build up, regular wave trains with moderate heights can
be more damaging compared to freak waves with a
single or a few number of wave crests. In this regard,
the fully-nonlinear Stokes wave can be a good
mathematical model for a regular wave train in a wave
tank test, or a swell in the case of a fully-developed sea.

2
0 u(x, z ) dx = (2, z ) (0, z ) = 0.

(22)

On the other hand, when we are interested in Stokes


waves generated by a wave maker, the reference frame
should be the one where there is no net mass flux at
any vertical surface along the wave train, i.e.,

In the linear wave theory, it is well known that


wave diffraction is insignificant when the wavelength
of incoming wave is much larger than the horizontal
dimension of floating structures. In case of highly
nonlinear Stokes wave, the higher-harmonic contents
of the incoming wave can interact with each column to
cause significant diffraction even when incoming
wavelength is large. The diffracted waves may further
interact with the other columns and incoming waves to
generate trapping mode in certain conditions. For
accurate description of the incoming linear waves, the
Stokes wave solution has been calculated here
numerically by solving the problem by the twodimensional version of the present finite-element
method and then compared with the analytic solution
provided by Cokelet (1977).

0 h u(x, z ) dzdx = 0.

(23)

In the reference frame defined by the relation (22),


there is non-zero mass flux toward the direction of
wave propagation. We denote the mean velocity due to
the mass flux by u , which is defined by
u

1 2
u (x, z ) dzdx.
2d 0 d

(24)

As a result, the Stokes wave solution at the wavemaker reference frame can be obtained by adding
uniform backward flow with flow velocity, - u , to the
Stokes wave solution in the reference frame defined by

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

(22). Then the celerity of the Stokes wave solution in


the new reference frame, C ' , and the water depth, h,
have the following relation with the celerity, C, and
water depth, d, in the old reference frame:

0.5

0.4

0.2
0.3
0.4

0.3

kA

C' = C u ,
u2
,
h = d + R0 +
2

(25)

0.7
0.8

0.1
0.9

0
0

The same finite-element method and Newtons


method that were used for the solitary wave problem
by Bai & Kim (1995) have been used here for Stokes
waves. We first obtain the Stokes wave solution in the
original Cokelets reference frame and then transform
it to the wave-maker reference frame. A nonlinearity
parameter, , is introduced by following Cokelet
(1977):

0.2

0.4

0.8

1.2

(a)
0.5

0.3

(26)

T(g/d)0.5 = constant
T'(g/h)0.5 = constant

0.4

0.2

which varies from 0 for a linear wave to 1.0 at the


breaking limit. In Eq. (26), qcrest and qtrough denote the
fluid particle velocities at the crest and trough of the
Stokes wave, respectively.

12
14

0.1
70

40

30

20

0
0

The Stokes wave solution is obtained for given


values of and kd. The relation between the wave
amplitude H/2=A and the celerity, C, are compared
with the semi-analytic solution of Cokelet (1977) in
Fig. 4. The agreement is very good. Faster convergence
is attained in shallower depth, or for higher values of
e kd .

0.2

0.4

0.6
kC2/g

0.8

1.2

(b)
Figure 5 Celerity-amplitude curve of the Stokes waves
in the two different reference frames along the (a)
constant water depth, and (b) constant wave period.
The Stokes wave input has been used to study the
wave trapping between a three-leg GBS operating in
the North Sea by Kim et al (2004). The numerical
solution agrees very well with the wave tank test by
Swan et al. (1997). Very steep trapped wave crest,
which is more than three times higher than the crest
height of the incoming Stokes wave, has been observed
both in the numerical simulation and model test. Fig. 6
shows the snapshot of the surface profile at the
moment of maximum wave elevation. Clearly seen is
the short diffracted waves with short wavelength
interacting with the incoming Stokes wave, which
cannot be explained by the linear theory. The time
history of the surface elevation measured at the
location of maximum elevation is plotted in Fig. 7
compared with the measured data. Excellent agreement
can be seen.

In Fig. 5, the same relation is plotted along the


constant value of two water depths d and h, and two
wave periods T and T1, which are defined by
2
2
, T1 =
.
kC
kC '

0.6
2
kC /g

kA

2
2
2 = 1 q crest
q trough
/C4,

0.5
0.6

0.2

where R0 is the Bernoulli constant shown in Eq. (5), in


the reference frame where z = -d denote the sea bottom.

T=

0.1

-kd

e = constant
-kh
e = constant

(27)

It can be seen that the difference between the two


reference frames is significant for steep waves in the
intermediate water depth. At the deep water, where
e kd = 0 and there is no mass flux, and at the solitary
wave limit, e kd = 1 , where the wave length is infinite,
the Stokes wave solutions at the two reference frames
become identical.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

8 (c) shows the strongly modulated wave after


evolution of 320 periods. The difference between wave
elevation at wave trough and crest is about 2.5 times of
the initial wave height. The wave profile also shows a
severe asymmetry in the crest and trough height. When
there is a freak wave, a deep hole in the ocean surface
is often observed just before the extreme crest of the
freak wave erupts (see, e.g., Bascom, 1980 and
Faulkner, 2003). Similar phenomenon has been
observed in the numerical simulation of selfmodulation. Fig. 8(b) shows the wave profile at t = 319
T, or one period ahead of the moment when the wave
elevation reaches the maximum value and shown in Fig.
8(c). As opposed to usual nonlinear waves with flat
trough, the focused wave shows a deep trough that can
be interpreted as deep-hole at this time step. One of
the possible explanation of this phenomena can be the
un-locking of the second harmonic component of the
Stokes wave from the primary carrier wave. The
second harmonic component of the Stokes wave travels
in the same speed of the primary wave, or locked to the
primary wave, and always makes the crest sharper and
trough flatter in the regular wave train. But in the
highly-modulated freak wave, the second harmonic can
be unlocked and travel with the its slower free wave
speed.
To show the time evolution of amplitude and phase
of each wave component and possible unlocking of the
second harmonic component, the amplitude of the
spatial Fourier components a n (t ) and its celerity c n (t )
of the modulated Stokes wave shown in Figure 8 are
investigated. They are defined by

Figure 6 A snapshot of wave elevation due to


interaction between nonlinear Stokes wave and three
vertical cylinders.
0.12

Wave Elevation (m)

0.08

0.04

0.00

-0.04

-0.08
10

11

12

13

14
t (sec)

15

16

17

18

1 2
(x, t ) exp[inx ]dx ,
2 0
t +T
1
1
d
Im
c n (t ) =
a n (t ')dt '.
t T na n (t ') dt '
2T
a n (t ) =

Figure 7 Time history of wave elevation at the location


of maximum elevation shown in Figure 6. Solid line:
wave elevation without cylinders; Dotted line:
computed wave elevation with the cylinders; Symbols:
measured wave elevation (Swan et al, 1997)

(28)
(29)

The celerity is averaged over two periods of the carrier


Stokes wave to remove the fluctuations that have been
observed otherwise. The amplitudes of the Fourier
components and celerities of the same wave system
given in Figure 8 are shown in Fig. 9(a) and 9(b),
respectively. Also shown is the spatial maximum and
minimum of the wave elevation in Fig. 9(c). The
energy of the carrier wave (n = 10) is transferred to the
side-band instability (n = 9, 11) seeded initially. The
energy is also transferred to the other side-band
components, n = 8 and 12. Actually, the wave
components n = 8 and 12 are closer to the most
unstable mode of the Type-1 instability described in
McClean (1982). After t = 320T, permanent downshifting of the energy to the lower side-band
instabilities (n = 8 and 9) are observed.

FREAK WAVE BY SELF-MODULATED STOKES


WAVE
The self-focused waves of extreme height can be
generated from a perturbed Stokes wave train. Fig. 8
shows evolution of a Stokes wave train that was
slightly modulated by superimposing a side-band
disturbance at t = 0. Stokes wave of kA being 0.1 is
modulated by adding sinusoidal waves with wave
numbers 90% and 110% of the carrier wave, and with
amplitudes 5 % of the carrier wave. The mechanical
energy transfer from the carrier Stokes wave to the
side-band wave components results in a strongly
modulated wave packet with extreme wave height. Fig.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

0.4
0.0
-0.4
0.0

2.5

5.0

7.5

10.0

7.5

10.0

x/

(a) t = 0
k

0.4
0.0
-0.4
0.0

2.5

5.0
x/

(b) t = 319T
k

0.4
0.0

Figure 9 Time histories of (a) amplitude, |an|, of the


primary ( n = 9, 10, 11) wave components, (b) celerity,
cn, of the primary and the primary and the secondary (n
= 18, 20, 22) wave components and (c) the maximum
and minimum values of wave elevation at each time.
The numbers on the curves denote the mode number.

-0.4
0.0

2.5

5.0

7.5

10.0

x/

(c) t = 320T
Figure 8 Self-focusing of modulated Stokes wave in
deep sea (Kim & Ertekin, 2001)

where H(x) is the Heaviside function that equals to


zero when x < 0 and unity when x > 0. Further, k n is
the wave number of a linear Stokes wave for a given
wave frequency n . When the magnitude of pressure
is small, the linear solution of the waves generated by
the pulsating pressure patch can be derived from an
analytic solution given in Wehausen & Laitone (1960):

When the energy transfer from the primary mode, n =


10, to the other modes are maximal and when the
strong modulation of the Stokes wave leads to the peak
values of the maximum and minimum surface elevation,
the celerity of the second harmonic components are
dropped to certain values which are close to the
celerity of the free waves of the same length. In other
words, the phase of the second harmonic components
are unlocked from the primary components.
Since the secondary wave components are
unlocked from the primary wave components during
the emerging of the freak waves, the kinematics of
each wave components will be similar to that of the
freak waves generated by phase focusing, which will
be discussed in the next section.

L ( x, t ) =

A H ( / k
n

P A
n

sin (k n x n t n ) (31)

n =1

where
Pn =

sinh 2k n
2
3k n n 2 + sinh 2 k n

(32)

The phase of each component waves, n, are selected to


make crests of all wave components arrive at a given
location, x = xc at a given time t = t c , i.e.,

DIFFRACTION OF FREQUENCY-FOCUSED
FREAK WAVE
Waves are generated by a pulsating pressure patch
in a half-infinite domain where a vertical wall is placed
at x = 0. The pulsating pressure is given by
p (x, t ) =

n = k n x c n t c . When the generated wave is not


small, its wavelength and phase speed differ from the
linear solution. An iterative scheme described in
Chaplin (1996) has been used to consider the phase
change due to nonlinearity. Fig. 10 shows the time
history and instantaneous profile of wave elevation

x ) cos 2 2k n x sin ( n t n ) (30)

n =1

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

compared with the experimental data. This wave


corresponds to the case 765 in Chaplin et al. (1997)
that was used to study the ringing of a vertical cylinder.
The agreement between the experimental data and
numerical results is good. The wave profile near the
sharp crest looks similar to the freak wave by selffocusing shown in Fig. 8(c).

y/L

0.1
0
-0.1
-0.2
-0.4

0.3

Incident wave elevation at (0,0)


Numerical wave
Exp. (Chaplin et al., 1997)

0.2

W / h

0.2

-0.2

x/L

0.2

0.4

Figure 12 Plan view of a flat vertical cylinder

0.1

0.03

Total force on the shi p


Heading = 90o (Beam Sea)
Heading = 100o
Heading = 120o
Heading = 150o

0.02

-0.1

4
6
t / (h/g)1/2

Fx /gh 3

-0.2
10

(a) Time history


0.3

0.01
0
-0.01

/h

0.2
-0.02

0.1

0
-0.1

-3

-2

-1

0
1
(x - xc) / h

Figure 10 Time history and wave profile of focused


wave around focusing time and location

maximum value at around t / h / g = 3.5. After then,


the numerical value slightly underestimates the
magnitude of the wave load. Presumably, this is due to
the viscous effect in the experiments conducted with

0.3
Wave-exciting Moment

My / gVh

The diffraction of the focused wave by a circular


cylinder has been simulated by Kim et al. (2003) and
the wave-exciting moment is compared with the
experimental result of Chaplin et al. (1997). Fig. 11
shows the time history of the wave-exciting moment.
The numerical result agrees well until it reaches the

(b) Wave profile

Circular cylinder, a/h = 0.09524


Exp. (Chaplin et al., 1997)

0.2

Figure 13 Time history of wave exciting force at


various heading angles

Wave profile at t = 4(h/g)1/2


Numerical Wave
Exp. (Chaplin et al., 1997)

-0.2

4
t/(h/g) 1/2

small model size. At around t / h / g = 4.5 short-

0.1

duration increase of wave load can be observed both in


the numerical and experimental result. This is known
as the secondary oscillation and plays an important
role in the ringing response of the structure (Chaplin et
al., 1997; Grue & Huseby, 2002).
A vertical cylinder of a simple geometry that
has beam-length ratio similar to FPSO floating
structure is considered next (see Fig. 12). The wave
exciting force due to the freak wave profile given in
Fig. 10 is calculated for various heading angles.

0
-0.1
-0.2
0

4
t / (h/g)1/2

Figure 11 Comparison of wave exciting moment on the


circular cylinder compared with experiment

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

structures. Numerical results for the vertical cylinder


have been presented. Nonlinear wave run up and
loadings have been calculated and compared to the
available experimental results with good agreements.
Extension of the present numerical method to simulate
nonlinear wave diffraction around more practical
floating structures is in progress.
REFERENCES

(a) at t / h / g = 3.7

Bai, K. J. and Han, J. H. , A Localized Finite-Element


Method for the Nonlinear Steady Waves due to a TwoDimensional Hydrofoil, J. Ship Research, 38(1) ,1994,
42-51.
Bai, K. J., Kim, J. W. and Lee, H. G., A Localized
Finite-Element Method for Nonlinear Free Surface
Wave Problems, Proc. 19th Symp. On Naval
Hydrodyn., Seoul, Korea,1992, 113-139.
Bai, K. J. and Kim, J. W. ,A Finite-Element Method
for Free-Surface Flow Problems, J. Theoretical and
Applied Mechanics, 1(1) ,1995, 1-27.

(b) at t / h / g = 3.9

Bai, K. J., Kyoung, J. H. and Kim, J. W. , Numerical


Computations for a Nonlinear Free Surface Problem in
Shallow Water, J. Offshore Mechanics and Arctic
Engineering, 125(1) ,2003, 33-40.
Bascom, W. Waves and Beaches, Anchor Books,
Garden City, New York, 1980.
Chan, E. and Melville W., Plunging Wave Forces on
Surface-Piercing Structures, Trans. of the ASME,
Vol.111, , 1989 pp. 97-100.

(c) at t / h / g = 4.1
Fig. 14 Surface elevation around the vertical cylinder.
Heading = 120 deg.

Chaplin, J., On Frequency-Focusing Unidirectional


Waves, Int. J. of Offshore and Polar Eng., Vol. 6, ,
1996 pp. 131-137.

As shown in Fig. 13, the maximum force occurrs at the


heading angle of 100 degrees. Significant secondary

Chaplin, J., Rainey R., and Yemm R., Ringing of a


vertical cylinder in waves, J. Fluid Mech. 350, 1997,
pp. 119-147.

oscillation at around t / h / g = 4.5 can be seen. Fig.


14 shows the surface elevation around the cylinder at
120 heading. Higher run-up at the stern area is
observed than the bow area. Also observed is the
nearly vertical wave slope around the stern area at

Clauss, G., Task-related wave groups for seakeeping


tests or simulation of design storm waves, Applied
Ocean Research 21, 1999, 219-234.

t / h / g = 4.1.

Cokelet, E. D., Steep Gravity Waves in Water of


Arbitrary Uniform Depth, Phil. Trans. R. Soc.
London, A 286, 1977, pp. 183-230.

CONCLUSIONS
A finite-element method has been proposed to
simulate nonlinear wave-body interaction between
steep nonlinear waves with extreme height and floating

10

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Doneal, M., Research needs for better wave


forecasting: LEWEX panel discussion, Directional
Ocean Wave Spectra (Edited by Beam, R.C.), Johns
Hopkins Univ. Press, 1991,pp.196-204.

Swan, C., Taylor, P. H. and van Langen, H.


"Observation of wave-structure interaction for a multilegged concrete platform" Applied Ocean Research, 19,
1997, pp. 309-327.

Faulkner, D. Survival Design of Cargo Hatch


Structures, Int. Conf. Design and Operation for
Abnornal Conditions II, RINA, London, UK, 2001, pp.
43-75.

Smith, S. and Swan C., Extreme two-dimensional


water waves: an assessment of potential design
solutions, Ocean Engineering 29, 2002, 387-416.
Trulsen, K and Dysthe, K.B. Freak waves: a 3-D
wave simulation, Proc. 21st Int. Symp. on Naval
Hydrodyn., Trondheim, Norway, 1997.

Grue, J., On four highly nonlinear phenomena in


wave theory and marine hydrodynamics, Applied
Ocean Research 24, 2002, pp. 261-274.

Wehausen, J.V., and Laitone, E.V., Handbuch der


Physik, Edited by W.Flgge, Springer-Verlag, Berlin,
Vol.9,1960.

Grue, J. and Huseby M. , Higher-harmonic wave


forces and ringing of vertical cylinders, Applied
Ocean Research 24, 2002, pp. 203-214.

Zhang S.,Yue, D. & Tanizawa, K., Simulation of


plunging wave impact on a vertical wall, J. Fluid
Mech. 327, 1996, 221-254.

Kim, J. W. and Bai, K. J. A Finite-Element Method


for Two-Dimensional Water Wave Problems,
International Journal for Numerical Methods in Fluids,
30(1), 1999, pp. 105-121.

Zou, J. and Kim, C., Extreme Wave Kinematics and


Impact Loads on a Fixed Truncated Circular Cylinder,
Proc.5th Int. Offshore and Polar Eng. Conf., 1995,
pp.216-225.

Kim, J. W. and Ertekin, R. C. "A Numerical Study of


Nonlinear Wave Interaction in Irregular Seas:
Irrotational Green-Naghdi Model," Marine Structures,
13(4-5), 2000, pp. 331-347.
Kim J. W. and Ertekin R. C.,"The Study of the Spatial
Coherence of Surface Waves by the Nonlinear GreenNaghdi Model in Deep Water", 2001, Report No.:
UHMORE-01115, SOEST, Univ. of Hawaii at Manoa.
Kim, J.W., Kyoung, J.H., Ertekin, R.C. and Bai, K.J.
A Finite-Element Computation of Wave-Structure
Interaction between Long-Crested Waves of Extreme
Height and Vertical Cylinders, Proc. the 8th
International Conference on Numerical Ship
Hydrodynamics, Busan, Korea, 2003.
Kim, J. W., Kyoung J. H., Ertekin, R. C. and Bai, K. J.,
A Finite-Element Computation of Wave-Structure
Interaction between Steep Stokes Waves and Vertical
Cylinders, To appear in J. of Waterway, Port, Coastal
and Ocean Engineering, 2005.
McLean, J. W., Instabilities of Finite-Amplitude
Water Waves, J.Fluid Mech., 114,1982, pp.315-330.
Osborne, A. R., "The Random and Deterministic
Dynamics of "Rogue Waves" in Unidirectional, DeepWater Wave Trains," Marine Structures Journal, 2001.

11

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics




25 Symposium on Naval Hydrodynamics


St. Johns, Newfoundland and Labrador, CANADA, 8-13 August 2004

Prediction of Slamming Loads for Ship Structural Design


Using Potential Flow and RANSE Codes
Ould el Moctar, Andreas Brehm, Thomas E. Schellin
(Germanischer Lloyd, Germany)
Nomenclature


 
 




 







 
 
 
!
" 
#%$&$
'
)
+*+*
,

*+*
-/.10
2

.10

=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=

Pitch angle [  ]
Amplitude of pitch motion [  ]
Wave elevation [m]
Amplitude of wave elevation [m]
Heave [m]
Amplitude of heave motion [m]
Phase shift of wave
Phase shift of heave
Circular frequency of encounter in waves [s  ]
Wave number
Wave height [m]
Wave length [m]
Slamming pressure [kPa]
Reference pressure [kPa]
Vertical force [kN]
Reference force [kN]
Time [s]
Reference time [s]
Length between perpendicular [m]
Vertical accelaration [m / s ( ]
Vertical velocity [m / s]
Forward perpendiculars
Aft perpendicular
Central difference scheme
Upwind difference scheme

Abstract
The paper presents a procedure to obtain spatial mean
slamming pressures suitable for design purposes of ships
subject to slamming. The first step of this procedure consisted of using a linear seakeeping code to select equivalent
design waves by systematically computing motions and
relative velocities for different forward speeds and wave
conditions that subject the ship to slamming loads. The selection of equivalent design waves was based on the magnitude of relative normal velocity between ship and wave.
A nonlinear strip theory based seakeeping code yielded accurate predictions by simulating the motion behavior of the
ships advancing in the selected design waves. Computed
1 Reynolds

ship motions then served as part of the input for a RANSE1


code to predict slamming loads. The procedure was applied
to two ships of modern design as well as to a generic design
of a RoRo Ferry. Favorable agreement with experimental
data from model tests validated the used methods. For one
of the ships the influence of wave height, wave length and
ship speed on slamming loads was investigated.

1 Introduction
Slamming loads can induce high stresses and cause deformations of local structural components. The accurate assessment of slamming loads is essential for the design of
the ships structure. Classification society rules contain
formulas for slamming loads (e.g., Germanischer Lloyd,
2002). Generally, these formulas are adequate for conventional ships, as they are based on operational experience.
However, for many modern ships it becomes necessary to
resort to direct computations of slamming loads.
A satisfactory theoretical treatment of slamming
has been prevented so far by the complexity of the problem. Most theories and their numerical procedures were
applied on two-dimensional bodies; however, slamming is
a strongly three-dimensional nonlinear phenomenon that is
sensitive to the relative motion between the ship and the
water surface (Tanizawa and Bertram, 1998). Slamming is
characterized by highly peaked local pressures of short duration. Hence, slamming peak pressures can not be applied
on larger areas to estimate structural response to slamming
impacts. Moreover, the influence of hydroelasticity, compressibility of water and air pockets may have to be accounted for as well. Mainly because of these phenomena,
potential flow methods are not well suited to accurately
predict slamming loads (ISSC, 2003). Recent progress has
been made to develop numerical methods that predict slamming pressures (ISSC, 2000). Kinoshita et al. (1999) predicted ship motions and loads with a Navier-Stokes solver.
Although no extreme load cases were presented, the applied method has the potential to compute slamming loads
in extreme wave conditions.

Averaged Navier-Stokes Equations

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

A number of research projects during the recent past were initiated to develop numerical techniques
that reliably predict slamming loads on ships. The federally sponsored project B7 (2000), for example, dealt with
impact loads on ship forebodies. The developed technique relied on methods of computational fluid dynamics
to numerically simulate the two-dimensional water entry of
bow sections (Sames et al., 1999). Although the resulting
slamming pressures compared favorably with experimental
data, they were strongly affected by entry velocities that
had to be determined separately. Within the framework
of the European research project DEXTREMEL (2001),
slamming loads on the bow door of a generic RoRo Ferry
design were investigated. To predict bow door loads on
this ferry, Sames et al. (2001) applied two methods that
both start with linear seakeeping predictions of ship motions. Their first procedure, using a finite volume code,
relies on computed impact pressure coefficients obtained
from two-dimensional water entry simulations of vertical
bow sections. Their second procedure, using a boundary
element code, comprises two-dimensional water entry simulations of tilted bow sections. Comparison with model
test measurements showed that neither method is capable
of accurate predictions although the two methods are able
to define upper and lower bounds.
Methods that directly solve the Reynoldsaveraged Navier-Stokes equations (RANSE), possibly including the two-phase flow of water and air, are better
able to describe the physics associated with slamming.
However, the computational effort for a three-dimensional
RANSE method to simulate motions and loads on a ship at
small, successive instances of time over a long time period
appears beyond current computational capabilities. This
paper presents a numerical procedure to predict slamming
loads by first making use of the potential flow seakeeping
codes GLPANEL (Papanikolaou and Schellin, 1991; stergaard and Schellin, 1995) to select design waves and
SIMBEL (Pereira, 1988) to determine the corresponding
ship motions. The resulting ship motions then serve as part
of the input for the RANSE code COMET (ICCM, 1999)
to yield slamming loads.
The procedure was applied to predict slamming
loads for three ships. For two ships, designated Hull 1
and Hull 2, slamming loads were examined at the ships
flared bow. For the third ship, a generic design of the RoRo
Ferry investigated under research project DEXTREMEL,
extreme wave loads on bow doors were of interest. For
Hull 1 and Hull 2, computed slamming loads were used to
specify design loads for the ships bow structure, whereas
for the RoRo Ferry, predictions served mainly to validate
the procedure used to compute slamming loads. Model test
measurements of two of the ships were available for comparison with computations.

2 Computational Procedure
The objective was to obtain spatial mean slamming pressures that can be applied as equivalent static design loads
to determine scantlings of hull structural elements for
ships. This was accomplished by integrating computed
local slamming pressures over selected critical areas of a
ships hull. For modern ships with heavily flared bows,
the area under the bow flare even when there is no keel
emergence is generally subject to slamming, and for RoRo
ships, the strength of bow doors subject to extreme wave
loads is of concern. Consequently, for the first two reference ships, Hull 1 and Hull 2, selected plate fields located
under the flared bow region above the bulb were defined
as critical areas, and for the third reference ship, the RoRo
Ferry, the area covered by the bow doors was considered
critical. The following steps comprised the computational
procedure:
1. Seakeeping code GLPANEL computed ship responses in unit amplitude regular waves. Wave frequency and wave heading were systematically varied to cover all possible combinations that are likely
to cause slamming. Results were then linearly extrapolated to obtain responses in wave heights that
represent severe conditions, here characterized by
steep waves close to breaking. Under such conditions, added wave resistance and large accelerations
force ships to reduce speed. Therefore, a conservative one-third speed reduction was assumed be realistic for Hull 1 and Hull 2. For the RoRo Ferry, the
ship speed used in the computations was governed
by the model test runs.
2. Regular design waves were selected on the basis of
maximum magnitudes of relative normal velocity
between ship critical areas and wave, averaged over
the critical areas, defined as follows:
8:9<;>? = @BAC>DFE ? = G D
35476

MN

8 9IH ; ? = @
H
H

A C D HH
H

(1)

where JLK 4O is the area vector normal to a surface


patch i on
the ships hull defined by the computa9
tional M grid generated for the GLPANEL computa9
the relative velocity at this patch i, and
tions, M N K 3 is
8 9QP
P
4O P
P
P J K
P is the examined critical area made up of
i surface patches.
3. The nonlinear seakeeping code SIMBEL determined
motions of the ship under design wave conditions,
thereby accounting for the ships forward speed, the

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

swell-up of water in finite amplitude waves, as well


as the ships wake that influences the wave elevation
around the ship, particularly at moderate to high forward speeds.
4. These nonlinearly computed ship motions constituted part of the input for the RANSE code. To
compute slamming loads, numerical volume grids
surrounding the ships were generated, appropriate
boundary and initial conditions were specified, and
convergence criteria were defined.

3 Numerical Methods
3.1 Linear Panel Method
GLPANEL is a linear frequency-domain panel code that
uses zero-speed Green functions and a forward speed correction based on the so-called encounter frequency approach. A velocity potential is found by distributing singularities (sources and sinks) of constant strength over the
mean wetted surface of the hull. The velocity potential
is separated into a time-independent steady contribution
caused by the ships forward speed and a time-dependent
part associated with the incident wave system and the oscillating ship motions. For a ship advancing at constant
mean forward speed, the incident waves undergo scattering
(diffraction), leading to a diffraction wave potential that oscillates harmonically and induces a wave field of the same
frequency radiating away from the ship. The ship moving in the wave field creates body motion potentials corresponding to three translational and three angular ship motions. These body potentials and the diffraction potential
are superimposed on the incident wave potential.
The source strengths are found by satisfying the
body boundary conditions, leading to integral equations of
the second Fredholm type for each of the source strength
contributions. These equations are solved numerically by
replacing them with a system of linear equations. The
wetted hull is discretized into a finite number of small triangular or rectangular surface patches (panels) in a way
that represents the hull surface without creating leakage
gaps. The integral equations are replaced by sets of linear equations from which the desired source strengths are
determined. Dynamic pressure forces resulting from ship
motions follow from the pressure on the hull surface according to the linearized Bernoulli equation. Integration of
pressures over the hull surface yields hydrodynamic force
and moment amplitudes.

3.2 Nonlinear Strip Theory Based Method


The nonlinear strip theory code SIMBEL (Pereira, 1988),
dating back to Sding (1982) and extended over the past

two decades to account for internal loads and propulsion


system dynamics and maneuvering (Bttcher 1986), simulates large-amplitude rigid-body motions of a ship in six
degrees of freedom by time-domain integration of the motion equations. Considered are forces and moments caused
by gravity, Froude-Krylov pressure, radiation and diffraction pressure, speed effects, rudder and propeller actions,
and wind. The stationary seaway is approximated by superposition of a large number of regular waves of different
frequencies and directions. Forces and moments caused
by radiation and diffraction are deduced from the twodimensional potential flow at each of the transverse ship
sections (strips).
For large-amplitude motions, radiation and
diffraction forces and moments are obtained by integrating the pressure over the instantaneously wetted surface.
Pressure distributions depend not only on the instantaneous
position, velocity and acceleration of the ship, but also on
the motions history (memory effect). Terms account for
effects of viscous hull damping in roll using Froude number dependent coefficients (Blume, 1979). The influence of
bilge keels is approximated according to Gadd (1964) and
Martin (1958). Propeller thrust is considered, whereby a
motion equation for the propulsion system determines the
propeller rate of rotation. Rudder forces and moments are
determined by considering the orbital motion of water particles and the immersion of the rudder in the seaway.

3.3 Reynolds Averaged Navier-Stokes Solver


The RANSE solver COMET, a code that implements
interface-capturing techniques of the volume-of-fluid
(VOF) type, proved to be suitable for handling strong nonlinearities. Today, this kind of code is the obvious choice
for computing complex free-surface shapes with breaking
waves, sprays and air trapping, hydrodynamic phenomena
that should be considered to predict slamming pressures.
The conservation equations for mass and momentum in their integral form serve as the starting point. The
solution domain is subdivided into a finite number of control volumes that may be of arbitrary shape. The integrals
are numerically approximated using the midpoint rule. The
mass flux through the cell face is taken from the previous
iteration, following a simple Picard iteration approach. The
remaining unknown variables at the center of the cell face
are determined by combining a central difference scheme
(CDS) with an upwind differencing scheme (UDS). The
CDS employs a correction to ensure second-order accuracy for an arbitrary cell. A second-order central difference scheme (CDS) can lead to unrealistic oscillations if
the Peclet number exceeds two and large gradients are involved. On the other hand, an upstream difference scheme
(UDS) is unconditionally stable, but leads to higher numerical diffusion. To obtain a good compromise between accuracy and stability, the schemes are blended.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Near the ship hull, the blending factor is chosen


between 0.8 and 0.9. The Euler implicit and the three
time level methods are used to integrate in time. Pressure
and velocity are coupled by a variant of the SIMPLE algorithm (Ferziger and Peric, 1996). All equations except
the pressure correction equations are under-relaxed using
a relaxation factor 0.8. The pressure correction equations
are under-relaxed using a relaxation factor between 0.2 and
0.4 for unsteady simulations, finding in each case a suitable
compromise between stability and convergence speed.
The two-fluid system is modelled by a two-phase
formulation of the governing equations (Lafaurie et al.,
1994). No explicit free surface is defined during the computations, and overturning (breaking) waves as well as
buoyancy effects of trapped air are accounted for. The
special distribution of each of the two fluids is obtained
by solving an additional transport equation for the volume fraction of one of the fluids. To accurately simulate
the convective transport of the two immiscible fluids, the
discretization must be nearly free of numerical diffusion
and must not violate the boundedness criteria (Ferziger and
Peric; 1996). For this purpose, the high resolution interface capturing (HRIC) scheme is used (Muzaferija and
Peric, 1998). This scheme is a nonlinear blend of upwind
and downwind discretization, and the blending is a function of the distribution of the volume fraction and the local
Courant number. The free surface is smeared over two to
three control volumes. Fluid structure interaction effects
are presently not accounted for, i.e., the body is assumed to
be rigid, and the fluid is assumed to be viscous and incompressible.

A shipbound coordinate system (0, x, y, z) moving in the x-direction with the ships forward speed was
employed, with the origin 0 located at the intersection of
the baseline and the rudder axis, the x-axis directed positive towards the bow, the y-axis positive to port, and the
z-axis positive upward.

4.2 Boundary Conditions and Ship Motions


Front, side, bottom and top flow boundaries were specified
as inlets of known velocities and known void fraction distributions defining water and air regions (Azcueta, 2001). On
the hull surface a no-slip condition was enforced on fluid
velocities and on the turbulent kinetic energy. At the outlet
boundary a zero gradient pressure boundary condition had
to be satisfied. The wake flow boundary was specified as
a zero-gradient pressure boundary (hydrostatic pressure).
All computations were performed using the RNG-k- R turbulence model with wall functions. The time step size was
chosen such that the Courant number was unity on average.
The impulse equations were discretized using 85 percent
central differences and 15 percent upwind differences. As
usual, the transport equations for the turbulent kinetic energy (k) and its dissipation rate (R ) were discretized using
upwind differences.
Ship motions were realized by moving the entire
grid at each time step. Thus, all boundary conditions were
newly computed at each time step.

4 Numerical Grids
4.1 Grid Generation
The numerical volume grids surrounding the ships comprised about one million hexahedral control volumes. Because only head wave conditions were examined, one half
of the ships were modeled by setting a symmetry plane
at the centerline. To avoid flow disturbances at outer grid
boundaries, these boundaries were located at a distance of
one ship length ahead of the bow, two ship lengths aft of the
stern, and one ship length beneath the keel. The top grid
boundary was located above the deck at a distance equal to
the length of the ship. The large domain of the mesh, especially above the deck, was chosen to allow for large pitch
motions in head waves. Near the ship hull and ahead of the
ship grid density was high to resolve the wave, whereas aft
of the ship the grid became course to dampen the waves.
The grids on the surface, ahead and aft of the RoRo Ferry
are shown in Figs. 1 and 2, respectively. The innermost
cell thickness was chosen such that, on average, the dimensionless distance from the wall equaled 100.

Figure 1: Numerical grid on surface of RoRo Ferry

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

4.4 Discretization Errors


To investigate numerical errors associated with the use of
different schemes that discretize the momentum equations,
solutions were obtained on the basis of the pure UDS, the
40% CDS, and the 85% CDS approximations. In Figs. 5
and 6 sample results are presented as time series of, respectively, slamming pressure on a critical plate field at the bow
of Hull 1 and of total vertical force on the whole ship. Pressures (S TVU ) are here normalized against a maximum pressure of S&W 6 270 kPa; vertical forces (XZY ), against a maximum force of X%W 6 57540 kN; and time, against a ships
encounter period of [ W 6 6.0 s. The computations demonstrated that using the more exact 85% CDS approximation
increased pressures at the bow and decreased vertical forces
on the whole ship. For the samples shown in Figs. 5 and 6,
peak pressure increased by about seven percent, and peak
vertical force decreased by about three percent.
The total vertical force, obtained by integrating
pressures over the wetted hull surface, was only slightly
smaller using the more accurate 85% CDS approximation.
Thus, the influence of the approximation scheme on vertical force predictions was almost insignificant, in contrast to
horizontal force (resistance) predictions of bodies in calm
water (Azcueta, 2001; El Moctar, 2001).

F
K
L

GHI
J

M
N
O
PQ
R
STU
V

EDC

Figure 2: Part of numerical grid ahead (top) and behind


(bottom) of the RoRo Ferry

XY
X
Y
B

K LM

NO

RS
TUV
XW
Y
H
G
E
D B
IF
C

EDCB
F
G
M
IH
JQ
N
K
O
PL
R
TS
VU
W

TU

L
S
Y FE
D
JK
P
Q
NO
M
BR
C
GIH

4.3 Initial Conditions and Convergence

Simulation of the flow field continued until a periodic solution was reached. After a simulation time of two
to five encounter periods, depending on ship motions, ship
speed, wave height and position of the investigated plate
fields, periodically converging solutions were obtained as
seen by the typical time history sample of computed slamming pressure under the bow flare of Hull 1 (Fig. 4). For
each time step up to ten outer iterations were needed.

Figure 3: Wave generation for Hull 1


1.2
1
0.8

psl / po

Volume fractions (Albina, 2000) and velocities that initialized the flow field arose from superposition of ship speed
and orbital particle velocities of the linear Airy wave (Fig.
3). Because of natural dissipation originating from the
wave as well as numerical dissipation associated with the
applied VOF method, the profile of the generated wave
took on a more natural shape than the sinusoidal shape of
an Airy wave. The influence of numerical damping on the
wave height was taken into account. Numerical diffusion
caused by the course grid aft of the ships dampened the
incident wave to such an extent that no significant wave
reflection occurred at the outlet boundary.

0.6
0.4
0.2
0
-0.2

0.5

1
1.5
Time / To

2.5

Figure 4: Example of convergence history of slamming


pressure for Hull 1

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

85% CDS
40% CDS
UDS

psl / po

0.8
0.6
0.4
0.2
0

1.1

1.2
1.3
Time / To

1.4

1.5

Figure 5: Influence of approximation scheme on slamming


pressure for Hull 1
1.1

UDS
0.85 CDS

1
0.9

F v / Fo

0.8
0.7
0.6
0.5
0.4
0.3
0.2

0.8

1.2
1.4
Time / To

1.6

1.8

The Hamburg Ship Model Basin (HSVA) conducted model


tests of Hull 2. The model was constructed with a separated bow segment, located above the still waterline, allowing measurement of forces and moments acting on this
segment. One pressure sensor was installed on each side of
this bow segment to record local pressures. Both sensors
were located at a height of 5.0 m above the baseline, with
the port one positioned 2.0 m aft and the starboard one 0.5
m aft of the forward perpendicular (Fig. 14). A gyro unit
recorded pitch motions, three accelerometers recorded vertical accelerations, and a wave probe recorded wave elevation. Loads measured on the bow segment were corrected
for inertial effects. The free-running, self-propelled model,
constructed at a scale of 1:10, was hand operated by the
helmsman accommodated on the carriage of HSVAs large
towing tank. Tests were run in regular waves of various
heights and wave length to ship length ratios ranging from
0.8 to 1.4 at propeller turning rates corresponding to calm
water ship speeds of 14 and 16 knots.
The Maritime Research Institute Netherlands
(MARIN) performed model tests of the RoRo Ferry
(Kapsenberg et al., 1999) to measure local and global loads
on the bow and to record wave profiles just ahead of the
bow during water entry. Constructed at a scale of 1:40, the
model was attached to the towing carriage but free to move
in heave and pitch. Nine pressure sensors were installed
on the starboard side of a separated bow segment that was
in itself attached to a force gage, allowing the measurement
of forces and moments acting on this separated bow section
(Fig. 24).

Figure 6: Influence of approximation scheme on vertical


hull force for Hull 1

6 Design Waves and Ship Speeds

5 Reference Ships

For all three ships, only head wave conditions were investigated because this condition was rated most critical from
the standpoint of slamming loads on the bow.

Slamming loads were investigated at the forebody of three


reference ships. Table 1 lists their principal particulars.
Hull 1 and Hull 2 are newbuilds, where slam impacts under
the bow flare may be severe enough to cause local structural damage to the ships plating. For the RoRo Ferry,
the strength of bow doors subject to extreme wave loads
is considered a critical issue. Model test measurements of
wave-induced impact-related forces and associated slamming pressures were available for comparison with computations.
Table 1: Principal particulars of reference ships
Hull 1 (
Hull 2 (
Length
130.0 m
70.0 m
Breadth
20.0 m
15.0 m
Draft
6.0 m
5.0 m
Service speed
18 kn
16 kn
( Approximate values

RoRo Ferry
173.0 m
26.0 m
6.5 m
28 kn

For Hull 1 and Hull 2, systematic GLPANEL seakeeping computations were first performed. Based on the
resulting transfer functions of relative velocities, the wave
frequency (and wave length) of highest relative normal velocities at critical areas of the bow were identified for waves
with heights that represent the severest weather conditions
these two ships are likely to encounter. For Hull 1, a wave
height of 11.0 m was chosen; for Hull 2, a wave height
of 7.0 m. The added resistance and involuntary speed loss
under such severe conditions were assumed to cause these
ships to advance at reduced speed.
For the RoRo Ferry, the design wave conditions were taken to be those where the highest impactrelated bow forces and slamming pressures were recorded.
These conditions occurred during MARIN model test case
112006, representing a run in regular waves of 7.30 m

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

height with the ship advancing at a forward speed of 26


knots. Design wave conditions and ship speeds for the
three ships are summarized in Table 2.
Table 2: Design wave parameters and ship speeds
Wave height

#%$&$
/
Ship speed

Hull 1
11.0 m
1.0
12.0 kn

Hull 2
7.0 m
1.2
10.7 kn

RoRo Ferry
7.3 m
1.0
26.0 kn

Motions of the reference ships were determined under characteristic head wave design conditions. As head waves
were investigated, only pitch and heave motions were considered. For Hull 1 and Hull 2, SIMBEL computations
provided predictions of ship motions. For the RoRo Ferry,
GLPANEL computations yielded sufficiently accurate ship
motions by adding linearized damping terms in the damping matrix to account for viscous effects in heave and pitch.
The following time harmonic functions accurately
describe wave elevations and the resulting pitch and heave
motions at the ships center of gravity:
wave elevation [m]
heave [m]

\^]
\ 

\ ]
6

`_baLced
J

`_^aLced
\ 
J

pitch [ n ]
o

For Hull 1 and Hull 2, all pressure and force histories are
presented as nondimensional values. Pressures were normalized by a maximum value of p W = 270 kPa; forces, by
the maximum value of FW = 57540 kN for Hull 1 and F W =
2300 kN for Hull 2; and time, by the wave encounter period
of TW = 6.0 s.

8.1 Hull 1

7 Seakeeping Analyses

8 Slamming Pressures and Forces

_^ijlk
K/fhg
g
@ K/f _^ap
ced

(2)

_imjlk  O

(3)

_ri O

(4)

fqg

is the ships
where \] stands for the wave amplitude,
fZg
circular frequency of encounter in waves, and t is time.
Pitch angle was defined to be zero at time t = 0, resulting
k
k
in positive phase shifts ] and  of, respectively, wave
elevation and heave against pitch. The average value of
maximum and minimum heave and pitch motions specified their amplitudes. Table 3 summarizes the resulting
parameter values of these functions, listing heave and
@

pitch amplitudes ( \ 
and o ), their normalized values
DxA O
 |~}   |hx
u

m
t
L
v
w
u

t
 
( Js\ 
/ yz] and {
/ { 
),
\ 
k
k
K
and heave and wave phase shifts (  and ] ). Maximum
vertical velocities and accelerations at the forward perpendicular (3V and m ) are given as well.

The three-dimensional flow field surrounding the full-scale


_
hull (Reynolds number = 7.5 10 ) under the influence of
the design wave was computed as a transient process. As
seen in Fig. 8, the ships bow was exposed and the deck
emerged. The reentry of the bow after emergence not only
resulted in slamming under the ships bow flare, but it also
caused water to flow onto the deck. Figure 9 shows the
pressure distribution in the bow region, demonstrating that
impact related slamming pressures occurred in areas under
the flared bow directly above the bulb. At selected plate
fields, shown in Fig. 7, computed average pressures acting
at plate fields 1 and 2 resulted in peak values of up to 270
kPa (Fig. 9).
The functional relationships of the computed pressures correspond to the classical time history of slamming
pressures that were observed in numerous measurements
(Fig. 10). At first, the plate fields were not wetted, and
the pressure was small. The incident wave then caused the
plate fields to immerse, and slamming occurred. The pressures suddenly increased to theirs peak values and then decreased rapidly to about 75 kPa. Then the pressures slowly
decreased further until they reached atmospheric pressure.
Because of its location ahead of plate fields 2 and 3, plate
field 1 experienced the highest pressure peak of 270 kPa.
The pressure acting at plate field 3 was significantly less
than the pressures acting at plate field 1. The spurious small
amplitude pressure variations (peaks) were a consequence
of numerical inaccuracies and, therefore, did not need to be
interpreted physically.

Table 3: Motion parameters for design wave conditions


e |~} : |h

e |~}  |h

 '5
 5

'
&
$
$ 7r
)b
' $&$ zb
(

Hull 1

Hull 2

RoRo Ferry

0.43

1.33

0.34

0.43
0.49
0.52
10.8
11.3

0.98
0.79
0.59
9.6
13.5

0.94
1.14
-0.19
7.66
8.37

1
2

Figure 7: Selected plate fields for Hull 1

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Influence of Wave Length, Wave Height and Ship Speed

2.5s
F

G H
IJ

KLM
N
OP
RQ

L M

ST
UVW
XY
Za

EDC

UW
V
X
YZY X
W X Y
aZ a
b b c
c
d
G
FE
D
IC
J HB

AC
A
B
D
E
FHJI G
K
L
M
N
PO
Q
VSR

T
WU

6.000e+00

4.800e+00

3.600e+00

2.400e+00

1.200e+00

7.749e07

1.200e+00

2.400e+00

3.600e+00

4.800e+00

6.000e+00

To demonstrate the dependence of wave length on slamming pressures, computations were performed for Hull 1
advancing in waves of differing lengths, with wave height
and ship speed kept constant. The resulting pressures at
plate field 1 are shown in Fig. 11. In longer waves at /L =
1.2 ( = wave length, L = ship length), although ship motion amplitudes as well as vertical velocities at the forward
perpendicular were larger than under design wave conditions ( /L = 1.0), pressures turned out to be less. This was
because slamming pressure at plate field 1 was influenced
by the normal relative velocity, and this velocity depended
also on the phase relationship between wave and ship motions and to a lesser degree on absolute ship motions. Pressures behaved similarly at plate fields 2 and 3.

4.5s

R
ST
UV
WX

Y Z

FEDC
LKJIHG
PONM
Q

BC
ED
F
G
aA
KIJ H
B
O
PLNM
A
R
TS Q
U
W
JZ M
FGHI J K
F IH K
LN
R
S
QT
CBAG
Y
VPOU
A BCDE
b ED
L

M
NO

P
Q
R
TS
U
V

WX
YZ
ab

cd

6.000e+00

4.800e+00

3.600e+00

2.400e+00

1.200e+00

7.749e07

1.200e+00

2.400e+00

3.600e+00

4.800e+00

6.000e+00

Figure 8: Water on deck and bow emergence of Hull 1

5.90s

P/Po

M
L KJI HGF E

C B A
A
B

L
J

F
HG

ED

1.000e+00

8.333e01

6.667e01

5.000e01

3.333e01

1.667e01

0.000e+00

K
M
L
K

Figure 9: Pressure distribution on Hull 1 while slamming

1.2

Panel 1
Panel 3

psl / po

0.8

Slamming pressures under design wave conditions were obtained also for three different ship speeds,
namely, for Hull 1 traveling at one-third and two-third
design speed as well as at full design speed. The corresponding motion parameters for design wave conditions at
these ship speeds are listed in Table 4. The influence of
ship speed on slamming pressures depended strongly on
the location of the plate fields. For the sample shown here,
the maximum pressure increased at higher ship speeds.
Doubling ship speed from a relatively slow 6 kn to 12 kn
led to a 35 percent increase in peak slamming pressure at
plate field 1. Increasing the ship speed from 12 to 18 kn,
on the other hand, caused an increase of only about ten
percent of the peak slamming pressure. Depending on the
location of the investigated plate field, it is possible that
an increase in ship speed leads to a decrease in slamming
pressure.
Table 4: Motion parameters of Hull 1 for different ship speeds
at design wave conditions

0.6
0.4

Ship speed

|~}  |h

e |~}  |h

 'Vr

0.2
0
-0.2

To quantify the influence of wave height on slamming pressures, computations were conducted by varying
the height of the design wave, with ship speed kept constant. The resulting pressure at plate field 1, shown in
Fig. 12, indicates that the computed pressure peak increased weakly nonlinear with wave height. The influence
of wave height depended also on other parameters, such as
the location of the investigated plate field. The influence
of wave height on peak pressures was similar at the other
plate fields.

0.8

1.2
1.4
Time / To

1.6

1.8

  '5
$ $ 7r
) &
' $&$ 7r
(

Figure 10: Time history of slamming pressure for Hull 1 at


design conditions

Copyright National Academy of Sciences. All rights reserved.

6 kn

12 kn

18 kn

0.28

0.43

0.48

0.40
0.40
0.99
8.0
6.0

0.43
0.49
0.52
10.8
11.3

0.41
0.82
0.11
11.7
14.0

Twenty-Fifth Symposium on Naval Hydrodynamics

8.2 Hull 2

1.2
/L=1.0

1
0.8
psl / po

Extreme loads caused by slamming were predicted at two


selected plate fields located on the separated bow section of
Hull 2. Figure 14 shows locations of the investigated plate
fields under the flare of the ships bow.

/L=1.2

0.6
0.4

/L=0.8

0.2
0
0.6

0.8

1.2
1.4
Time / To

1.6

1.8

Figure 11: Influence of wave length on slamming pressure


at plate field 1 for Hull 1 at design wave height of 11.0 m
and ship speed of 12.0 kn

1
HW = 11m
HW = 8m
HW = 5m

psl / po

0.8
0.6
0.4
0.2
0
0.6

0.8

1.2
1.4
Time / To

1.6

1.8

Figure 12: Influence of wave height on slamming pressure


at plate field 1 for Hull 1 at   6 1.0 and ship speed of
12.0 kn

1.2

v=18 kn
v=12 kn

psl / po

0.8

Based on the design wave conditions and the corresponding ship motions, the RANSE code computed the
flow around Hull 2. Time histories of predicted pressure averaged over the investigated plate fields are plotted in Fig.
15. The pressure peak at plate field 1 was twice as high
as the pressure peak at plate field 2. This was mainly because at plate field 1, located ahead of plate field 2, a water
jet impinged on the flared bow, caused by the accelerated
flow around the bulb. Also, the relative normal velocities at
plate field 2 were smaller than those at plate field 1.
Seakeeping model tests of Hull 2 yielded experimental data of ship response, including ship motions and
accelerations, that generally compared favorably with nonlinear computations from SIMBEL. Under the HSVA test
conditions (Table 5), the RANSE solver performed computations of wave-induced impact-related slamming loads
and local pressures acting on the separated bow segment
of the ship, using the computed motions as part of the input. Simulations of the flow field were performed for two
consecutive encounter periods, while the measurements
lasted over several periods. Computed forces and pressures
hardly changed after the first period; however, the corresponding measurements varied. Over the first two periods,
the computed vertical force on the separated bow section
compared favorably with the experimental data; see Figs.
16 and 17 for HSVA test runs 1 and 2, respectively. For
the corresponding slamming pressures, shown in Figs. 18
and 19, the functional relationship of computed values correlated well with measurements, whereas computed peak
values sometimes deviated from experimental data. This
deviation was largely attributed to the relatively strong
variation of the measured peaks, most likely caused by the
inability of the model to attain steady state conditions during tests in regular waves.

v=6 kn

0.6

Table 5: Conditions of HSVA test runs for Hull 2


 # $:$
 
Test run
Calm water ship speed
1
3.5 m
1.0
14 kn
2
3.5 m
1.2
14 kn

0.4
0.2
0
-0.2

0.8

1.2
Time / To

1.4

1.6

Figure 13: Influence of ship speed on slamming pressure at


plate field 1 for Hull 1 at 5 6 1.0 and wave height of
11.0 m

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

0.8

RANS
Experiment

0.7
0.6
Fv / F o

0.5
1

0.4
0.3
0.2
0.1
0
-0.1
0.5

Figure 14: Selected plate fields for Hull 2

1.5
2
Time / To

2.5

0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
-0.1
0.7

Panel 1
Panel 2

0.6

0.4

0.8

1.2

0.9
1
Time / To

1.1

1.2

0.2

0
0

0.5

1.5
Time / To

2.5

Figure 18: Time history of pressure on plate field 1 of Hull


2 for HSVA test run 1

0.5

RANSE
Experiment

RANSE
Experiment

0.4

0.8

0.3

0.6

psl / po

Fv / F o

0.3

0.1

Figure 15: Time history of slamming pressure for Hull 2 at


design conditions

0.4
0.2

0.2
0.1

0
-0.2
0.5

RANSE
Experiment

0.5

psl / po

psl / po

Figure 17: Time history of vertical force on bow section of


Hull 2 for HSVA test run 2

0
1

1.5
2
Time / To

2.5

Figure 16: Time history of vertical force on bow section of


Hull 2 for HSVA test run 1

0.5

1.5
Time / To

2.5

Figure 19: Time history of pressure on plate field 1 of Hull


2 for HSVA test run 2

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

300
200
100

RANSE
Experiment

100
50

Time [s]

20

Time [s]

200
150
100

Sensor 7

Time [s]

Experiment
RANSE

150

Pressure [kPa]

Pressure [kPa]

Sensor 8

200
Experiment
RANSE

100
50
0

100
50
0

Time [s]

Time [s]

Figure 21: Time histories of bow door pressures at selected


pressure sensors for RoRo Ferry
Vertical force

4000

modell test
RANSE

20000

Horizontal force

15000

Force Fx [kN]

Force Fz [kN]

For the forth period of simulations (between 18


and 24 s in Fig. 20), Fig. 21 presents sample plots of comparable pressure traces at selected sensor locations. Computed slamming pressures also agreed closely with measurements. Pressure sensor locations are listed in Table 6
and shown in Fig. 24. Coordinates x, y, z of the pressure
sensors refer to full-scale distances from the forward perpendicular, the baseline, and the centerline of the ship, respectively.
To obtain forces on bow doors, local pressures
were integrated over the area of the bow doors. The resulting time series of wave-induced vertical and horizontal
impact forces are shown in Fig. 22 together with comparable measurements from model tests. Close agreement with
measurements was achieved only for the horizontal force
component; vertical force predictions were significantly
higher than measurements. Although the vertical force predictions were obtained by integrating pressures over actual
(reduced) bow area, this apparent disagreement was mainly
caused by the flawed idealization of the ships forebody in
that the height to main deck of the idealized hull exceeded

Experiment
RANSE

150

Figure 20: Relative vertical motions at forward perpendicular for RoRo Ferry

0
2

200

30

50

-8
10

Time [s]

Sensor 5

250

150

300
Experiment
RANSE

-4

200

0
0

Sensor 4

250

Time [s]

200

300

100

0
0

Relative vertical motions at FPP

Experiment
RANSE

400

Pressure [kPa]

400

Sensor 2

500
Experiment
RANSE

Pressure [kPa]

12

Sensor 1

500

Pressure [kPa]

Rel. dist. betw. free surface and point at FPP [m]

To validate the employed methods, comparative computations were performed for the RoRo Ferry under conditions
corresponding to MARIN model test no. 112006, representing a run in regular head waves of 7.30 m height and
10.5 s period with the ship advancing at a constant forward
speed of 26 kn (Table 2).
For this ship, GLPANEL was used to obtain ship
motions, albeit with additional terms introduced in the
damping matrix to account for viscous effects in heave
and pitch. The subsequent COMET computations, using
these motions as input, yielded wave elevations, relative
motions along the hull, and local pressures acting at the
hull surface. Plots of computed and measured relative vertical motion at the ships forward perpendicular, shown
in Fig. 20, demonstrated that the simulated bow motion,
which was the motion that directly affected wave-induced
impact-related loads, generally agreed favorably with measurements.

the corresponding height of the model by 1.80 m, a significant amount of more than 11 percent. To obtain comparable results, such flow should have been simulated because
it occurred during the investigated model test. As this test
was run in relatively high waves (7.30 m), an additional
simulation was carried out in waves of lower height (4.46
m), where water did not flow over the deck. The resulting
bow door forces, shown in Fig. 23, now agreed favorably
with model test measurements.

Pressure [kPa]

8.3 RoRo Ferry

10000
5000

-4000

-8000
modell test
RANSE

0
0

Time [s]

-12000
0

Time [s]

Figure 22: Vertical (left) and horizontal (right) force history on bow door of RoRo Ferry in 7.30 m high waves

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Horizontal force

Vertical force

12000

Experiment
RANSE

10000

Force [kN]

Force [kN]

8000
6000
4000

-2000

-4000

2000
0
-2000

Experiment
RANSE

-6000
12

14

Time [s]

16

12

14

Time [s]

16

Figure 23: Vertical (left) and horizontal (right) force history on bow door of RoRo Ferry in 4.46 m high waves

Figure 24: Location of the bow door and positions of the


pressure sensors for the RoRo Ferry
Table 6: Positions of pressure sensors at bow of the RoRo Ferry
Sensor No.
1
2
3
4
5
6
7
8
9

x [m]
3.0
4.9
6.9
0.4
1.6
2.6
-2.7
-2.7
-2.7

y [m]
0
0
0
2.30
3.60
4.90
3.75
5.65
7.50

z [m]
10.7
12.4
14.1
10.7
12.4
14.1
10.7
12.4
14.1

9 Discussion
Accurate prediction of slamming loads continues to be a
difficult undertaking, mainly because of the influence of a
large number of parameters involved. The procedure presented here attempted to combine the physics of a ship
in a seaway together with the use of advanced numerical
techniques. Most methods used nowadays rely on potential theory or on two-dimensional RANSE solvers without
accounting for the effects of waves. The proposed procedure was used to predict slamming loads for the three reference ships, yielding pressure traces that compared favorably with model test measurements. Furthermore, it was

possible to demonstrate the influence of hull shape (i.e.,


bulbous bow) on slamming loads. From a practical standpoint, the procedure represents a compromise between attainable accuracy and computational effort.
To model the seaway, we relied on the definition
of an equivalent regular design wave. This approach has
been successfully applied to many kinds of sea load predictions for ships. Breaking wave criteria together with
wave scatter diagrams of operating sea areas together with
past experience with similar ships served as a basis to identify design wave conditions. Although a design wave only
roughly approximated the real wave conditions when slamming, there is at present no alternative available for practical application. Modeling the natural seaway by superimposing a large number of regular waves would have made
it necessary to deal with a prohibitively large number of
variables.
Although structural deformations affect impactrelated loads, this influence was not accounted for here.
Had this effect been included, most likely predicted slamming pressure peaks would have been less. Thus, from the
standpoint of structural safety, our computations were conservative. The purpose of integrating the RANSE computed local pressures over critical areas (structural plate
fields) to specify equivalent static design loads was meant
to at least partly account for this phenomenon. The generally favorable agreement of computed impact-related
forces with measurements validated our approach.
The potential flow theory based code, used to obtain normal relative velocities, was not able to account for
viscous effects leading to flow separation and vortex formation. As these relative velocities were a criterion to define
the design wave conditions, further investigations are necessary to ensure the relevancy of such predictions for the
procedure presented here.

10 Concluding remarks
Computed results of wave-induced slamming loads demonstrated that the procedure presented here was capable of
predicting slamming suitable for design of a ships structure. The generally favorable comparison of computations
with model test measurements validated the methods used.
The method, applied to two newbuilds Hull 1 and Hull 2,
obtained equivalent static slamming loads suitable for the
design of the ships bow structure.
The presented procedure relied on the combined
use of potential flow codes and a RANSE solver. It was
first necessary to obtain reliable predictions of ship motions. As slamming typically occurs in high waves, motions could not be obtained directly from the linear seakeeping code GLPANEL. However, the efficient application of GLPANEL served to identify the design wave conditions for Hull 1 and Hull 2. Subsequently, the nonlinear

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

code SIMBEL yielded sufficiently accurate motion predictions that were used as input for the RANSE code COMET.
This code is currently being modified to also solve the ship
motion equations, so that in future ship motions will no
longer have to be supplied as part of the input.
The RANSE solver yielded pressures at all surface
patches on a ships hull generated by the numerical volume
grids surrounding the ships. To determine slamming loads
useful for design, these pressures had to be integrated over
so-called critical areas. Selecting the location and size of
critical areas was based on structural design considerations
of the bow and on favorable correlation of computed slamming loads with model test measurements. The specified
plate fields located under the flare of Hull 1 and Hull 2
represented such critical areas. For Hull 1, the three selected critical areas represented total plate fields bounded
by frames and stiffeners of the ships bow structure. For
Hull 2, the smaller of the two critical areas represented a
single plate field; the other critical area, a total plate field.
Computations were performed in head seas only
because slamming loads were then expected to be large.
The computed pressure histories corresponded to the classical time history of impact related slamming pressure that
has been observed in numerous measurements. This not
only holds true for peak values, but also for the duration of
slamming pressures simulated.
The investigated dependence of wave length,
wave height and ship speed on slamming loads showed that
systematic computations were required to obtain equivalent
design wave parameters and the appropriate ship speed to
reflect extreme slamming conditions. Continued practical
experience in applying the presented procedure is essential
for its success.

Acknowledgement
Thanks are due to A. Khlmoos and R. Pereira for their
assistance with the numerical computations. The opinions
expressed herein are those of the authors.

References
Albina, F.-O., A Procedure to Set Volume Fraction for
Calculations with COMET, AB 3.13, 2000, Technical
University Hamburg-Harburg, Hamburg.
Azcueta, R., Computation of Turbulent Free-Surface
Flows Around Ships and Floating Bodies, Dissertation at
the Technical University Hamburg-Harburg, 2001, Hamburg (in German).
Blume, P., Experimentally Determined Coefficients for Effective Roll Damping and Application to Estimate Extreme
Roll Angles, Journal of Ship Technology Research , Vol.
26, 1979 (in German).
Bttcher, H., Ship Motion Simulation in a Seaway Using
Detailed Hydrodynamic Force Coefficients, Proceedings

of the 3 International Conference on Stability of Ships


and Ocean Vehicles, Gdansk, 1986.
Brackbill, J., Kothe, D., and Zemaach, C., A Continuum Method for Modelling Surface Tension, Journal of
Computational Physics, Vol. 100, 1992, pp. 335-354.
B7, Joint Project Life Cycle Design - Computation of Hydrodynamic Impact Loads, BMBF Contract MTK 0577 A,
Ministry of Education, Science, Research and Technology
of the Federal Republic of Germany, FG 2000.143, 2000,
Germanischer Lloyd, Hamburg (in German).
Demirdiz, I., Muzaferija, S. and Peric, M., Numerical
Method for Coupled Fluid Flow, Heat Transfer and Stress
Analysis Using Unstructured Moving Meshes With Cells
of Arbitrary Topology, Computer Methods in Applied
Mechanics and Engineering, Vol. 125, 1995.
DEXTREMEL, Design for Structural Safety under Extreme Loads, BriteEuram III project, contract BRPR-CT0513, 2001.
El Moctar, O., Numerical Evaluation of Hydrodynamic
Loads on Maneuvering Ships, Dissertation, Technical
University Hamburg-Harburg, Hamburg, 2001 (in German).
Ferziger, J. and Peric, M., Computational Methods for
Fluid Dynamics, Springer-Verlag, Berlin, Heidelberg, New
York,1996.
Gadd, G.E.: Bilge Keels and Bilge Vanes, 64, National
Physical Laboratory Ship Division, 1964.
Germanischer Lloyd, Rules for Classification and
Construction, I Ship Technology, 1 Seagoing Ships, 1 Hull
Structures, Edition 2002, Hamburg.
ICCM, User Manuel COMET Version 2.0, Institute of
Computational Continuum Mechanics GmbH, Hamburg,
1998.
ISSC, Technical
Committee I.2 LOADS, Proceedings

of the 15 International Ship and Offshore Structures
Congress, Mansour, A.E. and Ertikin, R.C., Edts., Vol. 1,
San Diego, 2003, pp. 87-90.
ISSC, Technical
Committee I.2 LOADS, Proceedings

of the 14 International Ship and Offshore Structures
Congress, Ohtsubo, H. and Sumi, Y., Edts., Vol. 1, Nagasaki, 2000, pp. 102-107.
Kapsenberg, G.K., van den Broek, N.A.M., and van Doeveren, A.G., DEXTREMEL Measurement of Bow Door
and Green Water Loads on a Ferry, Vol. 2, 14361-2-ZT,
1999, Maritime Research Institute Netherlands (MARIN),
Wageningen.
Kinoshita, T., Kagemoto, H., and Fujino, M., A CFD
Application to Wave-Induced Floating-Body Dynamics,
Proceedings of the International Conference on Numerical
Ship Hydrodynamics, Nantes, 1999.
Khlmoos, A., Otto, S., Sames, P.C., Schellin, T.E., and
Schiff, C., Part B7 of Joint Project Life Cycle Design - Computation of Hydrodynamic Impact Loads, FG

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

2000.143, 2000, Germanischer Lloyd, Hamburg (in German).

Martin, M., Roll Damping Due to Bilge Keels, Contract 1611 (01), 1958, Iowa Institute of Hydraulic Research,
Iowa.

drodynamic Panel Method for Practical Analysis of Ships


in a Seaway, Transactions of the Schiffbautechnische
Gesellschaft, Vol. 89, 1995, pp. 561-576.
Papanikolaou, A.D. and Schellin, T.E., A ThreeDimensional Panel Method for Motions and Loads of
Ships with Forward Speed, Journal of Ship Technology
Research, Vol. 39, 1991, pp. 147-156.
Pereira, R., Simulation of Nonlinear Sea Loads, Journal
of Ship Technology Research, Vol. 35, 1988, pp. 173-193.

Muzaferija, S. and Peric, M., Computation of FreeSurface Flows Using Interface-Tracking and InterfaceCapturing Methods, Nonlinear Water Wave Interaction,
Computational Mechanics Publ., Southampton, 1998, pp.
59-100.

Sames, P.C., Kapsenberg, G.K., and Corrignan, P., Prediction of Bow Door Loads in Extreme Wave Conditions,
Proceedings of the International Conference Design and
Operation for Abnormal Conditions II, Royal Institution of
Naval Architects, London, 2001.

Ogawa, Y., Matsunami, R., Minami, M., Tanizawa, K.,


Arai, M., Kumano, A., and Miyake, R., Estimation
Method for Probability Density Function of the Water
Impact Pressure of Post-Panamax Containership
due to

Bow Flare Slamming, Proceedings of the 7 International
Conference on FAST SEA TRANSPORTATION, University of Naples, Vol. II, Ischia, 2003, pp. 25-32.

Sames, P.C., Schellin, T.E., Muzaferija, S. and Peric, M.,


Application of a Two-Fluid Finite Volume Method to Ship
Slamming, Journal of Offshore Mechanics and Arctic
Engineering, Vol. 121, 1999, pp.47-52.
Sding, H., Prediction of Ship Steering Capabilities,Journal
of Ship Technology Research, Vol. 29, 1982, pp. 3-29.
Tanizawa, K. and Bertram, V., Slamming, Handbuch der
Werften, Vol. XXIV, 1998, Hansa-Verlag, pp.191-210.

Lafaurie, B., Nardone, C., Scardovelli, R., Zaleski,


S., and Zanetti, G., Modeling Merging and Fragmentation in Multphase Flows with SURFER, Journal of
Computational Physics, Vol. 113, 1994, pp.134-147.

stergaard, C. and Schellin, T.E., Development of a Hy-

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION

AUTHORS REPLY

Volker Bertram
ENSIETA, France

Forward speed significantly affects the


prediction of wave-induced loads, especially impact
related loads (slamming) in the bow region of ships.
When making load predictions, our experience
showed that it is necessary to account for speed
reduction in waves, and our choice of the 33 percent
speed reduction was a result of this experience.
Curiously enough, for the RoRo ferry, one of the
sample ships analyzed in our paper, loads obtained
from spectral analysis techniques by considering
speed reductions based on added resistance in
different sea states compare favourably (within 5
percent) with loads based on the 33 percent speed
reduction.
Regarding the Green function code, yes, for
the first step a linear strip method could have been
used. However, there are two aspects that speak for
using a Green function code. First, the discretization
of the hull can be made compatible with a finite
element idealization of the ships structure and,
second, computer time is no longer a critical issue.
With the exception of the RoRo ferry, a
generic design, all experimental data for validation of
seakeeping tests were confidential and thus not
available for researchers, as were lines plans and
electronic hull descriptions. From consistent
measurements, obtained for more than a dozen model
test runs performed under similar conditions, we felt
that the measurements were sufficiently accurate and
reliable for the intended purpose. We did not conduct
RANSE computations with initial heel angles.
However, we did make alternative runs with
successively refined grids. Furthermore, impact loads
were measured at plate fields under the bow flare
located on both sides of the hull not symmetrically
opposite to each other.
We added linearized damping terms to
account for viscous effects not only in roll, but also in
heave and pitch. Admittedly, viscous effects in heave
and pitch were small, and they were restricted to
frequencies in the neighbourhood of resonance. They
could have been neglected.

Seakeeping requires even more than other


simulations for naval architects intelligent
modeling to get the necessary information at
necessary accuracy at acceptable effort. The authors
succeeded
convincingly
in
this
endeavor
demonstrating impressively what is now frontier
state-of-the-art for three-dimensional slamming
analyses for ships. While research on twodimensional, fundamental slamming problems still
has its place, the progress in slamming analyses for
real three-dimensional ships is gratifying.
The first step of the presented analysis
employs a three-dimensional Green function method.
A 33% speed reduction was assumed for Hull 1 and
Hull 2. Why 33%, why not 10%, 25%, or 40%? Is the
choice based on some publication, GL experience, or
an engineering guestimate? Could the Green
function code be substituted by a linear strip method
for the first step?
Experimental data for local flow details for
validation of seakeeping tests are scarce. The shown
comparisons are all the more interesting. Are the
details of the model tests of HSVA and MARIN
public, i.e. are lines plans or electronic hull
descriptions available for other researchers? The
phenomena are probably strongly non-linear. Do you
have any information on the accuracy or repeatability
of these tests? Due to the high expense of seakeeping
model tests, these tests were probably not repeated,
but could you in your numerical simulations change
e.g. the initial heel angle by 1 and then publish the
influence on the pressure results? Also, the shown
grids appear to be reasonably fine, but have you tried
further refinement in the forebody?
In chapter 7, you mention adding linearized
damping terms [] to account for viscous effects in
heave and pitch. I am surprised that viscosity should
play a significant role for heave and pitch. Perhaps
geometric non-linearities in model tests have similar
effects are added viscosity in the numerical
simulation? If the viscosity influence is considerable
how do model-scale results differ from full-scale
results? The model tests violate Reynolds similarity
and at model scale 1:40 scaling effects would be
strong. While full-scale seakeeping tests to determine
viscous influence appear at least very difficult, could
you perform numerical simulations to estimate the
scaling effect for heave and pitch?

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Hoyte C. Raven
MARIN, The Netherlands
As the peak pressures occurring locally
during slamming seem to be a very rapid event, one
would suspect that their prediction would depend
strongly on the cell size and time step used. Have
you checked that dependence?
AUTHORS REPLY
The mesh size in the sensor areas was about
0.025 m (full scale). Increasing the mesh size
substantially decreases the pressure peak. The time
step used in the computation was about 0.001s. We
think that dt=1ms is small enough to capture
slamming.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

St. Johns, Newfoundland and Labrador, CANADA, 8-13 August 2004

Numerical and Experimental Analysis of Bow Flare


Slamming on a Ro-Ro Vessel in Oblique Waves
Ole A. Hermundstad and Torgeir Moan
(Centre for Ships and Ocean Structures,
Norwegian University of Science and Technology - NTNU, Norway)

ABSTRACT
A 2D numerical slamming calculation method
(Zhao et al 1996) is validated for a 20-knot, 120meter car carrier. Nonlinear strip theory is used to
calculate ship-wave relative motions, which are
input to the slamming analysis. Model tests of the
car carrier have been carried out in regular head,
bow and bow quartering waves of various heights.
Slamming on two panels in the upper part of the
bow flare is studied. It is found that the water pileup around the bow due to the forward speed of the
vessel increases the slamming pressures
significantly. A simplified way of including this
effect is presented. When the calculated slamming
pressures are corrected for 3D effects they compare
well with the measured data. Since the effect of the
wave elevation due to forward speed and the effect
of three-dimensional flow act in opposite
directions, excluding both of them produce results
that agree quite well with the experiments,
especially for the most severe slamming events.

INTRODUCTION
A proper estimate of the extreme slamming loads is
important in the design of hull plating and
scantlings in the bow flare of Ro-Ro vessels. This
is usually accomplished by using formulae given
by the classification societies. However, these
formulae are empirical and therefore not
necessarily valid for novel designs. Hence, there is
a growing need for direct calculation methods
Many methods have been developed for the
calculation of slamming pressure on a body that
penetrates the water surface with a prescribed
velocity. Among the earliest was the method of von
Karman (1929). Wagner (1932) presented an

asymptotic solution for water entry of twodimensional sections with small local deadrise
angles. Two fluid domains were used. The inner
domain contains a jet flow at the intersection
between the body and the free surface. In the outer
domain, the body boundary condition and the
dynamic free-surface condition (=0) were
transformed to a horizontal line. The kinematic
free-surface condition was used to determine the
intersection point between the free surface and the
body in the outer domain. Zhao et al (1996)
presented a simplified approach, which is a
generalized version of Wagners method, where the
body boundary condition is imposed on the real
body surface and not on a horizontal line. This is a
robust numerical method that is valid for quite
general section shapes. Small horizontal and
rotational velocity components are allowed in
addition to the vertical velocity. The method has
later been extended to include flow separation from
knuckles (Zhao et al 1997). It was found by Zhao et
al (1996) that for a bow flare section the simplified
method gave a peak pressure in the flare that was
about 4% higher than predicted with the more
refined numerical method of Zhao and Faltinsen
(1993).
General 3D methods for slamming analysis
have also been presented (Muzaferija and Peric
1998, Schumann 1998, Sames et al. 1998). These
methods solve the 3D Navier-Stokes equations.
Good agreement with drop tests has been
documented (Muzaferija et al 1998), but the
methods
require
significant
computational
resources.
In order to assess the slamming pressure on a
ship hull moving in waves, the slamming
calculations must be combined with calculations of

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

the ship and wave motions. We may distinguish


between two main approaches, namely the kfactor methods and the direct methods. The kfactor methods are based on the use of slamming
coefficients or so-called k-factors (e.g. Ochi and
Motter 1973). These relate the slamming pressure
to the square of the impact velocity, and they can
be calculated, or obtained experimentally, prior to
the ship motion analysis. In a direct method, one
starts out with the ship motion calculations, and
then applies a slamming calculation method each
time a slamming event takes place. Normally,
many slamming events need to be analyzed, and
the direct methods are therefore more
computationally demanding than the k-factor
methods. But they are presumably more accurate
since one simulates the entire slamming event and
uses the proper time-variation of the relative
velocity instead of a constant value. Pile-up effects
can also be properly included, provided that the
slamming calculation method is sufficiently
refined.
Sames et al (2001) compared results from a kfactor method (using local slamming coefficients
obtained from a 2D volume-of-fluid method) and a
direct method (using the slamming method of
Faltinsen and Zhao (1993)) with data from model
tests. Linear methods were used for seakeeping
calculations. Comparisons were made with results
from model tests of a Ro-Ro vessel. The k-factor
method produced lower values for the total vertical
force on the bow door of the vessel as compared to
the direct method. The experimental results were
typically between the two theoretical predictions,
and both methods agreed reasonably well with the
experiments for most cases. For pressures in the
bow flare, on the other hand, the agreement was not
good for any of the methods. Their direct method
could only handle head seas.
In the present paper we use a direct method.
For slamming pressure calculation we apply the
method of Zhao et al (1996). A conventional
nonlinear strip theory is used to calculate the ship
motions, and comparisons are made with results
obtained with a linear strip theory. We will also
include the effect of the steady wave elevation, and
we will account for 3D flow effects in a simplified
manner. The importance of these additional
refinements will be demonstrated. For validation
we use model test results for a 120 m Ro-Ro vessel.
Little validation has previously been done for
oblique seas, and we therefore present results also
for bow sea cases.

THEORY
General Approach
The present method follows the three steps
common to most direct slamming calculation
methods:
1.
2.
3.

Calculate ship motions.


Calculate the time-series of the relative
velocities between the ship sections and the
incident wave surface during a slam.
Calculate the slamming pressure on each
section when it enters a calm water surface
with the relative velocities from step 2.

Note that the ship motion calculations are


decoupled from the slamming load calculations.
Hence, we assume that the ship motions are
unaffected by the slamming forces. We will study
various refinements in the way the relative
velocities between the ship and the water surface
are calculated. Each of the three steps is described
below.
Calculation of Ship Motions
The ship motions are calculated using a nonlinear
strip theory method. The method is similar to that
presented by Wu and Moan (1996) and Wu and
Hermundstad (2002), except that the hull is
considered completely rigid in the present case, and
the only nonlinearities come from hydrostatics and
Froude-Krylov forces. This is similar to the
formulation denoted Nonlinear 1 in Wu and
Hermundstad (2002).
The method consists of two steps, in which
the linear transfer functions for the rigid body
modes are calculated first. The time-series of the
total responses in a selected sea-state are then
calculated from convolution of the linear impulseresponse functions, and by adding nonlinear
hydrostatics and Froude-Krylov forces.
Calculation of linear transfer functions is
performed with a conventional strip theory. The 2D
boundary value problems are solved with a hybrid
BEM method using Rankine sources in the inner
domain and analytical expressions in the outer
domain. An integral theorem is used when the
linear forces are calculated, in order to avoid
numerical differentiation of the velocity potentials
in the ships longitudinal direction (see
Hermundstad et al 1999).

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Calculation of Relative Velocities


The slamming calculations require the following
input data:

Geometry of the section(s) in consideration


Roll angle of the section(s) relative to the
water surface
Time-series of the relative vertical-, horizontaland roll-velocity during the impact
Initial draft of the section(s)

The relative roll angle is calculated at the start of


the slamming scenario. This is when the lowest
point of the section enters the water. If the section
never emerges completely, the slamming scenario
starts when the section emergence is at its
maximum. In the latter case, the initial draft of the
section is nonzero and the initial relative vertical
velocity is zero. Note that the relative roll angle is
constant during the impact, while the relative roll
velocity can be updated during the slamming event.
The relative vertical velocity, V(t), can be
calculated by considering the vertical velocity of
the section relative to the vertical velocity of the
incident water surface. When the ship has a
forward speed, U, it is common practice to include
the component of this speed that acts in the section
plane. This component originates from the pitch
angle, 5, and for small pitch angles it can be
written U5, when 5 is given in radians. The
relative velocity then becomes V(t) + U5.
As a further refinement in the calculation of
the relative vertical velocity, one may include the
vertical velocity of the wave elevation caused by
the ship. These are diffraction waves as well as
waves generated by the unsteady motions and the
steady forward motion of the ship. It is not
common to include any of these waves in slamming
calculations. However, we will include the waves
caused by the steady forward motion of the ship,
and we will demonstrate later that this steady wave
elevation can contribute significantly to the
slamming pressures in the bow flare.

For a given ship, the steady wave elevation around


the bow will depend on the forward speed as well
as the draft and trim. During a bow flare slamming
event, the draft in the bow may change from zero to
the weatherdeck. For typical bows of Ro-Ro ships
the steady wave elevation will be relatively small
for the design draft, and it will increase as the
waterline moves up into the flared area. In our
simplified approach we will neglect dynamic
effects on the steady wave elevation, and simply
calculate it for a number of drafts ranging from
nearly zero to the highest point in the flare where
we expect slamming to occur. We will use the
program WAVERES (Zhao 2003) for these
calculations. The program uses a linear 3D panel
method, and nonlinear effects in the bow region are
included by using a 2D approach intended for
high local Froude-numbers.
For a given section in the bow and a given
ship speed, the steady wave elevation, s, near the
hull can be plotted as a function of draft, d, as
illustrated in Figure 1. This can be repeated for all
relevant ship speeds. We have assumed zero trim in
the present case. The present Ro-Ro ship is
designed to operate at drafts around 5 to 5.5 meters.

Steady wave elevation [m]

The method is applicable to all wave headings.


Empirical predictions of viscous roll damping due
to bilge keels (Ikeda et al 1978a), skin friction
(Kato 1958, Himeno 1981) and eddy making (Ikeda
et al 1978b,c) are included. Rudders are modeled as
lifting surfaces, and controlled by an autopilot. The
method has been implemented into the program
VERES.

U=16.5 knots
U=17.7 knots
U=20.0 knots

2.5
2.0
1.5
1.0
0.5
0.0
0

6
Draft [m]

10

Figure 1: Steady wave elevation, s, at station 9 as a


function of draft, d, for our 120m case study Ro-Ro
vessel. Results for 3 speeds are included in this sample.

The total draft for the section can now be written,

d tot (t ) = d (t ) + s ( d (t ))
where d(t) is the distance from the keel to the
incident wave and s is the steady wave elevation
near the hull. The relative vertical velocity can be
taken as the time-derivative of the total draft. By
noting that d& (t ) equals the relative velocity, V(t),
we arrive at the following correction to V(t),

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Vtot (t ) = V (t )(1 +

s
)
d

The way we include the steady wave elevation does


not consider the effect of roll motion. The steady
wave elevation is considered identical on each side
of the hull. Finally, the U5-term can be added to
Vtot(t).
Calculation of Slamming Pressures
Slamming pressures are calculated by the
simplified 2D method presented by Zhao et al
(1996). This is a zero-gravity potential theory
formulation similar to the Wagner method, but with
body boundary conditions satisfied on the real body
surface. The method accounts for the important
pile-up of water on each side of the section.
Sections can be asymmetric and have an arbitrary
shape, provided that the beam of the sections
waterline increases monotonically as the section is
being gradually submerged. The section is allowed
to have arbitrary vertical, horizontal and rotational
velocities, but the two latter components are
assumed to be small compared to the first. The
method has been extended by Zhao et al (1997) to
account for separation from places where the
section geometry changes abruptly.
The problem is solved as an initial value
problem. For a number of waterlines, a twodimensional boundary value problem is established
and solved numerically by use of Greens second
identity. The dynamic free surface condition =0 is
applied on the two lines that extend horizontally to
each side from the intersection points between the
piled-up water surface and the body. For an
asymmetric body the pile-up is different on the two
sides, and the two horizontal lines will therefore
have different vertical positions. The body
boundary condition is applied on the exact position
of the wetted body surface below the two lines. A
local solution is used near the body-waterline
intersection point on the right hand side of the
section in order to find the vertical velocity of the
free surface. This gives the amount of time required
for the piled-up water surface to climb from the
present position to the point where the next
waterline intersects with the body. This determines
the time-step and the position of the next waterline
on the left hand side. The pressure on the body is
found from Bernoullis equation with the gravity
term excluded. For each time-step the pressure can
be integrated over a specified area (segment) on the
section. The resulting force is then divided by the

segment size to produce an average pressure. When


this procedure is repeated for each time-step, we
get a time-series of the average pressure on the
segment. The slamming calculation method has
been implemented into the computer code Slam2D.

MODEL TESTS
The calculated results are compared with data from
model tests of a 120-meter car carrier. The service
speed of the vessel is 20 knots in calm water. The
main particulars and the body plan are given in
Table 1 and Figure 2, respectively. The mass-data
are given in Table 2, while hydrostatic properties
are presented in Table 3.

Figure 2: Bodyplan of MV Autoprestige. Dimensions in


[m]. Full-scale. The positions of the two slamming
panels in the bow flare are indicated.

An overview of the instrumentation is given in


Figure 3, and the slamming panels in the bow flare
are depicted in Figure 4. Two slamming panels are
located above each other on port side at station 9 .
The centers of the panels are 4.0 m and 5.3 m
above the calm water level, respectively. The
positions of the panels are indicated in Figure 2.
The local deadrise angle in the area where the
panels are located is about 37. Both panels are
circular with a diameter of 1.45 m, giving an area
of 1.65 m2.
The thin and stiff panels are mounted on a
force transducer. The other end of the transducer is
connected to the bottom of a heavy steel capsule,
which in turn is connected to the hull. A thin,
elastic rubber membrane covers the small gap
between the circular panel and the hull surface.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

When the measured force in the transducer is


divided by the panel area we obtain the average
slamming pressure on the panel. This pressure will
be used in the following.
Table 1: Main particulars of M/V Autoprestige and
model. Scale ratio: 21.62
Parameter
Length over all LOA
Length
between
perpendiculars
Breadth moulded
Displacement in model tests

Value
Fullscale
125.5 m
118.2 m

Value
Modelscale
5.80 m
5.47 m

18.8 m
6902 tons

0.87 m
0.683 tons

propelled and the only physical connections with


the carriage were the cable-bundle and ropes fore
and aft. None of these connections were tight
during the tests, and they are not believed to have
had any significant influence on the vessels
responses.
The model was tested in both regular and
irregular waves, but only regular waves are
considered in this paper.

Table 2: Mass-distribution data. Full-scale values.


Parameter
Total mass m
LCG from AP
VCG from BL
TCG

Value
6902 tons
52.3 m
9.06 m
0m

Parameter
r44
r55
r66
r64

Value
6.55 m
31.1 m
31 m
6m

Table 3: Hydrostatic properties. Full-scale values.


Parameter
Draft FP
Draft AP
Trim angle

Value
5.5 m
5.5 m
0 deg

AP

Parameter
GMt
GMl
Roll period

Value
0.96 m
227 m
15.1 s

Midship cut (x=52.5m) Bow cut (x=109.9m)

Figure 4: The two slamming panels in the bow flare at


station 9 on port side. These are located at x=111.9,
y=4.95, z=9.50 and x=111.9, y=6.70, z=10.8 [m].
FP

RESULTS AND DISCUSSION

MRU

16.1

Some Observations from the Model Tests

Wave tape on port skin


x=114.1m
Vertical wave probe
x=118.4m, y=7.4m

Acc AP
x=0, z=18.7m

10

Acc FP
x=116.7, z=14.3m

Figure 3: Instrumentation of the 120-meter Ro-Ro


vessel. The two slamming panels in the port bow flare
are shown in Figure 4.

Ship motions are measured by an optical system


(and also by an inertial system MRU) and
sampled at 20 Hz (model scale). Relative wave
elevations are sampled at 500 Hz and slamming
forces at 2500 Hz (model scale values).
All tests were performed in the 80m x 50m
Ocean Basin at MARINTEK. The model was self-

The model was tested in regular waves in headings


= 0 (head seas), 22.5, 45, 135, 157.5 and
180. Slamming on the panels was observed only in
headings between 0 and 45 and only for the wave
period that caused heave/pitch resonance, namely
T=9s (/L = 1.06). For this period, various waveheights were tested, and significant slamming
occurred only for wave-heights from 2.5m and up.
The model was tested with the same propeller pitch
and revolution rate in all wave conditions. This rate
corresponds to a speed of 20 knots in calm water.
The speed loss in waves caused the actual speed to
lie between 13 and 18 knots, depending on the
wave condition.
Figure 5 shows pictures from the model tests
in head regular waves with height 5m and period
9s. This condition caused high slamming pressures
in the bow flare. Time-traces of pitch motions,

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

A close-up of the slamming event at t=435 seconds


is shown in Figure 7. We see that the pressure on
the panel is not completely zero between each
slam. In addition to the noise, the pressure
oscillates with the wave encounter frequency.
These oscillations are due to the mass connected to
the force transducer behind the panel. This mass is
accelerated as the bow of the vessel moves up and
down. The corresponding inertia forces are
transferred into the force transducer, which in turn
stretches the thin rubber membrane that connects
the panel and the hull surface. Hence, this is not a
phenomenon that would occur on a real vessel.

Pitch [deg]

4
Experiments
Calculations - nonlinear
Calculations - linear

2
0
-2
-4
400

Relative wave at bow [m]

relative wave elevation at bow and slamming


pressures for the same condition are shown in
Figure 6. The relative wave elevation is defined as
the distance from a point on the hull located 5.5m
above the keel (i.e. at the mean waterline) to the
surface of the incident waves. The relative wave
elevation is positive when this point is submerged.
The conditions in Figure 6 caused the bulb to
emerge about 5m out of water. The relative
velocity between the ship and the water is about 6
m/s as the bottom impacts the water, and it
increases to about 7-7.5 m/s as the mean waterline
passes the water surface. Thereafter the relative
velocity starts to decrease. It was shown by
Hermundstad et al (2004) that the slamming force
induces a large sagging moment in the hull girder
with subsequent vibrations. The vibrations are
damped out before the next slamming incident. In
this case it is therefore reasonable to assume that
the global vibrations from the previous slam can be
disregarded in the slamming pressure calculation.
However, real vessels can have smaller damping,
and vibrations from the previous slam should then
be considered.

410

420

430
time [s]

440

450

-5

-10
400

410

420

430
time [s]

440

450

Pressure - lower panel [kPa]

60
50
40
30
20
10
0

400

410

420

430
time [s]

440

450

Figure 6: Measured and calculated pitch motion, relative


wave elevation at bow and pressure at the lower panel.
Conditions: Regular waves, H=5m, T=9s (/L = 1.06),
= 0 (head seas).
Figure 5: Maximum bow emergence and submergence
during slamming in head regular waves with H=5m and
T=9s.

When the water surface reaches the lower panel,


the load on the panel rises quickly to its peak value.
The duration of the slamming load is about 0.5 s.
After the peak value has been reached, the panels

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

The calculated pressure is also included in Figure


7. The duration and magnitude of the pressure
peaks compare quite well. The numerical model
does not include the hydrostatic pressure, and the
hull girder vibrations are also disregarded. The
behavior after the high slamming peak is therefore
not reproduced by the calculations. The difference
in phase is primarily due to the variation in forward
speed of the model. In the ship motion calculations,
a constant speed was used.

are submerged for 1.5-1.7 s. During this latter


phase, the panels are exposed to an oscillating
pressure with nonzero mean value. The mean value
is caused partly by the hydrostatic pressure and
partly by the added-mass pressure due to the pitch
acceleration. The maximum submergence of the
lower panel is about 2 meters for the present
condition, and we should therefore expect a
hydrostatic pressure that rises steadily to about 20
kPa before it decays to zero. However, the pitch
acceleration is directed upwards during the
slamming event. The associated added mass
pressure is therefore negative, and it counteracts the
hydrostatic pressure. The added-mass pressure also
has an oscillating component that is caused by the
vibration of the hull girder. We see that the
pressure ceases to act immediately after the panels
exit the water. For the upper panel, there is a
secondary peak during water exit.

Comparison of Measured and Calculated Ship


Motions
In the strip theory calculations, 46 sections are used
in the longitudinal direction. Bilge keels, rudders
and viscous roll damping are included in the model.
The time-domain simulations are carried out for 60
s full-scale time, with a time-step of 0.05 s.

Secondary peak during water exit phase


Pressure - upper panel [kPa]

60

Measured and calculated amplitudes of heave,


pitch and relative wave elevation at the bow are
presented in Figure 8 for head waves and in Figure
9 for bow quartering waves. They are plotted to a
base of waveheight. Regular waves with a period of
9 s (/L = 1.06) are used. During the experiments, a
run across the basin started with an acceleration
phase followed by a phase of relatively steady-state
conditions. Finally there was a deceleration phase.

Experiment
Calculations

50

Inertia forces in
transducer caused by WF
motion (pitch/heave)

40
30
20
10
0

430

432

434

436

438

440

time [s]

Sum of hydrostatic pressure, added-mass pressure due


to pitch acceleration and oscillating added-mass
pressure caused by 2-node hull girder vibrations (2Experiment
2.5 Hz)

Pressure - lower panel [kPa]

60

Calculations

50
40
30
20
10
0

430

432

434

436

438

440

time [s]

Encounter period, Te=6s


Bow has started to move upwards relative
to the wave. Panel exits the water.

Figure 7: Close-up the measured slamming pressures at


the upper and lower panels during a severe slam.
Calculated pressures have also been included. Effects of
forward speed and three-dimensional flow are not
included in these calculations (see method Nonlinear
in Figure 12).

When results from the acceleration and


deceleration phases are disregarded, the steadystate time-window contains 7-10 wave encounters.
We have included one black diamond for each
positive and one for each negative amplitude that
were measured. This illustrates the scatter in the
experimental data. Runs for 3 different
waveheights are included for each of the two wave
headings. For H=5m in head seas, the run was
repeated once, and the plot therefore contains more
black diamonds for this condition.
Under ideal conditions, the ship- and wave
motions would have been completely sinusoidal
with constant amplitudes. The black diamonds for
one run would then lie on top of each other. Some
of the observed scatter can be attributed to
variations in the amplitude of the incident waves.
The wave amplitudes in the basin vary within 510%. Transients from the acceleration phase may
also contribute to the scatter. The nonlinear
calculations do not include the effect of slamming
forces on the ship motions. Since the calculations
agree quite well with the experiments, it indicates
that the slamming forces do not influence the ship

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

motions significantly. The time-series of the pitch


motion in Figure 6 is also quite sinusoidal.
However, the slamming forces may still have a
small or moderate influence on the motions, and
some of the observed scatter may be due to
variations in the slamming forces.

Head seas (0 deg)

Bow quartering seas (45 deg)

Experiments
Calculations - nonlinear
Calculations - linear

Heave amplitude [m]

2
Heave amplitude [m]

For comparison, linear strip theory calculations are


also included in Figure 8. Linear predictions do not
display the asymmetry between positive and
negative amplitudes that is clearly present in the
experimental data. The linear calculations give a
too large downward motion of the bow.

2
1

Experiments
Calculations - nonlinear
Calculations - linear

0
-1
-2

-1

2.0

2.5

3.0

3.5
4.0
Waveheight [m]

4.5

5.0

2.0

5.5

2.5

3.0

3.5
4.0
Waveheight [m]

4.5

5.0

5.5

Pitch amplitude [deg]

Pitch amplitude [deg]

2
0

4
2
0
-2

-2
-4
-4
2.5

3.0

3.5
4.0
Waveheight [m]

4.5

5.0

5.5

Relative wave elev. at bow [m]

Relative wave elev. at bow [m]

2.0

-5

2.0

2.5

3.0

3.5
4.0
Waveheight [m]

4.5

5.0

5.5

2.0

2.5

3.0

3.5
4.0
Waveheight [m]

4.5

5.0

5.5

-5

-10
2.0

2.5

3.0

3.5
4.0
Waveheight [m]

4.5

5.0

Figure 8: Measured and calculated amplitudes of heave,


pitch and relative wave elevation at the bow. Wave
period 9 s (/L = 1.06). Head waves. Positive relative
wave elevation means that the water surface is higher
than the calm water draft of 5.5m.

For the higher waves in bow quartering seas, the


bow moves not as deeply into the water as
predicted by the theory. A reason for this
discrepancy may be that we have neglected
slamming forces when calculating the ship
motions. However, in head waves the downward
motion is well predicted also for the highest waves.

5.5

Figure 9: Measured and calculated amplitudes of heave,


pitch and relative wave elevation at the bow. Wave
period 9 s (/L = 1.06). Bow quartering waves. Positive
relative wave elevation means that the water surface is
higher than the calm water draft of 5.5m.

The autopilot that controls the rudders is


programmed to keep the model on the preset
course. The rudder actions will also induce roll
motions. Hence, some roll motions were also
present during tests in head seas, but they were
within +/- 2. The roll motion varied typically
between -4 and +2 in bow quartering waves, with
extremes between -10 and +4. The case with the
largest roll motions is presented in Figure 10.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

In the calculations, there is no acceleration


phase and no response at the roll natural period.
This implies that we are unable to reproduce the
roll motions that were measured during the tests.
We will see later that this influences the slamming
pressures in the flare.

Figure 10: Measured and calculated roll motion and


measured rudder angle. Conditions: Regular waves with
H=5m, T=9s (/L = 1.06), = 45 (bow quartering seas).
Roll towards port side (i.e. towards the waves) is
positive. A positive rudder angle turns the ship towards
starboard; i.e. it gives a negative roll moment.

It is seen in Figure 11 that the roll motion is to


some extent influenced by the slams. The extent of
this influence depends on the roll angle and the roll
velocity during the impact. The influence is mutual
- the slamming pressures at the panels generally
increase for increasing roll angle and roll velocity.
In the experiments, the roll angle and the roll
velocity is negative (i.e. moving away from the
waves) for all slams in the lower waves (H = 2.5m).
In the two highest waves, the roll angle and/or the
roll velocity is positive for some of the slams. In
the calculations, the roll angle changes sign from
positive to negative during the impact for all
waveheights in bow quartering seas. Hence, the roll

Experiment - roll angle


Experiment - press. lower panel

2.5

0.0

30

20

-2.5

10

-5.0

Pressure [kPa]

In Figure 10 there is a large slowly varying


roll motion superimposed on the wave-frequency
response. There is much energy at the natural
period in roll. As the ship model accelerates, it
starts rolling at its natural period, and this also
induces some small yaw motions at the same
period. The rudders also start working at this
period, and they in turn influence the roll motions.
The vessel has two rudders, and both have the same
angle. After band-pass filtering of the measured roll
motion retaining only the energy around the wave
encounter frequency it compares much better
with the calculated roll motion.

velocity is always negative for the slamming


calculations. In the experiments, the ship rolls more
towards the leeward side, while in the calculations,
ship rolls slightly more towards the windward side.

Roll angle [deg]

Negative roll motion means that the vessel rolls


towards the leeward (starboard) side.

-7.5
300

310

320

330
time [s]

340

350

Figure 11: Measured roll motion and slamming pressure


at the lower panel. Conditions: Regular waves with
H=3.8m, T=9s (/L = 1.06), = 45 (bow quartering
seas). Roll towards port side (i.e. towards the waves) is
positive. Slamming panels are located on port side.

Comparison of Measured and Calculated


Slamming Pressures
The sensitivity of the slamming pressures with
respect to wave steepness is illustrated in Figure
12. Experimental results for regular waves of
different heights and with headings 0 and 45 are
included. All waves have a period of 9 s, which
means that they are 6% longer than the ship. One
run across the basin produced 7-10 slams within the
time-window of relatively stable conditions. The
peak value of the panel pressure for each individual
slam is included as a black diamond shape in the
plots. For the highest head wave, results from a
repeated run are also included.
It is seen from the experimental data in Figure
12 that the pressures increase with increasing
waveheight for both wave headings. The pressures
are generally higher at the lower panel, even if the
upper panel has a slightly smaller local deadrise
angle. The reason is that the relative velocity is
higher when the water surface passes the lower
panel. We also observe that the pressures are higher
in head seas than in bow quartering seas. We have
seen that there is some experimental scatter in the
ship-wave relative motions, and the scatter
observed in the measured slamming pressures is
therefore not unexpected.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Table 4: Description of methods presented in Figure 12.


Nonlinear

Description of method
Nonlinear ship motions
Linear ship motions
Sinkage incl. in calculating ship-wave relative motions
U5-term included in calculating relative velocity
Steady wave elevation incl. in calculating relative velocity
Slamming pressures corrected for 3D effects

Name of method (used in legend of Figure 12)


Linear
Nonlin
Nonlin Nonlin
Nonlin
sinkage
sink
sink
sink ueta5
ueta5
ueta5 Zs
Zs 3D
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x

Lower panel
100

Upper panel

Experiments
Nonlinear
Linear
Nonlin sinkage
Nonlin sink ueta5
Nonlin sink ueta5 Zs
Nonlin sink ueta5 Zs 3D

80
60

80
60

0 deg

40

40
20

0 deg (Head seas)

20
0

3
Waveheight [m]

80

60

45 deg (Bow quartering seas)

60
45 deg
40

40

20

20

0
1

3
Waveheight [m]

Figure 12: Measured and calculated slamming pressures [kPa] for lower (left column) and upper (right column) bow flare panels
plotted to a base of waveheight [m]. Regular waves with period 9 s (/L = 1.06). Upper row: 0 (head seas), and lower row: 45
(bow quartering seas). See Table 4 for explanation.

Calculated results obtained with different approaches


are included in Figure 12. The legend names are
described in Table 4.
Since the model tests were performed with a
constant thrust for all waveheights, the ship speed
decreased significantly when the waveheight
increased from 1.3m to 5m. The reduction was largest
for head waves.
Sinkage is included in the ship motion

calculations by forcing the ship to have an increased


draft without changing the mass. It is seen from
Figure 12 that the effect of including sinkage is
negligible for the lower panel, while it increases the
pressures at the upper panel in oblique seas.
Including the U5-term causes a reduction of the
slamming pressures. The effect is quite large for the
upper panel in oblique seas. The reason is that the
pitch angle changes from positive (bow up) to
negative during the slamming events. Since the
slamming pressure reaches the panels at the end of
the event, the pitch angle is always negative then, and

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

it serves to decrease the effective relative velocity at


this stage.
If linear theory is used to calculate the ship motions,
we get much higher slamming pressures for the
highest waves (curve labeled Linear). The calculated
pressure is now significantly higher than the mean of
the experimental pressures, but it compares quite well
with the highest experimental outliers. We saw from
Figure 8 and Figure 9 that linear ship motion theory
gave too large bow submergence for the higher
waves.
By including the steady wave elevation when
the relative velocities are calculated, the slamming
pressures increase significantly. The effect is only
included in a simplified quasi-static manner, as
described above. For the present case, the effect is
most pronounced for the upper panel in oblique seas.
It is seen that the effect of the steady wave elevation
decreases for the higher waves, due to the reduction
in ship speed.
Comparison with the experimental data shows
that the calculated pressures are too high when we
account for the steady wave elevation. This is what
one would expect from a two-dimensional method.
Scolan and Korobkin (2001) studied the water entry
of an elliptic paraboloid and compared the total force
obtained by solving the three-dimensional Wagnerproblem with that obtained by solving a set of twodimensional problems in a strip-wise manner. The
ratio of the forces obtained with these two methods
was presented for aspect ratios between 0 and 1. In
our case, we can use their results (Figure 17 in their
paper) to obtain a very rough estimate of 3D effects.
The forward parts of the waterline contours where the
two panels are located may each be regarded as one
half of an ellipse with aspect ratios 0.62 or 0.76. The
results of Scolan and Korobkin (2001) then indicate
that 3D effects reduce the 2D slamming pressures by
around 30%.
When the 3D correction is done, we see that the
calculated slamming pressures agree well with
experimental data for both panels and for both wave
headings. In bow quartering waves, the calculated
pressures are generally higher than the experimental
pressures, but this can be explained by the inaccuracy
in the prediction of the relative vertical motions and
the roll motions. It was observed in Figure 8 that the
nonlinear strip theory gave a slightly too large bow
submergence in bow quartering waves. It has also
been found that the calculated roll motion was

slightly biased towards the waves, while the


experimental roll motion was clearly biased away
from the waves. Using the calculated roll motion to
provide input to the slamming calculations will
therefore give too small deadrise angles and
subsequently too high slamming pressures. The
relative roll angle when an impact started was
typically 0-2 towards the waves in the calculations.
From the experiments it has been found that a roll
angle of -2 (i.e. away from the waves) is typical
when the bow enters the water in the highest waves.
For the lower waves the relative roll angle at initial
impact is typically around -1. In the lower part of
Figure 13 we have included results obtained using -2
as an initial relative roll angle in the slamming
calculations for bow quartering waves (Curve labeled
Nonlin sink ueta5 Zs 3D r=-2deg). This reduces the
slamming pressures at the upper panel by about 50%,
and the agreement with the experiments is now better.
The calculated pressures now become too low for the
lowest waves, since an initial relative roll angle of -1
would have been better for these cases. This study
shows that the pressures, particularly at the upper
panel, are very sensitive to small variations in the
relative roll angle. Variations in roll motions are
therefore an important reason for the experimental
scatter.
Also in head seas we measured roll angles
within +/-2, while the numerical ship motion
calculations give zero relative roll angle in head
waves. In the upper part of Figure 13 we have
included results for head waves using +2 and -2 as
initial relative roll angle. For the upper panel, most of
the experimental data lie between the curves formed
using these two roll angles. This indicates that the
variation in the relative roll angle is an important
reason for the experimental scatter also in head seas.
To study the sensitivity to speed variations, the
calculations for head seas were repeated with a 5%
lower speed. The results are presented in the upper
part of Figure 13 (Curve labeled Nonl sink ueta5 Zs 3D
95% speed). We see that a 5% reduction in speed
reduces the slamming pressures by typically 10-15%.
The sensitivity to changes in wave heading is studied
in the lower part of Figure 13 for bow quartering
waves. Results for 43 and 47 heading are presented
(Curves labeled Nonlin sink ueta5 Zs 3D 43deg and
Nonlin sink ueta5 Zs 3D 47deg). A change in heading
of 2 is seen to change the slamming pressures by
around 10-20%. Small variations in speed and
heading can therefore also have contributed to the
scatter in the measured slamming pressures.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Lower panel
80

Upper panel

Experiments
Nonlin sink ueta5 Zs 3D
Nonlin sink ueta5 Zs 3D r=-2deg
Nonlin sink ueta5 Zs 3D r=+2deg
Nonl sink ueta5 Zs 3D 95% speed

60

70
60
0 deg

50
40

40

30
0 deg

20

20

10
0

0
1

60

Experiments
Nonlin sink ueta5 Zs
Nonlin sink ueta5 Zs
Nonlin sink ueta5 Zs
Nonlin sink ueta5 Zs

50
40

3D
3D r=-2deg
3D 43deg
3D 47deg

50
40
45 deg

30

30

20

20
45 deg

10
0

10
0

3
Waveheight [m]

3
Waveheight [m]

Figure 13: Measured and calculated slamming pressures [kPa] for lower (left column) and upper (right column) bow flare panels
plotted to a base of waveheight [m]. Regular waves with period 9 s (/L = 1.06). Upper row: 0 (head seas), and lower row: 45
(bow quartering seas). Sensitivity to changes in ship speed, heading and relative roll angle at impact.

The effects of including the steady wave elevation


and correcting for 3D flow will act in opposite
directions. In our calculations the effect of the
steady wave elevation is generally stronger, but for
the highest waves the two effects almost cancel
each other. Both effects have a large influence on
the results, so one should not include only one of
them. One may argue that both effects could be
neglected as a simplified way of calculating
extreme slamming pressures. This may be a useful
approach if the steady wave elevation for various
drafts and speeds is not easily available. The results
obtained with a nonlinear ship motion calculation
where only sinkage is accounted for (Curve labeled
Nonlin sinkage in Figure 12), agree quite well with
the experimental data. A linear ship motion
analysis may also provide useful input for
calculation of extreme slamming pressures (curve
labeled Linear). Results for irregular waves must
be analyzed for a better evaluation of the practical
applicability of the methods.

CONCLUSIONS
Slamming pressures in the bow flare of a 120-meter
Ro-Ro vessel is studied numerically and
experimentally. A nonlinear strip theory method is
used to calculate ship-wave relative motions. The
relative vertical and roll velocities for a slamming
event are given as input to the slamming
calculation program, which is based on a
generalized Wagner formulation and solved by a
2D BEM (Zhao et al 1996). Simulation of a
slamming event for one ship section takes only a
few seconds on a conventional PC.
Various refinements in the calculations are
tested and comparisons are made with experimental
results in regular head and oblique waves. The
following observations are made:

A conventional strip theory with nonlinear


modifications of hydrostatics and FroudeKrylov forces gives a quite accurate prediction
of heave and pitch motions and vertical bow
motion relative to the undisturbed incident

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

waves. If a completely linear theory is used,


the predicted submergence of the bow becomes
too large.
Roll motions must be considered for flare
slamming in bow quartering seas. Slamming
pressures in a flared area are sensitive to small
variations in the roll angle. In oblique waves,
the physical ship model displayed roll motions
at its natural period, and this was not
reproduced by the calculations.
The forward speed of the ship causes a pile-up
of water around the bow, and the pile-up
changes as the bow is being submerged during
a slamming event. A simplified way of
accounting for this pile-up has been presented
in the paper. It is shown that this pile-up
significantly increases the slamming pressure
on the panels in the upper part of the flare,
particularly for the lower waves.
The flow in the bow flare is not twodimensional. By using results from Scolan and
Korobkin (2001) it has been estimated that the
results from the 2D slamming calculations
should be reduced by about 30% to account for
3D effects. When such a correction is made,
the final pressures compare well with the
measured data for various waveheights in head
and bow quartering waves.
The effects of the steady wave elevation and
three-dimensional flow act in opposite
directions. For the highest waves the two
effects almost cancel each other. Quite good
agreement with experiments was obtained
when both effects were neglected. The
practical significance of such a simplification
should be studied for realistic irregular wave
conditions.

ACKNOWLEDGEMENTS
This work has been funded by the Research
Council of Norway and MARINTEK. The model
tests were also supported by UECC.

REFERENCES
Hermundstad, O.A., Aarsnes, J.V. and Moan, T.,
Linear Hydroelastic Analysis of High-Speed
Catamarans and Monohulls, J. Ship Research.,
Vol. 43, No.1, 1999, pp 48-63.
Hermundstad, O.A., Moan, T. and Mrch, H.J.,
Motions and Slamming Loads on a Ro-Ro
Ship, Proc. 9th Int. Symp. Practical Design of
Ships and Other Floating Structures

PRADS04, Luebeck-Travemuende, Germany,


2004.
Himeno, Y., Prediction of ship roll damping
state of the art, Techn. Rep. 239, 1981, Dept.
of Naval Architecture and Marine Engineering,
Univ. Michigan, MI, USA
Ikeda Y., Himeno, Y. and Tanaka, N. On eddy
making component of roll damping force on
naked hull, Techn. Rep. 00403, 1978a, Dep. of
Naval Arch., Univ. Osaka Prefecture, Osaka,
Japan.
Ikeda, Y., Himeno, Y. and Tanaka, N., On roll
damping force of ship - effect of friction of hull
and normal force on bilge keels, Techn. Rep.
00401, 1978b, Dep. of Naval Arch., Univ.
Osaka Prefecture, Osaka, Japan.
Ikeda, Y., Himeno, Y. and Tanaka, N., On roll
damping force of ship - effect of hull surface
pressure created by bilge keels, Techn. Rep.
00402, 1978c, Dep. of Naval Arch., Univ.
Osaka Prefecture, Osaka, Japan.
Kato, H., On the frictional resistance to the rolling
of ships, Journal of Zosen Kiokai, 1958, pp.
102-115.
Muzaferija, S. and Peric, M., Computation of free
surface flows using interface-tracking and
interface-capturing methods. Chap 3. in O.
Mahrenholtz and M. Markiewicz (eds).
Nonlinear Water Wave Interaction, 1998,
Computational
Mechanics
Publications,
Southampton, England
Muzaferija, S., Peric, M., Sames, P. and Schellin,
T., A two-fluid Navier-Stokes solver to
simulate water entry, Proc. 22nd Symp. Naval
Hydrodynamics, Washington DC, USA, 1998,
pp. 277-289.
Ochi, M.K. and Motter, L., Prediction of
slamming characteristics and hull responses for
ship design, Trans. SNAME, Vol. 81, 1973,
pp 144-176.
Sames, P.C., Schellin, T.E., Muzaferija, S. and
Peric, M., 1998, Application of a two-fluid
finite volume method to ship slamming., Proc.
17th Int. Conf. Offshore Mechanics and Arctic
Engineering (OMAE98), Lisbon, Portugal.
Sames, P.C., Kapsenberg, G.K. and Corrignan, P.,
Prediction of bow door loads in extreme wave
conditions, Proc. RINA conf. Design and
Operation for Abnormal Conditions II, London,
UK, 2001, pp 111-123.
Schumann, C., Volume-of-fluid computations of
water entry of bow sections, Proc.
EUROMECH 374, Poitiers, France, 1998.
Scolan, Y.-M. and Korobkin, A.A., Threedimensional theory of water impact. Part 1.
Inverse Wagner problem, J. Fluid Mech.,

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Vol.440, 2001, pp. 293-326.


von Karman, T.,The impact of seaplane floats
during landing, NACA, Techn. note 321, 1929.
Washington DC, USA.
Wagner, H.,Uber stoss- und Gleitvergange an der
Oberflache von Flussigkeiten, Zeitschr.
Angew. Math. Mech., Vol. 12, No. 4, 1932, pp.
193-235.
Wu, M-K, and Moan, T., Linear and nonlinear
hydroelastic analysis of high-speed vessels, J.
Ship Research, Vol.40, No.2, 1996, pp. 149163.
Wu, M-K. and Hermundstad, O.A., Time-domain
simulation of wave-induced nonlinear motions
and loads and its applications in ship design,
Marine Structures, Vol. 15, 2002, pp 561-597.
Zhao, R. and Faltinsen, O.M., Water entry of twodimensional bodies, J. Fluid Mechanics, Vol.
246, 1993, pp. 593-612.
Zhao,R., Faltinsen, O.M. and Aarsnes, J.V., Water
entry of arbitrary two-dimensional sections with
and without flow separation, Proc. 21st Symp.
Naval Hydrodynamics, Trondheim, Norway,
1996, pp. 408-423.
Zhao,R., Faltinsen, O.M. and Haslum, H.A., A
Simplified Nonlinear Analysis of a High-Speed
Planing Craft in Calm Water, Proc. 4th Int.
Conf. on Fast Sea Transportation FAST97,
Sydney, Australia, 1997, pp. 431-438.
Zhao, R. A nonlinear method for predicting wave
resistance of ships with low block coefficients,
Proc. 8th International Conference on
Numerical Ship Hydrodynamics, 2003.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Jinzhu Xia
Australian Maritime College, Australia
It would be interesting to see a comparison
of predicted and measured global slamming force in
addition to the validation on slamming pressure.
Congratulations for an excellent contribution!
AUTHORS REPLY
During the experiments, the vertical shear
force at the bow (just behind station 9) was
measured. We plan to compare these measurements
with calculations and present the results in another
paper.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics


St. Johns, Newfoundland and Labrador, CANADA, 8-13 August 2004

On the Estimation of Torsional Loads


Acting on a Large-Container Ship
Ryuji Miyake, Tingyao Zhu and Atsushi Kumano
(Nippon Kaiji Kyokai, Japan)

ABSTRACT
Comprehensive tank tests to determine ship motions,
hydrodynamic pressures and wave-induced loads in
regular waves of different waveheights, wave angles
and wavelengths were conducted using a recent postPanamax container ship model. The nonlinear
characteristics of the container ship observed in the
ship motions and the wave loads were discussed.
Computations regarding ship motions, hydrodynamic
pressures and wave-induced torsional moments were
performed by a three-dimensional Rankine source
method program and a modified strip method program
in order to mainly verify the validity of the numerical
methods for estimating wave-induced torsional loads in
the design of container ship structures. The validity of
the Rankine source method and the strip method was
investigated in detail and discussed by comparing the
numerical results with the obtained experimental ones.
It was confirmed that the Rankine source method
estimates torsional moment, which is much smaller
than vertical and horizontal bending moments, with
good accuracy.
INTRODUCTION
The size of container ships continues to rise in recent
years. The increased size causes greater challenges in
technology aspects for ship designers. Warping
stresses and deformations of deck structure due to
wave-induced torsional loads, parametric roll, green
sea loadings, bow flare impacts and so on are
important issues that need to be assessed to have a safe
and well functioned ship. As warping stresses and
deformations of deck structure are very large in lager
waves, and they may be the most important items for
total structural strength from design viewpoints, it is
required to improve the accuracy of evaluation of
wave-induced torsional moments. Meanwhile, there are
few instances of experimental study (Nagamoto et al,

1971; Takaishi and Yoshino, 1974; Takezawa et al,


1981) on the wave-induced torsional load of container
ship models. Furthermore, experimental results of
recent large post-Panamax container ships have hardly
been reported.
On the other hand, strip methods are used
extensively to directly estimate wave-induced ship
motions and loads such as vertical and horizontal
bending moments with relative accuracy for practical
purposes recently. However, as they do not consider
the effects of diffraction waves due to the bow and
stern in short wavelength region, it may not be
expected that they can evaluate torsional moments with
good accuracy, because the torsional moment is very
smaller than vertical as well as horizontal bending
moment. Therefore, the authors newly developed a
three-dimensional Rankine source method program
which can consider the three-dimensional effect and is
expected to evaluate torsional moment in linear
frequency domain accurately.
Under these circumstances, the purpose of this
study is to mainly verify the validity of the developed
three-dimensional numerical method program for
estimating wave-induced loads especially the torsional
loads, which are very important in the design of
container ship structures. In order to achieve this, very
comprehensive tank tests using recent post-Panamax
container ship model were firstly carried out in regular
waves for various waveheights, wave angles and
wavelengths.
Additionally,
the
nonlinear
characteristics of the container ship observed in the
ship motions and the wave loads are discussed.
Secondly, computations regarding ship motions, waveinduced moments and hydrodynamic pressures were
performed by the three-dimensional Rankine source
method program and a modified strip method program.
Finally, the validity of the three-dimensional Rankine
source method program is investigated in detail and
discussed by comparing the numerical results with the
obtained experimental ones.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Force transducer

NUMERICAL COMPUTATION
STRIP METHOD
The strip method program used in the computations is
developed based on STFM. Additionally, the twodimensional hydrodynamic forces are computed by the
Close-fit method. Hydrodynamic pressures are
computed by Watanabe's method (1994) using
diffraction potential. Furthermore, modifications (Ito
and Mizoguchi, 1994) are made to the hydrodynamic
pressures based on the reflection of waves in the
shorter wavelength range.

S.S.2.5

S.S.1.75

TANK TEST
Tank tests were conducted at the Nagasaki
Experimental Tank (Length: 160m, width: 30m, depth:
3.5m) of Mitsubishi Heavy Industries. A scale model
was towed with a free motion guide having 6 degrees
of freedom fixed on an XY carriage. The towing point
was the center of gravity of the model. Some springs
were set to add the restoring forces regarding surging,
swaying and yawing motions in the free motion guide
system.
The model used in the tank tests is a typical postPanamax container ship designed by the National
Maritime Research Institute of Japan. Table 1 shows
the principal particulars of the model. The ship model
is separated into four segments being divided at the
cross sections located respectively at the square
stations (S.S.) 2.5, 5.0 and 7.5. Adjacent segments are
connected using a force transducer so that the sectional
forces and moments can be measured at these cross
sections, as shown in Fig. 1. The adjacent segments
have a clearance of about 5mm in order to permit

Pressure gauge

S.S.7.5

S.S.5.5

S.S.8.5

Fig. 1: The arrangement of pressure gauges and force


transducers
Table 1: Principal particulars of the model
Ship
Model

RANKINE SOURCE METHOD


Rankine source method was initially used to predict
wave resistance in steady-state free surface flow on an
advancing ship by Gadd (1976) and Dawson (1977). It
has been expanded to solve unsteady free surface flow
problems. It is now being used to directly estimate ship
motions and loads on an advancing ship in waves in
the linear frequency domain. This method formulates
boundary value problems by integral equations taking a
simple source as core function in frequency domain. In
the Rankine source method program used in the
computations, the upstream finite difference operator is
applied for radiation condition of Kelvin waves
(Dawson, 1977), while Rayleigh viscosity is employed
for the radiation condition of ring waves (Takagi, 1990,
1993) (Yasukawa, 1990). The numerical analysis code
has been developed for the estimation of wave loads by
Miyake et al (1999, 2001).

S.S.5.0

Length (Lpp) (m)

283.8

5.000

Breadth (B)

(m)

42.8

0.754

Draft (d)

(m)

14.0

0.274

GMT

(m)

1.08

0.019

109480 (ton)

584.1 (kgf)

Displacement

S.S.A

Table 2: Weight distribution of the model


S.S.B W(kgf) LCG(m) KG(m) xx/B yy/Lpp

0.0

2.5

116.27 -1.862

0.360

0.375

0.062

2.5

5.0

186.58 -0.634

0.318

0.364

0.057

5.0

7.5

192.69

0.633

0.304

0.351

0.057

7.5

10.0

88.56

1.722

0.303

0.315

0.061

584.10 -0.104

0.320

0.356

0.243

Overall

Table 3: Parameters for tank test


Incident waveheight (m)
3.5, 9.0,(15.0; only for 18.4knots)
0.3, 0.4, 0.5, 0.6, 0.7, 0.8, 0.9
/L
1.0, 1.2, 1.6
Incident wave angle (deg.) 0, 30, 60, 90, 120, 150, 180
Ship speed (knots)

18.4, 24.5

Fig. 2: The device to measure the center of gravity and the


radius of gyration of a segment

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Fig. 3: The panel arrangements on a still water surface around the model (/L=1.0)
Table 3: The computational domain on the still water surface
around the model
Longitudinal domain (x)
-1.5-0.5L <= x <= 2.0+0.5L
Transversal domain (y)

-1.0-0.5B <= y <= 1.0+0.5B

deformations of the force transducer. Each clearance


was made watertight by covering with a vinyl sheet. In
addition, 16 small force transducers each with a plate
of 30mm were attached to the hull located at the
square stations (S.S.) 1.75, 5.5 and 8.5 respectively for
the measurement of hydrodynamic pressures, as shown
in Fig. 1. The force acting on the plate of 30mm was
transformed to pressure per unit area.
Six degrees of motion, two shear forces (vertical
and horizontal shear forces) and three moments
(vertical and horizontal bending moments and torsional
moments) at the three cross sections and hydrodynamic
pressures acting on the hull at 16 locations were
measured. The surging, swaying and heaving motions
were measured with resistance-dynamometers. The
pitching, rolling and yawing motions were measured
with optical-fiber gyros.
Considering that the weight distribution along the
ship length is important for evaluating the torsional
moments, a typical weight distribution of the full load
condition obtained from the actual loading manual was
reproduced in the model. The center of gravity and
radius of gyration of each segment were measured with
a device similar to a swing to which each segment is
set, as shown in Fig. 2. Additionally, ballast weights
were applied to adjust the weight distribution of each
segment. The weight distribution of the model is
shown in Table 2. In the table, "xx/B" and "yy/Lpp"
represent the non-dimensional radius of gyration with
respect to rolling and pitching motions respectively at
the center of gravity of each segment or the overall
model. Furthermore, "LCG" represents the longitudinal
center of gravity from the midship, a positive sign
indicates forward from the midship.
The parameters of the tank tests are shown in
Table 3. The tank tests were carried out in regular

waves for 3 different incident waveheights, 10


different wave lengths, 7 different incident wave
angles from 180 deg. (head sea) to 0 deg. (following
sea) at 30 deg. intervals and 2 different ship speeds.
The tank tests in 15m waveheight could not be carried
out in shorter wave range since such incident waves
break at these waveheights.
COMPARISON BETWEEN NUMERICAL AND
EXPERIMENTAL RESULTS
Numerical computations were performed to investigate
the validity of the Rankine source method and the strip
method as the methods for estimating the torsional
moments on the recent large post-Panamax container
ships. The numerical results were compared with the
experimental ones.
Figure 3 shows the panel on the free surface used
in numerical computations by the Rankine source
method. Although it is preferable to take as large a
computational domain on the free surface as possible,
in practice, this domain is limited by the memory
capacity of the computer. Therefore, the computational
domain was varied by varying the wavelength, as
shown in Table 3, after considering the wave spread
with respect to wavelength within the scope of the
memory capacity of the computer.
SHIP MOTIONS
Figures 4 (a), (b), (c) and (d) show some comparisons
of the numerical and experimental results of the ship
motions (Sway, Heave, Roll, Pitch) which have an
effect on the torsional moments.
The horizontal axes in all the figures indicate the
wavelength normalized by ship's length L (/L),
while the vertical ones show the amplitude of motions
per unit wave amplitude (A) or wave slope (kA). The
experimental results were analyzed by Fourier
decomposition, and thus, the experimental values show
the amplitudes of first order components having the
same period as the encountering wave period. In the
figure, "Exp.(3.5m)", "Exp.(9.0m)", "Exp.(15.0m)",

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

1.50

1.50

Sway, =30deg.
Vs=18.4knot
Exp.(3.5m)
Exp.(9.0m)
Exp.(15.0m)
STRIP
RANKINE

0.75

RANKINE

0.25

0.25

0.25

0.00
1
/L

1.5

STRIP

0.75
0.50

0.5

Exp.(15.0m)

0.75
0.50

Exp.(9.0m)

1.00

0.50

0.00

Exp.(3.5m)

Sway, =120deg.
Vs=18.4knot

1.25

1.00
Sway/

1.00
Sway/

1.25

Sway/

1.25

1.50
Exp.(3.5m)
Exp.(9.0m)
Exp.(15.0m)
STRIP
RANKINE

Sway, =60deg.
Vs=18.4knot

0.00
0

0.5

1
/L

1.5

0.5

1
/L

1.5

(a) Swaying motion amplitude (=30, 60 and 120deg.)


1.50

1.50

Heave, =30deg.
Vs=18.4knot
Exp.(3.5m)
Exp.(9.0m)
Exp.(15.0m)
STRIP
RANKINE

0.75

Heave, =120deg.
Vs=18.4knot

1.25

1.00

1.00
Heave/

1.00
Heave/

1.25

Heave/

1.25

1.50
Exp.(3.5m)
Exp.(9.0m)
Exp.(15.0m)
STRIP
RANKINE

Heave, =60deg.
Vs=18.4knot

0.75

0.75

0.50

0.50

0.50

0.25

0.25

0.25

0.00

0.00

0.00

Exp.(3.5m)
Exp.(9.0m)
Exp.(15.0m)
STRIP
RANKINE

0.5

1
/L

1.5

0.5

1
/L

1.5

0.5

1
/L

1.5

(b) Heaving motion amplitude (=30, 60 and 120deg.)


3.00

3.00

Roll, =30deg.
Vs=18.4knot
Exp.(3.5m)
Exp.(9.0m)
Exp.(15.0m)
STRIP
RANKINE

1.50

RANKINE

0.50

0.50

0.50

0.00

1
/L

1.5

STRIP

1.50
1.00

0.5

Exp.(15.0m)

1.50
1.00

Exp.(9.0m)

2.00

1.00

0.00

Exp.(3.5m)

Roll, =120deg.
Vs=18.4knot

2.50

2.00
Roll/k

2.00
Roll/k

2.50

Roll/k

2.50

3.00
Exp.(3.5m)
Exp.(9.0m)
Exp.(15.0m)
STRIP
RANKINE

Roll, =60deg.
Vs=18.4knot

0.00

0.5

1
/L

1.5

0.5

1
/L

1.5

(c) Rolling motion amplitude (=30, 60 and 120deg.)


1.50

1.50

Pitch, =30deg.
Vs=18.4knot
Exp.(3.5m)
Exp.(9.0m)
Exp.(15.0m)
STRIP
RANKINE

0.75

RANKINE

0.25

0.25

0.25

0.00

1
/L

1.5

STRIP

0.75
0.50

0.5

Exp.(15.0m)

0.75
0.50

Exp.(9.0m)

1.00

0.50

0.00

Exp.(3.5m)

Pitch, =120deg.
Vs=18.4knot

1.25

1.00
Pitch/k

1.00
Pitch/k

1.25

Pitch/k

1.25

1.50
Exp.(3.5m)
Exp.(9.0m)
Exp.(15.0m)
STRIP
RANKINE

Pitch, =60deg.
Vs=18.4knot

0.00

0.5

1
/L

1.5

0.5

1
/L

(d) Pitching motion amplitude (=30, 60 and 120deg.)


Fig. 4: Comparison of experimental results and numerical results on the motion amplitude (Vs=18.4knots)

Copyright National Academy of Sciences. All rights reserved.

1.5

Twenty-Fifth Symposium on Naval Hydrodynamics

"STRIP" and "Rankine" represent the experimental


results in waves with three different incident
waveheights of the real scale, and the numerical results
by the strip method and the Rankine source method,
respectively.
The
results
for
"Exp.(3.5m)",
"Exp.(9.0m)" and "Exp.(15.0m)" should coincide with
each other if the phenomena were of linear ones.
Additionally, it is desirable that the results of both the
numerical methods are compared with the results of
3.5m waveheight in tank tests, since these numerical
methods employ liner theory.
For computing the non-linear viscous damping
forces against the rolling motion, extinction
coefficients obtained by a free rolling test with a
forward velocity were applied to both the numerical
methods.
As for the rolling motion, the nonlinear effect due
to waveheight is observed in the quartering sea
(30deg.) since the amplitudes normalized by the
incident-wave amplitude differ significantly among
those obtained in the three waveheights. The natural
period of the rolling motion lies in the quartering sea
(30deg.) since GMT of container ships is smaller than
those of tankers and bulk carriers generally. Therefore,
the waveheight effect around the natural period of the
rolling motion most likely attributed to the nonlinear
viscous-damping characteristics that are known to be
manifested in large rolling motion.
Moreover, as for heaving motion, it is also noticed
from the figure that the response per unit encountering
wave amplitude becomes slightly smaller as the
waveheight increases around the relative wavelength of
/L=0.8 in the bow sea (120deg.), where the heaving
motions have the natural period. Such nonlinear
characteristics were also confirmed in the experiments
of the VLCC model (Zhu et al 2002) and the Panamax
container ship model (Miyake et al, 2000). For this
reason, the numerical results of heaving motions in the
natural period by both the numerical methods are
slightly larger than experimental ones.
On the whole, the results of the motions by both
the numerical methods agreed well with the
experimental results, as shown in Fig. 4.
HYDRODYNAMIC PRESSURE
Figures 5, 6 and 7 show some comparisons of the
numerical and experimental results of the pressures on
the hull surface located around the cross section at the
square stations (S.S.) 1.75, 5.5 and 8.5 (see Fig. 1) in
the bow and quartering seas respectively. The
horizontal axes in all the figures indicate the
wavelength normalized by ship's length L (/L),
while the vertical ones show the amplitude of pressures
per unit wave amplitude (A). and g in the vertical

axes represent the fluid density and the gravity


acceleration respectively. If the pressure gauge is
located at just below the calm water surface, the time
history of the pressure is truncated at a certain pressure
level, which may correspond to the atmospheric
pressure. This is a well-known characteristic of the
pressure on the hull near the calm water surface, which
may come out of the water into the atmosphere as a
large displacement is induced in waves. Therefore, the
experimental results in those cases shown in Fig. 6 are
dealt with by the method proposed by Tanizawa et al
(1993).
In the figure, "11mW.L." and "Bottom X=10m"
represent the vertical location (11m above the keel line
or 3m below the calm water surface) of the pressure
gauge and the horizontal location (10m to the side from
the center line) of the pressure gauge on the bottom of
hull (see Fig. 1).
As shown in Fig. 5, the nonlinearities with respect
to waveheight are observed in the experimental
pressures at the locations except for the bottom of hull
at the three cross sections in the quartering sea (30deg.),
in which the pressure per unit incident-wave amplitude
becomes smaller as the waveheight increases,
especially the nonlinearities are larger at the locations
on the weather side than at those on the lee side. These
nonlinearities are most likely caused by the
nonlinearities of the rolling motions, because this wave
condition is the same as the nonlinearities that appear
in the rolling motions shown in Fig. 4 (c).
On the other hand, in the quartering sea (60deg.),
the nonlinearities are hardly confirmed from the
experimental results except for those near the waterline
on the weather side in the shorter wavelength range, as
shown in Fig. 6. On the other hand the nonlinearities
are clearly observed in the experimental results in the
bow sea (120deg.), in which the pressure per unit
incident-wave amplitude becomes smaller as the
waveheight increases, especially at the bow cross
section (S.S. 8.5), as shown in Fig. 7. In case of the
bow sea (120deg.), the pressures on the bottom of hull
at the bow cross section (S.S. 8.5) have the
nonlinearities due to waveheight around the relative
wavelength of /L=0.8, although it has been well
known that the scattering of waves is relatively small
on the bottom of hull. These nonlinearities around the
wavelength of /L=0.8 in the bow sea (120deg.) are
probably caused by the nonlinearities of the heaving
motions as shown in Fig. 4 (b). Furthermore, the
nonlinearities near the waterline on the weather side in
the shorter wavelength range of the quartering and bow
seas (60, 120deg.) probably arise from the
nonlinearities of diffraction waves when the ship
encounters waves of relatively shorter wavelength,
since the nonlinearities of the ship motions are hardly

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

3.5

3.5

=30deg., S.S.1.75, Weather Side, 11mW.L.


Vs=18.4knot
Exp.(3.5m)

=30deg., S.S.1.75, Center of Bottom


Vs=18.4knot

3.0

=30deg., S.S.1.75, Lee Side, 11mW.L.


Vs=18.4knot

2.5

Exp.(3.5m)

2.5

Exp.(3.5m)

Exp.(15.0m)
STRIP
RANKINE

1.5

Exp.(9.0m)
2.0

P A/ g

2.0

P A/ g

Exp.(9.0m)

1.5

Exp.(9.0m)

Exp.(15.0m)
STRIP
RANKINE

2.5

3.5

3.0

2.0

P A/ g

3.0

1.5

1.0

1.0

1.0

0.5

0.5

0.5

0.0

0.0
0

0.5

1
/L

1.5

Exp.(15.0m)
STRIP
RANKINE

0.0
0

0.5

1
/L

1.5

0.5

1
/L

1.5

(a) S.S. 1.75


3.5

3.5

=30deg., S.S.5.5, Weather Side, 11mW.L.


Vs=18.4knot

2.5

Exp.(3.5m)

Exp.(3.5m)

Exp.(15.0m)
STRIP
RANKINE

1.5

=30deg., S.S.5.5, Lee Side, 11mW.L.


Vs=18.4knot

2.5

Exp.(3.5m)

Exp.(9.0m)
2.0

P A/ g

2.0

P A/ g

Exp.(9.0m)

3.0

1.5

Exp.(9.0m)

Exp.(15.0m)
STRIP
RANKINE

2.5

3.0

2.0

P A/ g

3.0

3.5

=30deg., S.S.5.5, Weather Side


Bottom X=10m, Vs=18.4knot

1.5

1.0

1.0

1.0

0.5

0.5

0.5

0.0

0.0
0

0.5

1
/L

1.5

Exp.(15.0m)
STRIP
RANKINE

0.0
0

0.5

1
/L

1.5

0.5

1
/L

1.5

(b) S.S. 5.5


3.5

3.5

=30deg., S.S.8.5, Weather Side, 11mW.L.


Vs=18.4knot

2.5

Exp.(3.5m)

Exp.(3.5m)

Exp.(15.0m)
STRIP
RANKINE

1.5

=30deg., S.S.8.5, Lee Side, 11mW.L.


Vs=18.4knot

2.5

Exp.(3.5m)

Exp.(9.0m)
2.0

P A/ g

P A/ g

Exp.(9.0m)
2.0

3.0

1.5

Exp.(9.0m)

Exp.(15.0m)
STRIP
RANKINE

2.5

3.0

2.0

P A/ g

3.0

3.5

=30deg., S.S.8.5, Center of Bottom


Vs=18.4knot

1.5

1.0

1.0

1.0

0.5

0.5

0.5

0.0

0.0
0

0.5

1
/L

1.5

Exp.(15.0m)
STRIP
RANKINE

0.0
0

0.5

1
/L

1.5

0.5

1
/L

1.5

(c) S.S. 8.5


Fig. 5: Comparison of experimental results and numerical results on the hydrodynamic pressure amplitude in the quartering sea
(=30deg., Vs=18.4knots)

observed in these domains, as shown in Fig. 4.


In the quartering sea (30deg.), the numerical
results obtained by both the methods fairly agree with
the experimental ones in 3.5m waveheight (see Fig. 5).
Additionally, in case of the quartering sea (60deg.), the
results obtained by both the numerical methods show
good agreement with the experimental ones in 3.5m
waveheight (see Fig. 6). On the other hand, in the bow
sea (120deg.), the numerical results by the Rankine
source method show good coincidence with the
experimental ones in 3.5m waveheight except for those

at the bow cross section (S.S. 8.5), as shown in Fig. 7.


It is difficult for the pressures around the bow in the
head and bow seas to be estimated by the numerical
method with good accuracy, especially in case when
response becomes larger since the scattering of waves
becomes larger. As for the strip method, the numerical
results do not agree well with the experimental ones at
the bow cross section (S.S. 8.5) as well as the midship
cross section (S.S. 5.5), especially around the relative
wavelength of /L=0.8 of the bow sea (120deg.). This
is because the results of the heaving motions obtained

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

3.5

3.5

=60deg., S.S.1.75, Weather Side, 11mW.L.


Vs=18.4knot

2.5

Exp.(3.5m)

Exp.(3.5m)

Exp.(15.0m)
STRIP
RANKINE

1.5

=60deg., S.S.1.75, Lee Side, 11mW.L.


Vs=18.4knot

2.5

Exp.(3.5m)

Exp.(9.0m)
2.0

P A/ g

2.0

P A/ g

Exp.(9.0m)

3.0

1.5

Exp.(9.0m)

Exp.(15.0m)
STRIP
RANKINE

2.5

3.5

=60deg., S.S.1.75, Center of Bottom


Vs=18.4knot

3.0

2.0

P A/ g

3.0

1.5

1.0

1.0

1.0

0.5

0.5

0.5

0.0

0.0
0

0.5

1
/L

1.5

Exp.(15.0m)
STRIP
RANKINE

0.0
0

0.5

1
/L

1.5

0.5

1
/L

1.5

(a) S.S. 1.75


3.5

3.5

=60deg., S.S.5.5, Weather Side, 11mW.L.


Vs=18.4knot

2.5

Exp.(3.5m)

Exp.(3.5m)

Exp.(15.0m)
STRIP
RANKINE

1.5

=60deg., S.S.5.5, Lee Side, 11mW.L.


Vs=18.4knot

2.5

Exp.(3.5m)

Exp.(9.0m)
2.0

P A/ g

2.0

P A/ g

Exp.(9.0m)

3.0

1.5

Exp.(9.0m)

Exp.(15.0m)
STRIP
RANKINE

2.5

3.0

2.0

P A/ g

3.0

3.5

=60deg., S.S.5.5, Weather Side


Bottom X=10m, Vs=18.4knot

1.5

1.0

1.0

1.0

0.5

0.5

0.5

0.0

0.0
0

0.5

1
/L

1.5

Exp.(15.0m)
STRIP
RANKINE

0.0
0

0.5

1
/L

1.5

0.5

1
/L

1.5

(b) S.S. 5.5


3.5

3.5

=60deg., S.S.8.5, Weather Side, 11mW.L.


Vs=18.4knot

2.5

Exp.(3.5m)

Exp.(3.5m)

Exp.(15.0m)
STRIP
RANKINE

1.5

=60deg., S.S.8.5, Lee Side, 11mW.L.


Vs=18.4knot

2.5

Exp.(3.5m)

Exp.(9.0m)
2.0

P A/ g

P A/ g

Exp.(9.0m)
2.0

3.0

1.5

Exp.(9.0m)

Exp.(15.0m)
STRIP
RANKINE

2.5

3.0

2.0

P A/ g

3.0

3.5

=60deg., S.S.8.5, Center of Bottom


Vs=18.4knot

1.5

1.0

1.0

1.0

0.5

0.5

0.5

0.0

0.0
0

0.5

1
/L

1.5

Exp.(15.0m)
STRIP
RANKINE

0.0
0

0.5

1
/L

1.5

0.5

1
/L

1.5

(c) S.S. 8.5


Fig. 6: Comparison of experimental results and numerical results on the hydrodynamic pressure amplitude in the quartering sea
(=60deg., Vs=18.4knots)

by the strip method are a little larger than those


obtained by tank tests and Rankine source method in
this wavelength of the bow sea (120deg.).
TORSIONAL MOMENT
Figures 8, 9 and 10 show some comparisons of the
numerical and experimental results of the torsional
moments at the three cross sections (S.S.2.5, 5.0, 7.5)
in the bow and quartering seas respectively.

The horizontal axes in all the figures indicate the


wavelength normalized by ship's length L (/L),
while the vertical ones show the amplitude of torsional
moment per unit wave amplitude (A). The
experimental results were analyzed by Fourier
decomposition, and thus, similar to the previous
motions, the experimental values show the amplitudes
of first order components having the same period as the
encountering wave period. The torsional moments
were computed about the vertical location of the force
transducer, or slightly below the still-water level.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

3.5

3.5

3.5

=120deg., S.S.1.75, Weather Side, 11mW.L.


Vs=18.4knot
Exp.(3.5m)

=120deg., S.S.1.75, Center of Bottom


Vs=18.4knot

3.0

2.5

Exp.(3.5m)

2.5

Exp.(15.0m)
STRIP
RANKINE

1.5

2.0
1.5

STRIP
RANKINE

2.0
1.5

1.0

1.0

0.5

0.5

0.5

0.0
1
/L

1.5

Exp.(15.0m)
STRIP
RANKINE

0.0

0.0
0.5

Exp.(3.5m)
Exp.(9.0m)

Exp.(15.0m)

1.0

=120deg., S.S.1.75, Lee Side, 11mW.L.


Vs=18.4knot

Exp.(9.0m)
P A/ g

2.0

P A/ g

Exp.(9.0m)

2.5

3.0

P A/ g

3.0

0.5

1
/L

1.5

0.5

1
/L

1.5

(a) S.S. 1.75


3.5

=120deg., S.S.5.5, Weather Side, 11mW.L.


Vs=18.4knot

2.5

3.0
2.5

Exp.(3.5m)

Exp.(3.5m)

STRIP
RANKINE

Exp.(15.0m)

2.0

P A/ g

P A/ g

1.5

=120deg., S.S.5.5, Lee Side, 11mW.L.


Vs=18.4knot

2.5

Exp.(3.5m)

Exp.(9.0m)

Exp.(9.0m)
2.0

3.0

1.5

Exp.(9.0m)

Exp.(15.0m)
STRIP
RANKINE

3.0

3.5

=120deg., S.S.5.5, Wether Side


Bottom X=10m, Vs=18.4knot

2.0

P A/ g

3.5

1.5

1.0

1.0

1.0

0.5

0.5

0.5

0.0

0.0
0

0.5

1
/L

1.5

Exp.(15.0m)
STRIP
RANKINE

0.0
0

0.5

1
/L

1.5

0.5

1
/L

1.5

(b) S.S. 5.5


3.5

3.5

=120deg., S.S.8.5, Weather Side, 11mW.L.


Vs=18.4knot

2.5

3.0

3.0

=120deg., S.S.8.5, Lee Side, 11mW.L.


Vs=18.4knot

Exp.(3.5m)

2.5

Exp.(3.5m)

2.5

Exp.(3.5m)
A

2.0

Exp.(15.0m)
STRIP

1.5

Exp.(9.0m)

Exp.(9.0m)

P A/ g

P A/ g

Exp.(9.0m)

RANKINE

2.0

Exp.(15.0m)
STRIP

1.5

RANKINE

3.0

=120deg., S.S.8.5, Center of Bottom.


Vs=18.4knot

2.0

P A/ g

3.5

1.5

1.0

1.0

1.0

0.5

0.5

0.5

0.0
0.5

1
/L

1.5

STRIP
RANKINE

0.0

0.0

Exp.(15.0m)

0.5

1
/L

1.5

0.5

1
/L

1.5

(c) S.S. 8.5


Fig. 7: Comparison of experimental results and numerical results on the hydrodynamic pressure amplitude in the bow sea
(=120deg., Vs=18.4knots)

It has been well known that it is difficult to


measure torsional moments by tank tests as well as to
estimate them by numerical methods with good
accuracy, since torsional moments are very small
compared to the vertical/horizontal bending moments.
Therefore, the torsional moments rather than the other
moments may be affected by the accuracy of weight
distribution, and a typical weight distribution of the full
load condition obtained from the actual loading manual
was reproduced in the model as shown in Table 2. This
implies that the lateral motions such as roll and sway

may be predominant components in the torsional


moments. Accordingly, it is predicted that the torsional
moments become larger in the quartering sea in which
the lateral motions become larger (see Fig. 4). On the
contrary, it is confirmed from all the figures that the
experimental results in the shorter wavelength range of
the bow sea (120deg.) are the largest at all the cross
sections, while the values at the stern cross section
(S.S.2.5) are larger than those at the other cross
sections (S.S.5.0, 7.5), as shown in Fig. 10. Hence, as
for the torsional moments, the hydrodynamic forces

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

0.005

0.004

0.002

0.001

Exp.(3.5m)
Exp.(9.0m)
Exp.(15.0m)
STRIP
RANKINE

0.003

0.002

0.001

0.000
0.5

1
/L

1.5

0.002

0.000
0

Exp.(3.5m)
Exp.(9.0m)
Exp.(15.0m)
STRIP
RANKINE

0.003

0.001

0.000
0

=30deg., S.S.7.5
Vs=18.4knot

0.004

Tx/ gL B

Exp.(3.5m)
Exp.(9.0m)
Exp.(15.0m)
STRIP
RANKINE

0.003

Tx/ gL B

0.004

0.005

=30deg., S.S.5.0
Vs=18.4knot

=30deg., S.S.2.5
Vs=18.4knot

Tx/ gL B

0.005

0.5

1
/L

1.5

0.5

1
/L

1.5

(a) S.S.2.5
(b) S.S.5.0
(c) S.S.7.5
Fig. 8: Comparison of experimental results and numerical results on the torsional moment amplitude in the quartering sea
(=30deg., Vs=18.4knots)
0.005

0.004

0.002

0.001

0.003

0.002

Exp.(3.5m)
Exp.(9.0m)
Exp.(15.0m)
STRIP
RANKINE

0.001

0.000
0.5

1
/L

1.5

0.003

0.002

0.000
0

Exp.(3.5m)
Exp.(9.0m)
Exp.(15.0m)
STRIP
RANKINE

0.001

0.000
0

=60deg., S.S.7.5
Vs=18.4knot

0.004

Tx/ gL B

0.003

Exp.(3.5m)
Exp.(9.0m)
Exp.(15.0m)
STRIP
RANKINE

Tx/ gL B

0.004

0.005

=60deg., S.S.5.0
Vs=18.4knot

=60deg., S.S.2.5
Vs=18.4knot

Tx/ gL B

0.005

0.5

1
/L

1.5

0.5

1
/L

1.5

(a) S.S.2.5
(b) S.S.5.0
(c) S.S.7.5
Fig. 9: Comparison of experimental results and numerical results on the torsional moment amplitude in the quartering sea
(=60deg., Vs=18.4knots)
0.005

0.004

Exp.(3.5m)
Exp.(9.0m)
Exp.(15.0m)
STRIP
RANKINE

0.003

0.002

0.001

0.001

0.5

1
/L

1.5

Exp.(3.5m)
Exp.(9.0m)
Exp.(15.0m)
STRIP
RANKINE

0.003

0.002

0.001

0.000

0.000

=120deg., S.S.7.5
Vs=18.4knot

0.004

Tx/ gL B

0.002

Exp.(3.5m)
Exp.(9.0m)
Exp.(15.0m)
STRIP
RANKINE

0.003

Tx/ gL B

0.004

0.005

=120deg., S.S.5.0
Vs=18.4knot

=120deg., S.S.2.5
Vs=18.4knot

Tx/ gL B

0.005

0.000
0

0.5

1
/L

1.5

0.5

1
/L

1.5

(a) S.S.2.5
(b) S.S.5.0
(c) S.S.7.5
Fig. 10: Comparison of experimental results and numerical results on the torsional moment amplitude in the bow sea
(=120deg., Vs=18.4knots)

rather than the inertia forces have a great influence


since the lateral motions are apparently small in the
shorter wavelength range of the bow sea (120deg.).
This trend also accords with a trend that the warping
stresses become larger around in front of the engine
room (about S.S.2.5) due to the structural configuration
of container ships.

As shown in Fig. 10, the nonlinearities are clearly


observed in the torsional moments at the bow cross
section (S.S.7.5) in the middle wavelength range of the
bow sea (120deg.), in which the torsional moment per
unit incident-wave amplitude becomes larger as the
waveheight increases, whereas the three values in
waves of different waveheights are almost identical

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

with the torsional moments at the stern and midship


cross sections (S.S.2.5, 5.0) which implies very weak
nonlinearity. This is due to impact loads acting on the
bow flare of the model because the impact load
becomes larger as the waveheight increases in this
wavelength range. On the other hand, in case of the
quartering seas (30, 60deg.), the nonlinearities are also
shown in the torsional moments at all the cross sections.
This is mainly attributed to the nonlinearities of the
rolling motions in these wave angles.
The numerical results obtained by the Rankine
source method show good agreement with the
experimental ones in 3.5m waveheight in all figures,
especially in the shorter wavelength range, whereas the
numerical results obtained by the strip method do not
always agree with the experimental ones, especially at
the midship cross section (S.S. 5.0) in the quartering
and bow seas (60, 120deg.) as shown in Figs. 9 (b) and
10 (b).
It is noticed from Figs. 6 and 7 that the pressures
obtained by the strip method almost agree with
experimental results and the results by the Rankine
source method in these wave angles expect for the
cases slightly below waterline close to the midship
cross section (S.S. 5.5) in the bow sea (120deg.) as
shown in Fig. 7 (b). For this reason, in case of the bow
sea (120deg.), the accuracy with respect to the
torsional moments by the strip method is probably
affected by the accuracy with respect to the pressures.
Similarly, the strip method does not estimate the
torsional moments with good accuracy at the midship
cross section (S.S. 5.0) of the quartering sea (60deg.).
As a result, it could be concluded that it is difficult to
use the strip method to estimate the torsional moments,
which are very smaller than the vertical/horizontal
bending moments, with good accuracy.
CONCLUSION
Comprehensive tank tests to determine ship motions,
hydrodynamic pressures and wave-induced loads in
regular waves of different waveheights, wave angles
and wavelengths were conducted using a recent postPanamax container ship model. Some of the results
were compared with those obtained by a Rankine
source method and a strip method. The findings of the
study can be summarized as follows.
The followings may be concluded from these studies:
1) The ship motions are predicted fairly well by both
the Rankine source method and the strip method.
2) The Rankine source method estimates the
pressures except for those at the bow cross section in
the bow sea with good accuracy. On the other hand,
the strip method estimates the pressures except for

those at the bow cross section as well as the midship


cross section in the bow sea fairly well.
3) The measured torsional moments are the largest in
the shorter wavelength range of the bow sea, while
the values at the stern cross section are larger than
those at the midship and bow cross sections.
4) Nonlinearities with respect to waveheight are
clearly observed in such a way that the torsional
moment per unit incident-wave amplitude becomes
larger as the waveheight increases. These nonlinear
characteristics are especially manifested at the bow
cross section in the bow sea, whereas are hardly
manifested at the stern and midship cross sections.
5) The Rankine source method estimates the torsional
moments with good accuracy. On the other hand, the
strip method does not always estimate the torsional
moments well, especially at the midship cross section
in the quartering and bow seas.
Based on the above findings, it can be concluded
that the three-dimensional numerical method such as
the Rankine source method, which can properly
consider the three-dimensional effect, is an effective
method for estimating torsional moments in the design
of large container ship structures. However, problems
such as the memory capacity of computers and the
excessive computation time are to be solved in order to
make the Rankine source method more practical.
Nevertheless, the Rankine source method may replace
the strip method as the effective method for estimating
wave-induced loads in the near future.
ACKNOWLEDGEMENT
The tank tests have been carried out as part of the joint
research project among Mitsubishi Heavy Industries,
the National Maritime Research Institute of Japan and
Nippon Kaiji Kyokai. The authors wish to gratefully
acknowledge the support and valuable comments of Dr
Katsuji Tanizawa of the National Maritime Research
Institute of Japan, and Prof. Hironori Yasukawa of
Hiroshima University during the course of conducting
the tank tests. The authors also thank Mr. Yuji
Fukushima of Mitsubishi Heavy Industries, who took
charge of the tank tests.
REFERENCES
Nagamoto, R., Konuma, M., Iizuka, M., Aoki, M. and
Takahashi, T., "Theoretical calculation of lateral shear
force, lateral bending moment and torsional moment
acting on the ship hull among waves", Journal of the
Society of Naval Architects of Japan, Vol.132, 1972,
pp. 257-268 (in Japanese).

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Takaishi, Y. and Yoshino, T., "Midship Bending and


Torsional Moments of a Container Ship in Oblique
Waves", Journal of the Kansai Society of Naval
Architects, No. 152, 1974, pp. 79-91 (in Japanese).
Takezawa, S., Kobayashi, K. and Obokata, J., "Towing
Model Tests on Ship Motions and Wave Loads in
Oblique Waves", Journal of the Society of Naval
Architects of Japan, Vol.149, 1981, pp. 37-46 (in
Japanese).
Watanabe, I., "Practical method for Diffraction
Pressure on a Ship Running in Oblique Wave", Journal
of the Kansai Society of Naval Architects, No. 221,
1994, pp. 83-89 (in Japanese).
Ito, A. and Mizoguchi, S., "Hydrodynamic Pressure
Acting on a Full Ship in Oblique Short Waves",
Journal of the Kansai Society of Naval Architects, No.
222, 1994, pp. 125-132 (in Japanese).

of the Society of Naval Architects of Japan, Vol. 190,


2001, pp. 107-119 (in Japanese).
Miyake, R., Kinoshita, T., Kagemoto, H. and Zhu, T.,
"Ship motions and loads in large waves", Proceedings
of 23rd Symposium on Naval Hydrodynamics, Val de
Reuil, France, 2000, pp.48-61.
Zhu, T., Kumano, A., Shigemi, T. and Matsunami, R.,
"Consideration of Wave-Induced Loads for Direct
Strength Calculation under Extreme Waves",
Proceeding of 21st International Conference on
Offshore Mechanics and Arctic Engineering, Oslo,
Norway, 2002, OMAE02-28234
Tanizawa, K., Taguchi, H., Saruta, T. and Watanabe, I.,
"Experimental Study of Wave Pressure on VLCC
Running in Short Waves", Journal of the Society of
Naval Architecture of Japan, Vol. 174, 1993, pp. 233242 (in Japanese).

Gadd, GE., "A method of computing the flow and


surface wave pattern around full forms", Trans. Royal
Institution of Naval Architects, Vol. 18, 1976, pp. 207219.
Dawson, C.W., "A practical computer method for
solving ship-wave problems", Proc. 2nd Intl. Conf. on
Numerical Ship Hydrodynamics, Berkley, 1977, pp.
30-38.
Takagi, K., "An Application of Rankine Source
Method for Unsteady Free Surface Flows", Journal of
the Kansai Society of Naval Architects, No. 213, 1990,
pp. 21-29 (in Japanese).
Takagi, K., "Calculation of Unsteady Pressure by
Rankine Source Method", Journal of the Kansai
Society of Naval Architects, No. 219, 1993, pp. 47-56
(in Japanese).
Yasukawa, H., "A Rankine panel method to calculate
unsteady ship hydrodynamic forces", Journal of the
Society of Naval Architects of Japan, Vol. 168, 1990,
pp. 131-140.
Miyake, R., Kagemoto, H. and Fujino, M.,
"Calculation of hydrodynamic forces acting on a ship
in waves by Rankine source method", Journal of the
Society of Naval Architects of Japan, Vol. 185, 1999,
pp. 49-60 (in Japanese).
Miyake, R., Zhu, T. and Kagemoto, H., "On the
Estimation of Wave-induced Loads Acting on Practical
Merchant Ships by a Rankine source Method", Journal

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics


St. Johns, Newfoundland and Labrador, CANADA, 8-13 August 2004

H2O: Hierarchical Hydrodynamic Optimization


Chi Yang, Rainald Lohner
(George Mason University, U.S.A.)
al., 2003]).

ABSTRACT
This paper present a gradient-based hierarchical hydrodynamic optimization toolkit (H2O). The main components of this toolkit are: a) A simple CFD tool and
an advanced CFD tool; b) A finite-difference approximate cost-function formulation, a finite-difference costfunction formulation, and an adjoint formulation for the
gradient computations; c) A steepest descent technique
for the change of design variables. The present hierarchical hydrodynamic optimization toolkit consists of
three optimization tools that can be used at different
stages of the design. For purposes of illustration, these
three tools are used to determine optimized hull forms
that have the same displacement as the classical Wigley
hull.
INTRODUCTION
The relentless advance in numerical methods and computer power has made accurate flow simulations of realistic geometries a reality. Such simulations are increasingly reducing the amount of lengthy (and costly)
experiments in the aerospace, car, train and shipbuilding industries, substituting them for high fidelity CFD
runs. This way of utilizing CFD is nothing more than
an exchange of real for virtual experiment. However,
CFD and its underlying mathematics offers the possibility to step beyond the capabilities of any experiment.
While the experiment (or stand-alone CFD run) only
measures the performance of the product as is, numerical methods can also predict the effect of changes
in the shape of the product. This has led, over the last
decade, to a large body of literature on optimal shape
design (see, e.g. [Jameson, 1988 and 1995; Kuruvila
et al., 1995; Anderson and Venkatakrishnan, 1997; Elliott and Peraire, 1997; 1998; Korte et al., 1995; Mohammadi, 1997; 1999; Reuther, et al. 1997; Drela,
1998;, Medic et al., 1998; Nielsen and Anderson, 1998;
Dreyer and Matinelli, 2001; Gumbert et al.,2001; Li et
al., 2001; Mohammadi and Pironneau, 2001; Soto and
Lohner, 2001a,b; 2002; Soto, et al., 2002; Lohner, et

Although CFD-based hull-form optimization is not routinely used for ship design, applications of CFD tools
to hydrodynamic optimization mostly for reducing
calm-water drag and wave patterns have been reported in a significant number of studies. These studies
attest to a rapidly growing interest in hydrodynamic optimization (see, e.g. [Janson and Larsson, 1996; Tahara
and Himeno, 1998; Hino, 1999; Percival et al. 2001;
Peri et al., 2001; Peri and Campana, 2001; 2003; Yang
et al., 2000; 2002a;b]).
In order to compare the merit of different designs, a cost
function I is defined. This cost function depends on
design parameters , and the changes in flow variables
u(
) due to them. The aim is then to minimize (or
maximize) this cost function:
I(
, u(
)) min ,

(1.1)

subject to the following constraints:


- PDE Constraints:
R(u) = 0

(1.2)

- Geometric Constraints:
g(
) 0

(1.3)

- Physical Constraints:
h(u) 0

(1.4)

Examples for the cost function I are drag or prescribed


pressure, for PDE constraints R(u) the Euler/NavierStokes equations or Laplace equation, for geometric
constraints g(
) the displacement or transverse moment of inertia of the waterplane, and for physical constraints h(u) a minimal pressure to prevent cavitation.
The simplest (and most expensive) way to proceed is
by copying what nature has done in the course of evolution: try variations of , recalculate the flowfield and

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

keep the ones that minimize (i.e. improve) the cost


function I(
, u(
)). This class of optimization techniques are known as genetic algorithms. While simple to code and use, robust and suited for rough cost
functions, the number of the CFD runs required for N
design variables is at least of O(N 2 ). The speed of convergence is also strongly dependent on the crossover,
mutation and selection criteria.
The second class of optimization techniques is based
on evaluating gradients of I(
, u(
)). From a Taylor
series expansion we have
I + I I + I,
.

(1.5)

This implies that if one chooses:

= I,

(1.6)

for sufficiently small the new functional has to diminish:


I + I = I I,T I, I

(1.7)

There exist a variety of ways of computing the required


gradients I, . The easiest way is via finite differences. For each i , vary its value by a small amount,
recompute the cost function I, and measure the gradient
with respect to i . For central differences, this implies
O(2N ) field solutions for each gradient evaluation. An
alternative is to use a first order finite difference with
complex variables. This requires O(N ) field solutions
for each gradient evaluation, but at the cost of a flow
solver with complex variables. For noisy or rough
cost functions, gradients may be computed from socalled response surfaces. The parameter space in a region close to the present design is populated, and a loworder polynomial is fitted through these data points.
The gradients are then obtained from the low-order
polynomial. This type of technique also requires O(N )
field solutions for each gradient evaluation.
The only alternative to obtain gradients in a more expeditious manner is via adjoint solvers. The PDE constraints are used in adjoint solvers to obtain all gradients at once analytically. The variation of cost function
I exhibits only derivatives with respect to , i.e. no explicit derivatives with respect to u appear. The cost for
the evaluation of gradients is independent of the number of design variables. This type of technique requires
O(1) evaluation of the cost function to get all gradients.
So it is extremely fast, especially for a case with many
design variables. However, one needs to write an adjoint solver for adjoint equations in addition to the flow

solver, and the boundary conditions for adjoint equations are not general.
Hydrodynamic design of ships (like any design) involves several stages, from preliminary and early-stage
design to late-stage and final design. Gradient-based
techniques represent a very powerful class of optimization techniques that are well suited for the late stage design. This paper present a hierarchical hydrodynamic
optimization toolkit. The main components of this
toolkit are: a) A simple CFD tool and an advanced CFD
tool; b) A finite-difference approximate cost-function
formulation and a finite-difference cost-function formulation for the gradient computations, c) A steepest
descent technique for the change of design variables.
The present hierarchical hydrodynamic optimization
toolkit consists of three optimization tools that can be
used at different stages of the design. The first optimization tool (computer code H2O 1) is based on a
simple CFD tool. This CFD tool is used for evaluating
cost functions and cost-function gradients via finite differences. The second optimization tool (computer code
H2O 2) is based on an advanced CFD tool. Both cost
functions and cost-function gradients are evaluated using an advanced CFD tool via finite differences. The
third CFD tool (computer code H2O 3) is based on
both a simple CFD tool and an advanced CFD tool.
Cost functions are evaluated using an advanced CFD
tool and cost-function gradients are evaluated using a
simple CFD tool via finite differences.
For evey design cycle, the following steps are required:
- A flow solver;
- The evaluation of cost functions;
- The evaluation of cost-function gradients;
- A mesh movement module
Each of these steps will be discussed in the following
sections. For purposes of illustration, the present study
considers a simple hydrodynamic-optimization problem: minimization of the wave drag of a monohull ship
while keeping the displacement unchanged. The above
three tools are used to determine optimized hull forms
that have the same displacement as the classical Wigley
hull.
FLOW SOLVERS
A Simple CFD Tool
A simple CFD tool (computer code FKS) based on potential flow theory and Fourier-Kochin representation

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

of ship waves and slendar-ship approximation is integrated into the optimization toolkit. The flow computed by FKS is defined explicitly in terms of the ship
speed and the ship hull form. Specifically, the wave
drag associated with the wave energy transported by the
waves trailing the ship is determined from the Havelock formula in terms of the wave spectrum function.
The Fourier-Kochin representation of waves defines the
wave spectrum in terms of the velocity distribution at
the ship hull surface. As a result, this code is very fast,
can be used for any ship speed, and is ideally suited
for preliminary design, where information content is incomplet. It takes about 1 second CPU time on a Pentium 4 processor PC to compute the wave drag using
about 10,000 triangular elements on the hull. The application of FKS to optimize trimarans and monohulls
are reported by authors (Yang et al. Yang et al., 2000;
2002a;b) and the others (Percival et al., 2001). Code
FKS is used for evaluating cost functions and costfunction gradients via finite differences for a gradientbased optimization tool (computer code H2O 1).
An Advaced CFD Tool
The advanced CFD tool (computer code FEFLO) based
on Euler/RANS equations and nonlinear free surface conditions is also integrated into the optimization toolkit. The overall scheme of this free surface
solver combines a finite-element, projection-type threedimensional incompressible flow solver with a finite element, two-dimensional advection equation solver for
the free surface equation.
The numerical schemes chosen to solve the incompressible Euler equations are based on the following
criteria:
- Spatial discretization using unstructured grids (in
order to allow for arbitrary geometries and adaptive refinement);
- Spatial approximation of unknowns with simple
finite elements (in order to have a simple input/output and code structure);
- Temporal approximation using implicit integration of viscous terms and pressure (the interesting
scales are the ones associated with advection);
- Temporal approximation using explicit integration
of advective terms;
- Low-storage, iterative solvers for the resulting systems of equations (in order to solve large 3-D
problems); and

- Steady results that are independent from the


timestep chosen (in order to have confidence in
convergence studies).
A detailed description of the numerical solution procedure can be found in authors previous work (Lohner et
al., 1998; Yang and Lohner, 1998; Lohner et al., 1999;
Lohnerl. 2003).
The free surface equation is treated as a standard scalar
advection equation with source terms for the x, y plane.
The faces on the free surface are extracted from the 3D volume grid. Points and elements are renumbered
locally to obtain a 2-D triangular finite element mesh in
x, y.
One complete timestep consists of the following steps:
- Given the boundary conditions for the pressure,
update the solution in the 3-D fluid mesh;
- Extract the velocity vector v = (u, v, w) at the
free surface and transfer it to the 2-D free surface
module;
- Given the velocity field, update the free surface;
- Transfer back the new free surface to the 3-D fluid
mesh, and impose new boundary conditions for
the pressure.
In order to minimize the costs associated with geometry recalculations and grid repositioning along surfaces,
we do not move the grid in the present calculation, but
only change the pressure boundary condition after each
update of the free surface.
The computer code FEFLO has been used to compute
the steady ship flows around different hull forms that
are fixed or free to sink and trim (Yang and Lohner,
2002). The code FEFLO has also been used in several
studies for advanced carrier hulls, trimarans, tankers
and ferries. The three-dimensional nonlinear steady
free-surface flow simulation can be performed on a
Pentium 4 processor PC. The CPU time is about 5 minutes for computing the flow using Euler equations with
half a million tetrahedral elements. The computer code
FEFLO can also be used to solve RANS equations.
The effects of turbulence are modeled using either the
Baldwin-Lomax or the k-e model.
The code FEFLO will be used for evaluating cost functions and cost-function gradients via finite differences
for a gradient-based optimization tool (H2O 2). Code
FEFLO and code FKS will be used together for a new
optimization tool (H2O 3) to minimize wave drag,

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

where the cost functions are evaluated using code FEFLO, and the cost-function gradients are evaluated using code FKS via finite differences. The optimization tool H2O 3 requires much less computing time in
comparison with the optimization tool H2O 2.
MESH MOVEMENT MODEL
For purposes of illustration, three optimization tools are
applied to the Wigley hull. The waterline is represented
by a six-point spline. Four of six spline points are allowed to change. The hull length is fixed during optimization. These four spline points are chosen as design
variables. A steepest descent technique is used for the
changes of design variables, thus update the hull surface definition.
The discretization of a general three-dimensional computational domain into an unstructured assembly of
tetrahedra is accomplished by means of an advancing
front grid generation procedure. This procedure requires that the geometry of the computational domain
be defined in terms of an assembly of surface patches,
and that the spatial variation of element size and shape
be prescribed. The first step in the process is the triangulation of the computational boundary surfaces. The
assembly of resulting triangles forms the initial front
for the three-dimensional grid generation process. The
advancing front method is then used to fill the computational domain with tetrahedra, which are generated so
as to meet a user-prescribed distribution of element size
and shape.
Both analytical surface patches (planes, Coons patches
etc.) and discrete surface patches (defined by a surface
triangulation) are used to describe the present computational domain boundaries, which consist of the hull surface, free surface, inflow plane, exit plane and bottom
plane. Specifically, the preprocessor FECAD first reads
in the hull offset data. A triangulation is then generated
from the offset data. This triangulation is subsequently
used to define the hull surface in a discrete manner. The
rest of the boundary surfaces are defined analytically.
The desired element size and shape are prescribed via
background grids and sources. The computational domain is covered by a coarse background grid of tetrahedral elements. The desired element size and shape
are then specified at the nodes of this background grid.
During grid generation, the local element size and
shape are obtained using linear interpolation. In addition, both line and surface sources are used on the
hull surface and the free surface to further define the element size. The sources on the hull surface ensure the

generation of a finer mesh to accurately capture the hull


geometry and the complex flow in the bow and stern
regions. The sources on the free surface yield a finer
mesh in the Kelvin wave pattern region.
Once a new hull surface is updated, the above automatic
unstructured grid generator (code GEN3D) is used to
generate triangular surface grids for CFD tool FKS and
tetrahedral volume grids for CFD tool FEFLO.
NUMERICAL RESULTS
The classical Wigley hull is taken as initial hull in
the optimization process. Three optimization tools
H2O 1, H2O 2 and H2O 3 are applied to the Wigley
hull. The hull length and displacement are fixed while
the wave drag are minimized during the optimization
process. The Wigley hull is optimized for three Froude
numbers, F = 0.25, 0.289, 0.316, respectively. For a
given Froude number, all three optimization tools are
used to determine the optimal hull form.
Figs. 1a and 1b show the comparison of framelines and
wave patterns for the original Wigley hull and the optimized hull obtained using optimization tool H2O 1.
Figs.2a and 2b show the comparison of framelines and
wave patterns for the original Wigley hull and the optimized hull obtained using optimization tool H2O 2.
Figs. 3a and 3b show the comparison of framelines and
wave patterns for the original Wigley hull and the optimized hull obtained using optimization tool H2O 3.
Figs 4a, 4b and 4c show the comparison of waveprofiles of original hull and optimized hull obtained using
H2O 1, H2O 2, H2O 3, respectively. In summary,
Figs. 1a - 4c show the framelines, wave patterns and
waveprofiles obtained from the original hull, optimized
hulls via three optimization tools at F=0.25. Figs. 5a
- 8c show the corresponding results for F=0.289, and
Figs. 9a - 12c for F=0.316.
The present simple CFD tool and advanced CFD tool
have been validated in authors previous work. The
waveprofiles and wave drag obtained using present
CFD tools have shown good agreement with experimental data (see Yang et al., 2002a;b;c).
For all three Froude numbers, it can be seen that that the
optimization tools based on simple CFD tool (H2O 1)
and advanced CFD tool (H2O 2) have produced the
similar optimal hull forms. The computing time for
H2O 1 is less than 10 minutes, H2O 2 in the order
of 10 hours. The new approximate-gradient-based optimization tool H2O 3 generate almost exact same optimized hull form as gradient-based optimization tool

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

-0.01

-0.01

-0.02

-0.02

-0.03

-0.03

-0.04

-0.04

-0.05

-0.05

-0.06

-0.06

-0.07
-0.08

-0.06

-0.04

-0.02

0.02

0.04

0.06

0.08

-0.07
-0.08

-0.06

-0.04

-0.02

0.02

0.04

0.06

0.08

Fig. 1a. Comparison of framelines of original


hull and optimized hull via H2O 1 (F=0.25).
(solid line: original; dashed line: optimized)

Fig. 2a. Comparison of framelines of original


hull and optimized hull via H2O 2 (F=0.25).
(solid line: original; dashed line: optimized)

Fig. 1b. Comparison of wave patterns generated by


original hull and optimized hull via H2O 1 (F=0.25).
(left: original; right: optimized)

Fig. 2b. Comparison of wave patterns generated by


original hull and optimized hull via H2O 2 (F=0.25).
(left: original; right: optimized)

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

0.015

Wave Elevation (Fr=0.25)

0
-0.01
-0.02
-0.03
-0.04
-0.05
-0.06
-0.07
-0.08

original
optimized-H2O_1

0.01
0.005
0
-0.005
-0.01
-0.5 -0.4 -0.3 -0.2 -0.1

-0.06

-0.04

-0.02

0.02

0.04

0.06

0.08

Fig. 3a. Comparison of framelines of original


hull and optimized hull via H2O 3 (F=0.25).
(solid line: original; dashed line: optimized)

0.1

0.2

0.3

0.4

0.5

X-Coordinates

Fig. 4a. Comparison of waveprofiles of original


hull and optimized hull via H2O 1 (F=0.25).
(solid line: original; dashed line: optimized)

Wave Elevation (Fr=0.25)

0.015

original
optimized-H2O_2

0.01
0.005
0
-0.005
-0.01
-0.5 -0.4 -0.3 -0.2 -0.1 0 0.1
X-Coordinates

0.2

0.3

0.4

0.5

Fig. 4b. Comparison of waveprofiles of original


hull and optimized hull via H2O 2 (F=0.25).
(solid line: original; dashed line: optimized)

Wave Elevation (Fr=0.25)

0.015

original
optimized-H2O_3

0.01
0.005
0
-0.005
-0.01
-0.5 -0.4 -0.3 -0.2 -0.1 0 0.1
X-Coordinates

0.2

0.3

0.4

0.5

Fig. 4c. Comparison of waveprofiles of original


hull and optimized hull via H2O 3 (F=0.25).
(solid line: original; dashed line: optimized)

Fig. 3b. Comparison of wave patterns generated by


original hull and optimized hull via H2O 3 (F=0.25).
(left: original; right: optimized)

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

-0.01

-0.01

-0.02

-0.02

-0.03

-0.03

-0.04

-0.04

-0.05

-0.05

-0.06

-0.06

-0.07
-0.08

-0.06

-0.04

-0.02

0.02

0.04

0.06

0.08

-0.07
-0.08

-0.06

-0.04

-0.02

0.02

0.04

0.06

0.08

Fig. 5a. Comparison of framelines of original


hull and optimized hull via H2O 1 (F=0.289).
(solid line: original; dashed line: optimized)

Fig. 6a. Comparison of framelines of original


hull and optimized hull via H2O 2 (F=0.289).
(solid line: original; dashed line: optimized)

Fig. 5b. Comparison of wave patterns generated by


original hull and optimized hull via H2O 1 (F=0.289).
(left: original; right: optimized)

Fig. 6b. Comparison of wave patterns generated by


original hull and optimized hull via H2O 2 (F=0.289).
(left: original; right: optimized)

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

0.015

Wave Elevation (Fr=0.289)

0
-0.01
-0.02
-0.03
-0.04
-0.05
-0.06
-0.07
-0.08

original
optimized-H2O_1

0.01
0.005
0
-0.005
-0.01
-0.5 -0.4 -0.3 -0.2 -0.1

-0.06

-0.04

-0.02

0.02

0.04

0.06

0.08

Fig. 7a. Comparison of framelines of original


hull and optimized hull via H2O 3 (F=0.289).
(solid line: original; dashed line: optimized)

0.1

0.2

0.3

0.4

0.5

X-Coordinates

Fig. 8a. Comparison of waveprofiles of original


hull and optimized hull via H2O 1 (F=0.289).
(solid line: original; dashed line: optimized)

Wave Elevation (Fr=0.289)

0.015

original
optimized-H2O_2

0.01
0.005
0
-0.005
-0.01
-0.5 -0.4 -0.3 -0.2 -0.1 0 0.1
X-Coordinates

0.2

0.3

0.4

0.5

Fig. 8b. Comparison of waveprofiles of original


hull and optimized hull via H2O 2 (F=0.289).
(solid line: original; dashed line: optimized)

Wave Elevation (Fr=0.289)

0.015

original
optimized-H2O_3

0.01
0.005
0
-0.005
-0.01
-0.5 -0.4 -0.3 -0.2 -0.1 0 0.1
X-Coordinates

0.2

0.3

0.4

0.5

Fig. 8c. Comparison of waveprofiles of original


hull and optimized hull via H2O 3 (F=0.289).
(solid line: original; dashed line: optimized)

Fig. 7b. Comparison of wave patterns generated by


original hull and optimized hull via H2O 3 (F=0.289).
(left: original; right: optimized)

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

-0.01

-0.01

-0.02

-0.02

-0.03

-0.03

-0.04

-0.04

-0.05

-0.05

-0.06

-0.06

-0.07
-0.08

-0.06

-0.04

-0.02

0.02

0.04

0.06

0.08

-0.07
-0.08

-0.06

-0.04

-0.02

0.02

0.04

0.06

0.08

Fig. 9a. Comparison of framelines of original


hull and optimized hull via H2O 1 (F=0.316).
(solid line: original; dashed line: optimized)

Fig.10a. Comparison of framelines of original


hull and optimized hull via H2O 2 (F=0.316).
(solid line: original; dashed line: optimized)

Fig. 9b. Comparison of wave patterns generated by


original hull and optimized hull via H2O 1 (F=0.316).
(left: original; right: optimized)

Fig. 10b. Comparison of wave patterns generated by


original hull and optimized hull via H2O 2 (F=0.316).
(left: original; right: optimized)

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

0.015

Wave Elevation (Fr=0.316)

0
-0.01
-0.02
-0.03
-0.04
-0.05
-0.06
-0.07
-0.08

original
optimized-H2O_1

0.01
0.005
0
-0.005
-0.01
-0.5 -0.4 -0.3 -0.2 -0.1

-0.06

-0.04

-0.02

0.02

0.04

0.06

0.08

Fig. 11a. Comparison of framelines of original


hull and optimized hull via H2O 3 (F=0.316).
(solid line: original; dashed line: optimized)

0.1

0.2

0.3

0.4

0.5

X-Coordinates

Fig. 12a. Comparison of waveprofiles of original


hull and optimized hull via H2O 1 (F=0.316).
(solid line: original; dashed line: optimized)

Wave Elevation (Fr=0.316)

0.015

original
optimized-H2O_2

0.01
0.005
0
-0.005
-0.01
-0.5 -0.4 -0.3 -0.2 -0.1 0 0.1
X-Coordinates

0.2

0.3

0.4

0.5

Fig. 12b. Comparison of waveprofiles of original


hull and optimized hull via H2O 2 (F=0.316).
(solid line: original; dashed line: optimized)

Wave Elevation (Fr=0.316)

0.015

original
optimized-H2O_3

0.01
0.005
0
-0.005
-0.01
-0.5 -0.4 -0.3 -0.2 -0.1 0 0.1
X-Coordinates

0.2

0.3

0.4

0.5

Fig. 12c. Comparison of waveprofiles of original


hull and optimized hull via H2O 3 (F=0.316).
(solid line: original; dashed line: optimized)

Fig. 11b. Comparison of wave patterns generated by


original hull and optimized hull via H2O 3 (F=0.316).
(left: original; right: optimized)

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

H2O 2. However, the computing time of H2O 3 is


less than half of that of H2O 2. Numerical results have
also shown that the wetted surface and displacement remains almost unchanged, while more than 50% wave
drag reduction is achieved with optimized hulls for every Froude number. The optimized hull had a reduced
displacement in both bow and stern region and an increased displacement in the middle of the hull.
An adjoint solver and code FEFLO have been
integrated into an adjoint-based optimization tool
(H2O 4). Some preliminary results have been obtained
for KRISO container ship (KCS) using this tool. The
objective is to modify the shape of bulb bow in order to
reduce wave drag. Figs. 13a,b show both original and
final mesh in the bulb bow region and the comparison
of wave drag coefficient obtained for both hull forms.

Modified hull

CONCLUSIONS AND OUTLOOK


A gradient-based hierarchical hydrodynamic optimization toolkit (H2O) has been developed and tested. The
optimization has been performed for a single-spped and
multi-speed optimization for a few hull forms. The
present hierarchical hydrodynamic optimization toolkit
consists of three optimization tools that can be used at
different stages of the design. For purposes of illustration, these three tools are used to determine optimized
hull forms that have the same displacement as the classical Wigley hull.
We are currently investigating in more depth a number
of nontrivial issues, including selection of an appropriate objective function, choice of optimization scheme,
representation of hull surface and choice of related design variables and corresponding constraints, and decision to perform optimization for a single-point design
or for multiple-point design, e.g. for a single ship speed
or for a range of speeds.

ACKNOWLEDGEMENTS
This work was partially funded by NRL LCP&FD. Dr.
William Sandberg was the technical monitor.

REFERENCES

Original hull
Fig. 13a. Surface mesh of original
and optimized KCS hull

Drela, M., Pros and Cons of Airfoil Optimization,


Frontiers of CFD98 (D.A. Caughey and M.M. Hafez
eds. ) World Scientific, 1998.

0.5

10000xCw

Elliott, J. and Peraire, J., Aerodynamic Optimization


on Unstructured Meshes with Viscous Effects, AIAA97-1849, 1997.

Original hull
Modified hull

0.4

Anderson, W. and Venkatakrishnan, V., Aerodynamic


Design Optimization on Unstructured Meshes With
a Continuous Adjoint Formulation, AIAA-97-0643,
1997.

0.3

0.2

Elliott, J. and Peraire, J., Constrained, Multipoint


Shape Optimization for Complex 3-D Configurations,
Aeronautical J. -102, 365-376, 1998.

0.1

0
0

0.2

0.4
Time

0.6

Fig. 13b. Wave drag coefficient of original


and optimized KCS hull

0.8

Dreyer, J.J. and Matinelli, L., Hydrodynamic Shape


Optimization of Propulsor Configurations Using a Continuous Adjoint Approach, underlineAIAA-01-2580,
2001.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Gumbert, C.R. Hou, G.J.W. and Newman, P.A., Simultaneous Aerodynamic and Structural Design Optimization (SASDO) of a 3-D Wing, AIAA-01-2527,
2001.
Hino, T., Shape Optimization of Practical Ship Hull
Forms
Using
Navier-Stokes
Analysis, Proc. of 7th Intl. Conf. on Numerical Ship
Hydrodynamics, Nantes, France, 1999.
Jameson, A., Aerodynamic Design via Control Theory, J. of Scientific Computing 3, 233-260, 1988.
Jameson, A., Optimum Aerodynamic Design Using
Control Theory, CFD Review 1995, J. Wiley & Sons,
1995.
Janson, C. and Larsson, L, A Method for the Optimization of Ship Hulls from a Resistance Point of
View,
Proc. of the 21st Symp. on Naval
Hydrodynamics Trondheim, Norway, 1996.
Korte, J.J., Salas, A.O., Dunn, H.J., Alexandrov, N.N.,
Follett, W.W., Orient, G.E. and Hadid, A.H., Multidisciplinary Approach to Aerospike Nozzle Design,
AIAA95-0478 (1995).
Kuruvila, G., Taasan, S. and Salas, M.D., Airfoil
Design and Optimization by the One-Shot Method,
AIAA-95-0478, 1995.
Li, W., Huyse, L. and Padula, S., Robust Airfoil Optimization to Achieve Consistent Drag Reduction Over a Mach Range, NASA/CR-2001-211042,
underlineICASE Report No. 2001-22, 2001.
Lohner, R., Yang, C. and Onate, E., Viscous
Free Surface Hydrodynamics Using Unstructured
Grids, Proceedings of the 22nd Symposium on Naval
Hydrodynamics, Washington, D.C., 1998.
Lohner, R., Yang, C., Onate, E. and Idelssohn, S.,
An Unstructured Grid-Based, Parallel Free Surface
Solver, Applied Numerical Mathematics, 31, 1999,
pp. 271-293.
Lohner, R., FEFLO Users Manual, GMU-CSI/CFD01-01, 2003.
Lohner, R., Soto, O. and Yang, C., An Adjoint-Based
Design Methodology for CFD Optimization Problems,
AIAA-03-0299, 2003.
Medic, G., Mohammadi, B., Moreau, S. and Stanciu,

M., Optimal Airfoil and Blade Design in Compressible and Incompressible Flows, AIAA-98-2898, 1998.
Mohammadi, B., A New Optimal Shape Design Procedure for Inviscid and Viscous Turbulent Flows,
Int. J. Num. Meth. Fluids, 25, 183-203, 1997.
Mohammadi, B., Flow Control and Shape Optimization in Aeroelastic Configurations, AIAA-99-0182,
1999.
Mohammadi, B. and Pironneau, O., Applied Shape
Optimization for Fluids; Oxford Univ. Press, 2001.
Nielsen, E. and Anderson, W., Aerodynamic Design Optimization on Unstructured Meshes Using the
Navier-Stokes Equations, AIAA-98-4809, 1998.
Percival, S., Hendrix, D. and Noblesse, F., Hydrodynamic Optimization of Ship Hull Forms,
underlineApplied Ocean Research, 23, 337-355, 2001.
Peri, D., Rossetti, M. and Campana, F., Design
Optimization of Ship Hulls Via CFD Techniques,
J. of Ship Research, Vol. 45, No. 2, 2001.
Peri, D. and Campana, F., Multidisciplinary Design Optimization of a Naval Surface Combatant,
J. of Ship Research, Vol. 47, No. 1, 2003.
Peri,
D.
and
Campana, F., High Fidelity Models in Simulation Based
Design, Proc. of 8th Intl. Conf. on Numerical Ship
Hydrodynamics, Busan, Korea, 2003.
Reuther, J., Jameson, A., Alonso, J.J., Rimlinger, M.J.
and Saunders, D., Constrained Multipoints Aerodynamic Shape Optimization Using and Adjoint Formulation and Parallel Computers, J. of Aircraft, 36(1), 5174, 1999.
Soto, O. and Lohner, R., CFD Shape Optimization
Using an Incomplete-Gradient Adjoint Formulation,
Int. J. Num. Meth. Eng. 51, 735-753, 2001.
Soto, O. and Lohner, R., General Methodologies
for Incompressible Flow Design Problems, AIAA-011061, 2001.
Soto, O. and Lohner, R. A Mixed Adjoint Formulation
for Incompressible RANS Problems, AIAA-02-0451,
2002.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Soto, O., Lohner, R. and Yang, C., A Stabilized


Pseudo-Shell Approach for Surface Parametrization in
CFD Design Problems, Comm. Num. Meth. Eng. 18,
4, 251-258, 2002.
Tahara, Y. and Himeno, Y. An Application of Computational
Fluid
Dynamics to Tanker Hull Form Optimization Problem,
Proc. of the 3rd Osaka Colloquium on Advanced CFD
Applications to Ship Flow and Hull Form Design, Osaka, Japan, 1998.
Yang, C. and Lohner, R., Folly Nonlinear Ship Wave
Calculation Using Unstructured Grids and Parallel
Computing, Proceddings of 3rd Osaka Colloquium on
Advanced CFD Applications to Ship Flow and Hull
Form Design, Osaka, Japan, 1998.
Yang,
C.,
Noblesse,
F.,
Lohner, R. and Hendrix, D., Practical CFD Applications to Design of a Wave Cancellation Multihull Ship;
Proc. of the 23rd Symp. on Naval Hydrodynamics, Val
de Reuil, France, September, 2000.
Yang, C. and Lohner, R. Calculation of Ship Sinkage
and Trim Using a Finite Element Method and Unstructured Grids, Int. J. CFD 16, (3), 217-227, 2002.
Yang, C., Lohner, R. and Soto, O., Optimization of
a Wave Cancellation Multihull Ship using CFD Tools,
J. of Hydrodynamics B, 1, 1-7, 2002.
Yang, C., Soto, O., Lohner, R. and Noblesse,
F. Hydrodynamic Optimization of a Trimaran;
Ship Technology Research,49, 2, 70-92, 2002.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Solomon C. Yim
Oregon State University, USA
Thank you very much for an interesting
presentation. Pressure was one of the parameters for
optimization. However, in hull design, extreme
stress may not be directly associated with extreme
pressure location. Hull elasticity will need to be
included to determine extreme stress. How much
more computational effort will be needed to include
hydroelasticity into the optimization process to
determine extreme stress?
AUTHORS REPLY
Thank you very much for your interest in
our paper and for the good comment. A single
variable, wave drag coefficient, is chosen as the
objective function in present study. The purpose of
such an optimization is to minimize the wave drag,
which is part of the hydrodynamic design of hull
forms. We agree that the prediction and minimization
of the extreme stress is very important in the
structural design of hull forms, in which hull
elasticity will need to be included. Multiple objective
functions will be required. One shape optimization
example will be the minimization of wave drag and
extreme stress, as indicated by the discussor. In this
case both wave drag (pressure integration) and
extreme stress (load distribution) will be chosen as
objective functions. The rigid hull boundary
condition can still be used if the hull deformation is
relatively small. The additional computational effort
is mainly from the structural analysis under a given
hydrodynamic load distribution. The authors have
developed a loose-coupling algorithm for the fluidstructure interaction problem (Lohner, et al., 1995).
This algorithm can be used if the hull deformation
has to be taken into account in the hydrodynamic
analysis. The computational effort will al least be
doubled.
REFERENCES
Lohner, R., Yang, C., Cebral, J., Baum, J. D., Luo,
H., Pelessone, D. and Charman, C. "Fluid-Structure
Interaction Using a Loose Coupling Algorithm and
Adaptive Unstructured Grids", Computational Fluid
Dynamics Review 1995, (M. M. Hafez Editor), John
Wiley and Sons, 1995.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Emilio F. Campana
Istituto Nazionale per Studi ed Esperienze di
Architettura Navale, Italy
The
Authors
address
the
challenging problem of reducing the CPU time
required for numerical optimization when using
expensive high fidelity flow solvers. The approach
uses high- and low-fidelity solvers to compute the
objective function and its gradient, respectively.
Do the authors use any estimate or
control to check if the low-fidelity model is correctly
guiding the optimizer in the right descent direction?
Recently Alexandrov (Alexandrov and Lewis 1996,
Alexandrov et al. 2000, Alexandrov and Lewis 2003)
proposed a first-order approximation and model
management technique for monitoring the use of
low-fidelity solvers in variable-fidelity methods. Via
a trust-region method, their approach guaranties (to
the first order) global convergence of the overall
optimization process to a solution of the high-fidelity
problem.

test cases. However, it would be very interesting for


us to try the first-order approximation and model
management technique recommended by the
discussor with our low- and high-fidelity solvers and
to study global convergence of the overall
optimization process to a solution of the high-fidelity
problem.

REFERENCES
N.M. Alexandrov and R.M. Lewis. A trust region
framework for managing approximation models in
engineering approximations, Sept. 1996. AIAA
Papers 96-4101 and 96-4102
N.M. Alexandrov, R.M. Lewis, C.R. Gumbert, L.L.
Green, and P.A. Newman. Optimization with
variable-fidelity models applied to wing design, Jan.
2000. AIAA Paper 2000-0841. Accepted for
publication on the AIAA J. of Aircraft.
N.M. Alexandrov and R.M. Lewis. First-order
approximation and model management in
optimization,
in Large-Scale PDE-Constrained
Optimization, L.T. Biegler, O. Gattas, M.
Heinkenschloss, B. van Bloemen Waanders editors,
Springer, 2003.
AUTHORS REPLY
Thank you very much for your interest in
our paper and for providing information on
monitoring the use of low-fidelity solvers in variablefidelity methods. We would like to address your
comments as follows.
We used high- and low-fidelity solvers to
compute the objective function and its gradient. We
found that low-fidelity solver is correctly guiding the
optimizer in the right descent direction in the present

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics


St. Johns, Newfoundland and Labrador, CANADA, 8-13 August 2004

Quantitative Characterization of the Free-Surface Around


Surface Ships
James R. Rice, Don Walker, Thomas C. Fu, Anna Karion, Toby Ratcliffe
(Naval Surface Warfare Center, Carderock Division, USA)
ABSTRACT
Ship hull optimization is an extremely important field
in the Navy today. Goals of optimization include
decreasing wave resistance and minimizing wake
generation. The ability to accurately measure or
predict the flow field around a hull form in the near
and far field is a necessity in the optimization
process.
In order to make these types of
measurements accurately, it became necessary to
develop minimally intrusive and non-intrusive
measuring techniques and instrumentation. Various
instrumentation and techniques of near and far field
wake measurements have been developed and
improved at the David Taylor Model Basin of the
Naval Surface Warfare Center, Carderock Division
(NSWCCD) over the past few years. This paper will
provide a description and assessment of the various
instruments and techniques used to make these flow
field measurements at model scale. It will also
describe recent work extending these techniques to
full-scale, at-sea measurements and to a remote
application, where the instrumentation was located
~30.5 m from the free-surface.
INTRODUCTION
Three systems used for the measurement of near and
far field surface ship wakes will be described, and
compared in results from several experiments at both
model and full scale. These measurement systems
have been successfully used to obtain a complete
description of the free-surface around a model. The
systems used were Wave Cuts, Finger Probes, and
Quantitative Visualization (QViz).
The Wave Cut system consists of several
stationary capacitance probes attached to a boom
extending from the side of the model basin to
measure longitudinal wave heights. The Finger
Probe system is a system that was developed at
NSWC, which uses mechanical probes attached to a
traversing positioning system at the stern of the
model to measure both transverse and longitudinal

wave heights. Finally, QViz is a laser imaging


technique developed at NSWCCD (Furey & Fu,
2002) that uses a traversing light sheet and video
imaging to measure the free-surface along the side of
a model with particular detail in the vicinity of bow
wave breaking (Karion et al., 2003).
The
combination of all three systems allows for a
complete mapping of the near and far field freesurface around a model.
WAVE CUT MEASUREMENT SYSTEM
Theory of Operation
The sensing element of the capacitance probe is a 30
gauge (AWG) solid silver-plate copper wire with 0.1
mm kynar insulation, approximately 38.1 cm in
length. Attached to the sensing element is a weighted
1.22 m length of mylar fishing line, used to provide
probe stability in waves. The sensing element is
suspended with half its length submerged below the
surface of the basin water. The basin water is used as
the ground reference of the sensing elements circuit
card. With the copper wire completely insulated from
the water, the sensing element behaves as a capacitor,
with one plate being the copper wire, the second plate
the water, and the wire insulation acting as the
dielectric. As a wave in the water changes the
submergence height of the sensing element it also
changes the effective capacitor plate size, which
results in a change in capacitance. The capacitance
of the probe is governed by the following equation:

C=

2l
x10 6 F ,
b
ln
a

(1)

where = dielectric constant of the wire insulation


(F/m), a = diameter of wire conductor, b = diameter
of wire including insulation thickness, and l =
submerged wire length (m). In order to use this
property to create a wave height sensor, an electronic
circuit was designed that outputs a pulse train at a
fixed frequency with a duty cycle that is dependent

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

on the value of a given capacitor. The pulses are


integrated and filtered to produce a d.c. voltage. A
50% duty cycle, when averaged, produces a zero d.c.
voltage. As the duty cycle changes the d.c. voltage
will vary accordingly. By attaching the sensing
element, a varying capacitor, to the electronic circuit
the capacitance of the probe regulates the duty cycle
of the pulse train. As a result the final d.c. voltage
output is directly proportional to the capacitance of
the probe, and by extension, the wave height being
measured.

Calibration Procedures
In situ static calibrations are performed after the
completion of the test setup. To calibrate the probes,
the motorized uni-slide is traversed in 2.5 cm
increments for a total of +/- 7.6 cm. Data are
collected at each increment step for each of the
probes. A straight line fit is performed and a slope is
calculated and stored for each probe. An in situ
calibration allows us to calibrate the probes, the
signal conditioning amplifiers, and the A/D converter
as a system.

Experimental Setup
A truss section (wave boom), cantilevered from the
basin wall over the water, provides a structure on
which instrumentation is mounted. The wave boom
extends 6.83 m from the basin wall, which is
approximately 0.92 m short of the basin centerline.
Mounted vertically on the wave boom is a motorized
uni-slide with an attached horizontal bar. The
capacitance probes electronics are mounted on the
horizontal bar of the uni-slide. The uni-slide allows
precise placements of the probes vertical positions
(probe submergences) used during static calibration
(Figure 1). Four to six probes are generally used for
an experiment. Transverse probe positions are
referenced to centerline of the model, with probe #1
being the closest inboard and probe #4 the farthest
outboard.

Operating Procedures
Probe zeroes are collected in calm water before each
run. The model is then run through the test section,
past the probes, at a constant speed. As the model
approaches the test section, a strip of reflective tape
positioned on the carriage triggers a photo sensor at
the side of the basin, which starts data collection.
The position of the photosensor and the duration of
data collection are adjusted to insure that the
maximum amount of data is collected before tank
wall reflections occur. Data are filtered at 10 Hz.
with a 3 pole Bessel filter and collected at a sampling
rate of 100 Hz. for 20 to 30 seconds depending on
model speed and photo sensor position.
An example of these data is shown in Figure 2.
The results from the capacitance probe measurements
are used to obtain free wave spectra, from which
wave resistance coefficients are calculated. Previous
experiments using this technique are reported in
Ratcliffe, Mutnick & Rice, (2001). While the Wave
Cuts are useful for far field measurements, near field
measurements are more often obtained with Finger
Probe and QViz systems.

Capacitance Probe
Electronics

30knt Propelled - Probe 2


0.02

0.015

Ground probe

0.01

Capacitance Probe
WaveHeight (Ship Lengths)

Figure 1: Wave boom with capacitance probes and


uni-slide traverse, cantilevered from basin wall.
A photosensor is set to trigger data collection
when the forward perpendicular of the model is a
predefined distance from the capacitance probes. A
personal computer, using a 16-bit analog to digital
(A/D) converter, collects and stores the data.

0.005

Spot 39
NA
Spot 40

-0.005

-0.01

-0.015

-0.02
0

0.2

0.4

0.6

0.8

1.2

Distance (Ship Lengths)

Figure 2: Sample output from capacitance probes.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

STERN TOPOGRAPHY SYSTEM

Experimental Setup
To create a topology of the free-surface at the stern of
the model, four probes are mounted together with 5.1
cm spacing between probes. The set of probes is then
attached to an X-Y traverse that is mounted
horizontally to the carriage at the stern of the model.
The traversing area measures approximately 1.83 m x
2.74 m. Two string pots are attached to the traverse
and are used to track the longitudinal (X) and
transverse (Y) positions of the probes. A Pentium
class National Instruments PXI system personal
computer, using a 16-bit A/D converter and a motion
control interface are used to traverse the probes, and
to collect and store the data. The collection computer
is networked with a 350 MHz Pentium II class laptop
computer, which is used for data analysis and
plotting.

Theory of operation
The finger probe is a vertically oriented, mechanized
probe, which continuously searches for the free
surface. The sensing element of the probe is a .04 cm
diameter, 5.1 cm long stainless steel wire. The wire
is mounted into a copper tube, which makes up the
body of the probe. A geared rack, attached to the
probe body, allows the probe to be driven up and
down in the vertical plane by a servomotor. An
electrical conduction, through the probe, is sensed by
an electronic circuit, which drives the servomotor.
When the probe is not in contact with the water
surface, there is no electrical conduction through the
probe and the servomotor drives the probe down
towards the surface of the water. Once contact is
made between the probe and the surface of the water
(circuit ground), an electrical conduction is sensed
and the probe is driven up out of the water. This
process is continuously repeated, causing the probe to
oscillate at the free surface at approximately 12 Hz.
in calm water. The probe is also geared to a
potentiometer to track its position along the z-axis
(wave height). Probe position is only recorded by a
sample and hold circuit at the instant the probe makes
initial contact with the water surface. This manner of
sampling, probe position, alleviates position error
from meniscus effects, due to surface tension.

Operating Procedures
Probe #1 is aligned longitudinally (X) and
transversely (Y) with the aft perpendicular and
centerline of the model respectively. Longitudinal
and transverse string pots are zeroed at this location,
and all future measurements are referenced to this
position. In order to collect the data needed to
generate a complete topology of the stern area, the
area is divided into a number of transverse cuts. The
possible number of transverse cuts per run is
dependent on model speed. Once the number of
traverse cuts per run is determined, a command file is
generated which controls the positioning of the
probes during the run. Using four probes spaced 5.1
cm apart along the x-axis, one transverse cut collects
an area of 15.2 cm x 116.8 cm. Starting as close to
the stern of the model as possible (1.3 cm),
successive transverse cuts are made with an
advancement of 20.3 cm along the x-axis between
cuts. Refer to Figure 3. Prior to each data run, a
zero run is performed. A zero run consists of
performing an identical collection of transverse cuts,
as in the data run, with the model sitting still in calm
water. This allows the subtraction of bias errors due
to misalignment or sagging of the traverses X-Y
plane and the surface of the water, which should be
parallel. After the zero run is performed, the model
is brought up to a constant speed and a collection of
transverse cuts is started. This process is repeated at
successive transverse locations until the desired stern
area of the model has been completely mapped. Data
is filtered at 10 Hz. with a 3 pole Bessel filter and
collected at a sampling rate of 100 Hz.

Calibration Procedures
Static calibrations are performed on the finger probes
in the lab, prior to the experiment. Probes are
mounted together on a bracket, and attached to a unislide. The probes are positioned over a container of
water, and allowed to track the calm free surface as
the uni-slide is traversed in 2.5 cm increments for a
total of +/- 7.6 cm. Data are collected at each
increment step for each of the probes. A straight line
fit is performed and a slope is calculated and stored
for each probe.
Potentiometers are used on the X and Y traverses
to record longitudinal and transverse positions of
finger probe #1, with all other probes having relative
offsets from the first. These potentiometers are
calibrated in situ by incrementally stepping the
traverses a known distance through the range of
interest while collecting the voltage of the
potentiometers. A straight line fit is performed and a
slope is calculated and stored for each axis.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Two sets of probes have been constructed. The


first set 43.2 cm. in length, provide a measuring
range of +/- 11.4 cm, and a second set 63.5 cm, with
a measuring range of +/- 21.6 cm. In tests that
required the measurement of larger displacements, a
1.22 m Z-traverse has been mounted to the X-Y
traverse. With the finger probes mounted to the Ztraverse, a computer monitors the positions of the
probes and adjusts the vertical position of the Ztraverse accordingly. The position of the Z-traverse
is recorded and added to the vertical positions of the
individual finger probes.

Stern of Model

Stern of Model

Figure 5: Finger probes mounted on z-traverse.

Figure 3: Diagram of traverse movements needed to


obtain a stern wave topology.

QUANTITATIVE VISUALIZATION SYSTEM


Theory of operation
The Quantitative Visualization system (QViz) is a
laser imaging technique. In this technique, a laser
sheet is oriented perpendicularly to the side of the
model and to the water surface. A video camera
records the resulting reflection of the laser sheet off
of the turbulent water surface at various longitudinal
positions along side of the model. The captured
images are analyzed using image-processing software
to extract the water surface heights.

0.2

Filt. Z/L

0.15

0.1

0.05

-0.1

0.1

Y/L

0.2

0.3

Figure 4: Example output from laterally traversing a


single finger probe at a given longitudinal position.

Experimental Setup
The QViz system has multiple configurations that are
primarily driven by experimental requirements, i.e.
the area to be surveyed, and the location of the
experiment, and secondarily by physical limitations
on the ability to mount any particular configuration
securely. Therefore, the setup of the QViz system
must be slightly customized to the required

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

experiment. Examples of system configurations


include 1) one camera and one light sheet mounted to
a traverse, 2) two cameras separated laterally which
view the same light sheet, to cover a wider viewing
angle, and mounted to a traverse, 3) two cameras, one
facing forward and one facing aft, and two light
sheets mounted to one traverse, 4) two separate QViz
systems, each matching configuration (1), but with
the two traverse systems offset longitudinally to
cover a longer axial measurement distance. A
description of the basic setup for three types of
experiments will be discussed: a carriage-towed
model basin experiment; a full scale, at-sea
experiment, and a remote location experiment, where
the instrumentation is located approximately 30.5 m.
from the free surface.

A typical carriage test setup is shown in Figure 6


and is described as configuration (3). In this setup,
two laser light sheets are formed by splitting the
beam from a single laser source and focusing those
beams into two fiber optic probes. The probes are
assembled with cylindrical lenses so that the outputs
form laser light sheets. The light sheets are oriented
orthogonal to the centerline of the model and the
water surface and mounted at opposite ends of a
support structure (linear rail). Two cameras, each on
pan and tilt units, are mounted at the center of the
linear rail. One camera faces forward and is focused
on the forward laser light sheet while the second
camera faces aft and is focused on the aft laser light
sheet. The rail support is then mounted to a
horizontal linear traverse, which allows the complete
system to be traversed to various axial locations.
Each camera is attached to a dedicated computer with
an image acquisition card. The two computers are
linked through the digital I/Os of their image
acquisition boards. One computer is chosen as the
master and controls the traverses axial positioning
and the triggering of image acquisition for both
systems. In this configuration, the complete system
can be controlled from one computer.

The QViz system uses a diode-pumped, solid


state Nd:YAG laser, with an output of 2.5-3.0 W at
532 nm (Model MLM-0532 by Melles-Griot) to
create a laser light sheet with either a scanning mirror
or a cylindrical lens using a fiber optic probe. A CVA10 black & white, low light (0.1 lux), progressive
scan camera with a resolution of 640x480 pixels is
mounted a fixed distance from the light sheet.
Images are acquired on a Pentium II NT based
personal computer system using a National
Instruments PCI-1409 image acquisition card.
Motion control is accomplished using a Compumotor
stepper motor indexer and driver, interfaced to the
serial port of the image acquisition computer. When
multiple views are needed as shown in Figure 5, the
laser beam can be split to form multiple sheets.
Although a single PCI-1409 image acquisition board
can acquire video from four analog cameras, the
sustained aggregate frame rate is only 30 fps unless
the image size is reduced. In order to maximize
frame rate, each camera has a dedicated computer
system with an image acquisition card. To time
synchronize the computers and camera images, each
computer is equipped with an Adrienne Electronics
timecode generator/reader card, model PCI21V/RG1. The timecode cards are synchronized to
an external GPS timecode generator. The timecode
cards also have the ability to synchronize and
maintain the PCs system clock. This allows multiple
computer systems and image files time stamps to be
synchronize, and provides the ability to imprint the
timcode on the actual video before saving it to disk.
The imprint video timecode provides synchronizing
resolution of 1 to 2 frames.

Umodel
Model
Rhodamine 6G

laser sheet 1

laser sheet 2

Camera 1

laser
probe 2

Camera 2

laser probe 1

63.25 in

48.25 in

(measurements are from laser sheet to pivot of pan and tilt unit holding camera)

142.75 in

NOT TO SCALE

Figure 6: Diagram of a QViz carriage setup with


two laser light sheets and two cameras.
A full scale QViz experiment has been
accomplished on R/V Athena I. In the setup of this
experiment, three cameras, one laser light sheet, and
a linear traverse was used. The traverse, comprised
of linear rails, was mounted to the forward deck and
manually operated to adjust axial position of the light
sheet and cameras. A light sheet was formed by
mounting a scanning mirror in the middle of a

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

support rail from which the cameras were attached at


either end. Two cameras were mounted side-by-side,
facing aft, to accommodate the requested lateral
viewing distance of the surface profiles. The third
camera faced forward to capture any in-board surface
features that may have been blocked as a result of
steep wave slopes. Figure 7 shows an illustration and
image of the Athena test setup.

approximately 109.7 m in length by 73.2 m in width,


with wavemakers orthogonally positioned on two of
the four basin walls. The generation of single
dimensional waves in either the longitudinal or
transverse directions, as well as two-dimensional
waves is possible at this facility. The remote setup of
the QViz system was accomplished by mounting a
scanning mirror on a walkway that is suspended from
the roof of the facility 71.3 m above the water surface
(Figure 8). The camera was positioned on the side of
the basin approx. 6.1m above the surface of the water
and 41.94 m from the laser light sheet. In this setup a
single camera and light sheet were used, along with
one PC with a timecode card.

QViz
Camera

Camera 1
Camera 3

50
.6
Camera 2

Figure 7a: Illustration of Athena test setup

Laser Sheets

Flat Plate

Catwalks

Beach

73.2 m

Spatial
Reference

Camera 3

Beach
109.7 m

Spatial
Reference

Catwalk

QViz Set-Up
Laser Sheet

Camera 2

Laser Sheet
QViz Camera

21.34 m

41.94 m
Camera 1

Figure 7b: Image of Athena test setup during a run


Three computer systems where used, each with
its own image acquisition and timecode reader cards.
One computer was designated as the master computer
and controlled the triggering of image acquisition.
The linear traverse system was manually positioned
to the appropriate axial positions. Detailed result
from this experiment can be found in Fu, et al (2004).

Figure 8: Diagram and picture of Qviz remote


location setup.

The QViz system has also been used to make


wave height and surface roughness measurements
from a test area that was approximately 30.5 m from
the system. These measurements where taken at
NSWCCDs maneuvering basin facility. The basin is

Calibration Procedures
Zero reference images are collected during calm
water conditions. Similar to the Finger Probe system,

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

a zero run is collected with the Qviz system. A series


of calm water images are collected for each axial
location used in the experiment. The detected water
surface is used as a zero reference during analysis.

following data into the program: 1) path for a root


directory i.e. E:\QVizData, 2) project name i.e.
SurfaceShip, 3) test name i.e. DeepDraft, 4)
calibration path i.e. E:\Calibration\SurfaceShipCal
5) next run
number i.e. 01.
Once this
information is entered and saved, a directory
structure
is
created
as
follows:
E:\QVizData\SurfaceShip\DeepDraft\Run001. If
more than one system is being operated, the steps are
repeated for each system. The default operating
mode for the Qviz computer is as a master. If the
computer is to be run in slave mode, i.e. externally
triggered to start collection, that option is selected on
the front panel of the program. The next step is to
enter the operating range of all systems, starting
position, ending position, and increment distance.
Finally, enter the number of images to be collected at
each axial location. With the program(s) initialized,
pressing the software run button will move the
traverse into the starting position and advance the
program to the run screen. For slaved computers, this
step readies the computer for an external trigger to
start collection. For a master computer, this screen
allows the user to trigger the start of data collection.
When the start button on the master computer is
pressed, data collections are started on all computers.
Images are collected and stored at 30 frames per
second (fps). Image file names are created by
concatenating the test name and an index (frame)
number i.e. SurfaceShip001.dat. These files are
then saved in the directory ..\Run001. At the end
of the run, the run number is advanced and a new run
directory is created in preparation for the next run.

Camera calibrations are taken in situ when


possible by holding a calibration target in the plane of
the laser light sheet and recording an image. The
calibration target has a series of dots oriented a
known distance vertically and horizontally. Using a
pre-packaged calibration routine in National
Instruments Labview image analysis software (IMAQ
Vision), the calibration target image is analyzed and
the calibration corrections are stored with the image
to a file. These calibration corrections are then used
to correct test images during analysis, post
experiment. Refer to Figure 9. In addition to
performing the calibration in place when possible, the
pan and tilt positions and the distance from the image
plane are recorded for each camera. Working with
Applied Microvideo Incorporated, NSWCCD was
able to develop a pan and tilt unit with potentiometer
feedback and pneumatic brakes on both axes. These
pan and tilt units allow the user to remotely position
and lock the cameras into a known orientation. Post
experiment, the cameras can be taken back to the lab
and repositioned to their experiment orientation. A
calibration target can then be placed the recorded
distance from the camera, an image taken and then
used to develop the calibration corrections.

Rhodamine 6G is used during some basin model


tests to increase the water surface contrast in the laser
light sheet. Rhodamine 6G has the property of
fluorescing at a wavelength of approximately 570 nm
when exposed to light with a wavelength of
approximately 532 nm (laser source). Narrow band
filters are attached to the cameras that are tuned to
pass the wavelength of the fluorescing Rhodamine.
Adding the filters to the cameras lessens the
incidence of capturing unwanted stray light. The data
is then post processed and assembled into a topology
map.

Figure 9: The image on the left is an image of a


calibration grid for a given camera orientation. The
image on the right is the same image after applying
calibration corrections.
Operating Procedures
Operating procedures are fairly common between
multiple configurations and setups. The software
program used for data collection, position control and
analysis was developed at NSWCCD using National
Instruments Labview with motion control and image
analysis toolboxes. To operate the system, the
software must first be initialized for the given test.
Program Initialization consists of entering the

RESULTS
Results will be presented for each of the wave height
measurement systems, Wave Cut, Finger Probe, and
QViz. A discussion on, how the data is being used,

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

sources of error, and possible system improvements,


will be given.

New hull forms typically have been optimized to


minimize resistance. This has led to smaller
amplitude wave features around the hull. The reduced
amplitude of these features has made it more difficult
to make accurate measurements, especially at low
speeds. Areas of improvement for the Capacitance
Probe system are: 1) Identifying a suitable wire, with
a coating that would minimize the meniscus effect. 2)
Evaluating better wave dampening systems. 3)
Dynamically evaluating the system for probe
contamination.

Wave Cuts
The results from the capacitance probe measurements
are used to obtain free wave spectra, from which
wave resistance coefficients are calculated and
compared to predictions (Figure 10). Previous
experiments using this technique are reported in
Ratcliffe, Mutnick & Rice, (2001)

Resistance lbs. (ModelScale)

Predicted Model Resistance

Finger Probe

35.000
30.000

Results from the Finger Probe measurements are used


to create a stern topology of a model ship wake for
comparison with computational fluid dynamic (CFD)
predictions (Figure 11 shows an example). Breaking
waves generated from submerged transoms make the
stern topology especially useful as a data set.

Rt (Pressure)

25.000

Rt (Wavecut)

20.000

Measured Bare

15.000

Measured Trip

10.000

Measured Stud

5.000
0.000
0.000

2.000

4.000

6.000

8.000

Knots (Model-Scale)

Required improvements needed for the Finger Probes


include faster sampling speed, more simultaneous
measurements (more probes), and greater system
robustness.

Figure 10: Plot comparison of measured resistance,


wave cut calculated resistance, and predicted
resistance (Courtesy of Don Wyatt, SAIC).
Sources of error for this system are meniscus
effects on probes, probe contamination, and basin
noise. The meniscus effect is a result of capillarity,
which is occurs because intermolecular forces cause
water to stick to itself (surface tension) and to many
other materials. The meniscus is the clinging section
of water that is drawn away from the surface of the
water as the surface moves up or down the
capacitance probe. Probe contamination results from
oils and dust on the water surface becoming
deposited on the surface of the probes. These deposits
can change the capacitive properties of the probe in
addition to changing the properties of the meniscus
effects that already exists. Model basin noise is a
result of the residual wave energy residing in the tank
after each model run. To minimize these errors the
following steps are taken. First, the water surface is
skimmed and filtered to reduce surface contaminates.
Second, small instrumentation fans are mounted near
the capacitance probes, blowing on the water surface
between data runs to keep contaminants away from
the probes. Third, probes are wiped clean at the
beginning of each test day, or whenever the operator
has an indication of contamination, and a static
calibration is performed. Lastly, wait time between
runs is extended as long as feasibly possible.

Figure 11: An example of a completed stern


topology survey using finger probes.
QViz
The results from QViz are used to create a topology
of the free-surface along the side of the model, and
full scale boat, for comparisons with computational
fluid dynamic (CFD) predictions. The data are also
being used to examine the characteristics of breaking

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

x 2.54 cm

waves, i.e. surface fluctuations and roughness


(Karion et al, (2004)). Additional information on
QViz analysis can be found in Karion, et al (2003).
Figure 12 shows a sample QViz image with the water
line as detected by the image processing algorithms
superimposed in red and blue. Figures 13 and 14
show examples of free-surface contours obtained
using the QViz system and compared with
computational predictions.

Figure 13: An example of the measured free surface


compared to the computed free surface (courtesy of
D. Wyatt)
SUMMARY
The combination of these three systems, Wave Cut,
Finger Probe, and QViz, has become an invaluable
tool for researchers. The results are used as a basis
for evaluation of ship performance, empirical
modeling, and validation and verification of
computational fluid dynamic (CFD) predictions.
Additionally, in the future these instruments will
provide the necessary data for next generation flow
prediction schemes based upon CFD and data
assimilation.

x 2.54 cm

Figure 12: Analyzed results showing the detected


surface in NSWCCDs SeaKeeping and Maneuvering
Basin facility, during remote operation of Qviz.
Error for the QViz system has been estimated to
be +/- 2 pixels. Although this error sounds small, it is
relative to the area being imaged. Typical pixel sizes
range from 0.001 m (model-scale) to 0.003 m. (fullscale). Model scale resolution is set by the transverse
spatial resolution (i.e. the size of the area to be
measured) while at full-scale the resolution is limited
by the amplitude of the wave being measured. One
source of error in in-situ calibrations is how
accurately the calibration target can be positioned.
Another source of error, for post calibration, is the
accuracy at which the orientation of the camera can
be measured. A final source of error comes from the
ability to make an accurate reference measurement
relative to the model or boat.
An interest in breaking wave characterization is
driving the need to update the QViz system for higher
frame rates and resolutions. Frame rates of 100 fps
and higher are desirable.

Figure 14: Comparison plot of CFD Predictions


(Top), QViz (bottom), and Finger Probe (bottom aft).

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

ACKNOWLEDGEMENTS
The authors would like to acknowledge Mr. Steve
McGuigan, lead engineer for the development of the
Finger Probe. They would also like to acknowledge
the support and encouragement of Don Wyatt, SAIC
Dr. Arthur Reed, NSWC, and Dr. Pat Purtell. This
work has been developed over many years. Funding
and support has been provided by ONR under various
contracts and programs. Dr. L. Patrick Purtell is the
current program manager.
REFERENCES
Fu, T.C., Karion A., Rice J.R., Walker, D.C.,
Experimental Study of the Bow Wave of the R/V
Symposium
on
Naval
Athena
I,
25th
Hydrodynamics, St. Johns, Newfoundland and
Labrador, Canada, August 8-13, 2004.
Furey, D.A. and Fu, T.C., Quantitative Visualization
(QViz) Hydrodynamic Measurement Technique of
Multiphase Unsteady Surfaces, 24th Symposium on
Naval Hydrodynamics, Fukuoka, Japan, July 8-13,
2002.
Karion, A., Sur, T., Fu, T.C., Furey, D., Rice, J., and
Walker, D., "Experimental Study of the Bow Wave
of a Large Towed Wedge, 8th International
Conference on Numerical Ship Hydrodynamics,
Busan, Korea. September 22-25, 2003.
Ratcliffe, T. J., Mutnick I., Rice J., Stern Wave
Topography and Longitudinal Wave Cuts Obtained
on Model 5415, With and Without Propulsion,
Naval Surface Warfare Center, Carderock Division,
Hydrodynamics
Directorate
R&D
Report,
NSWCCD-50-TR-2000/028, January 2001

10

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Dane Hendrix
Naval Surface Warfare Center, Carderock Division,
USA
The free surface is what makes
hydrodynamics unique from the rest of fluid
dynamics. The authors have provided a summary of
the current state of the art in measuring free-surface
characteristics around large models. It is probably
beyond the reasonable scope of the current paper, but
it would be interesting to reflect on some free-surface
measurement instruments that have not been used
recently: stereo photography, shadowgraphs, and
perhaps others.
More directly related to the current paper, I
would be interested in what you see as the principal
advantages and disadvantages of the three systems
you described? You addressed their use in terms of
location relative to the model, but it is not obvious to
me that they are all restricted to these uses. You
have, for example, used the finger probe to measure
the water surface alongside a model or even to
measure a wave cut.
Why do you still use
capacitance probes for wave cuts? Perhaps you could
address the difficulty of setting each system up, the
reliability of the measurements, and the difficulty of
the required analysis.
Finally, do you have some opinion on the
future of free-surface measurements that you would
like to share?
AUTHORS REPLY
My experience with other free-surface
measurement techniques like shadowgraphs and
stereo photography are somewhat limited. In my use
of shadowgraphs in the towing basin we had a
stationary light source underwater and a screen
attached to the towing carriage at the stern of the
model just above the water surface. A video camera
was mounted above the screen to image the shadows
produced by surface features from the trailing wake
of the model, as we towed it passed the light source.
In these experiments we were more interested in the
qualitative measurement of surface features rather
then the quantitative measurements of wave heights.
The use of stereo photography in the
measurements of wave heights has been used at The
David Taylor Model Basin as recently as 1990. The
difficulties of these experiments in the past have been
the cost and time of analysis, finding an appropriate
material to image on the water surface, and the ability
to make measurements close to the model. The cost
and time spent on the analysis started with the

development of the film which was completed inhouse. The film was then sent to the National Ocean
Survey (NOS) for analysis. Upon return of the data
to DTMB the data was further analyzed and then
plotted. The materials used to image the water
surface were chads from computer punch cards,
which would be sprinkled on the surface of the water
prior to a data run. The problems with the chads
were they did not stay positively buoyant
indefinitely, making it hard to identify if they were
actually at the surface or just below when imaged. It
was also difficult to make measurements close to the
model at higher speeds due to the bow wave diverting
the chads away from the model. With todays
technology of digital camera and gigahertz personal
computers it is no question that stereo photography
can be accomplished cheaper and completely inhouse, in near real-time. If a suitable material can be
identified to image the surface, this technology
should probably be revisited.
The
three
instrumentation
systems,
capacitance probes, finger probes, and Qviz systems
where all developed over the years for specific needs,
which generally included overcoming the short
comings of its predecessor.
The capacitance probe was developed to
collect Kelvin wake wave height measurements at a
fixed transverse location in order to obtain free wave
spectra to calculate wave resistance coefficients.
Because the capacitance probes have weighted ends
that hang in the water to stabilize the probes, they are
better suited for stationary measurements as opposed
to being dragged through the water. An advantage of
the capacitance probes are the ability to get
continuous wave height readings because the probes
are always in contact with the water.
The
disadvantages of the capacitance probes are: 1) the
probes must remain stationary or errors in position
and wave height will be introduced into the
collection, 2) due to the fact that the probes are in
contact with the water, errors from probe/water
interaction, meniscus effects and contamination, can
exists. Unlike the capacitance probe the finger probe
was designed to be run with the model.
The finger probe was designed to make
localized standing wave measurements that the
capacitance probes could not make. The advantages
of the finger probe are: 1) the probe makes minimal
contact with the water which allows it to survive and
make high speed, standing wave measurements, and
2) the probe is resistant to contaminants and
meniscus effects. A disadvantage of the finger probe
is its frequency response. Due to the slow frequency
response of the finger probe, it is not an ideal
instrument for capturing wave heights of a traveling
wave. This is why the finger probe is not suitable to

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

replace the capacitance probe for wave cut


measurements, until we are able to significantly
increase its response time. And finally unlike the
finger probe which needs to be traversed to obtain a
wave profile the Qviz system was designed to capture
the profile instantaneously.
The Qviz system was designed to collect an
instantaneous wave profile measurement at a fixed
location, alleviating the need to traverse the system to
obtain a profile. The advantages of the Qviz system
are: 1) it is a non-invasive measurement system, 2)
the ability to instantaneously capture a wave profile,
and 3) the ability to collect both model and full scale
data. The disadvantages of the Qviz system are: 1) in
some cases, the need to use fluorescent dyes to obtain
data, and 2) the difficulty of obtaining highly
accurate image calibrations and setup measurements
and maintaining them throughout the experiment.
As it has been shown, each new instrument
development generally overcomes the shortcomings
of its predecessor. I fully expect that the next
generation of finger probes will replace the
capacitance probes and I predict within the next two
generations of Qviz that some variation of the Qviz
system, which could include LIDAR (Light Detection
and Ranging) will make most or all of our surface
height measurements.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics


St. Johns, Newfoundland and Labrador, CANADA, 8-13 August 2004

Unsteady RANS Simulation of


a Manoeuvring Ship Hull
A. Di Mascio, R. Broglia, R. Muscari
(Istituto Nazionale per Studi ed Esperienze di Architettura Navale, Rome, Italy)
R. Dattola (Italian Navy General Staff, IV Dept. SPMM, Rome, Italy)
ABSTRACT

whole flow field.

The computation of the flow around a surface piercing ship hull in forced planar motion is presented.
The flow is simulated by the numerical solution of
the unsteady RANS equations. The algorithm used
for the simulation is a fully implicit finite volume
scheme, which is unconditionally stable for any time
step. Free surface effects are taken into account by
means of a surface fitting approach. The numerical calculation of the flow around a ship hull with
prescribed yaw and sway motion is reported; the
simulation mimics the towing tank test in which the
planar motion mechanism (PMM) is used to enforce
the ship motion. The unsteady wave pattern and velocity field are shown; the dependence of the global
coefficients on the simulation parameters is also discussed.

The main objective of the present work is the


analysis of the unsteady flow field around a ship
hull with forced motion. Numerical solutions of
the unsteady incompressible NavierStokes equations are obtained by means of a second order multi
block finite volume solver. The algorithm is based
on the steady RANS solver developed at the INSEAN (Di Mascio et al., 2001): convective fluxes in
the momentum equations are evaluated by means of
a second order TVD scheme, whereas diffusive fluxes
are discretized by a centered scheme. The one
equation model by Spalart and Allmaras (1994) is
used to compute turbulent viscosity. At the free surface both dynamic and kinematic boundary conditions on the actual position are enforced; the movement of the surface is handled by a surface fitting
technique. Physical time-derivative in the governing
equations is approximated by a second order accurate, threepoint backward finite difference formula.
In order to obtain a divergence free velocity field at
every physicaltime step, a dual or pseudo time
derivative is introduced in the discrete system of
equations (Merkle and Athavale, 1987). Acceleration of the convergence ratio for the inner iteration
is achieved by means of local time stepping, an implicit Euler scheme with approximate factorization
(Beam and Warming, 1978) and an efficient multigrid technique (Favini et al., 1996).

INTRODUCTION
The numerical simulation of the steady and unsteady viscous turbulent flow field around surface
ship hulls is one of the major field of research in the
framework of marine hydrodynamics. For example,
the correct prediction of the hydrodynamic forces
and moments acting on a manoeuvring ship hull is
of great interest for safety and comfort reasons, and
it can be of fundamental importance for the estimation of the loads on the structures. Analysis of such
kind of problems by means of numerical investigation requires ad hoc methodologies capable to deal
with unsteadiness of the flow field and with the presence of moving bodies. Moreover, in the case of large
horizontal plane motions and/or roll motions, effects
due to viscosity and turbulence, vorticity production in the boundary layer, as well as flow unsteadiness like, for example, the vortex shedding, are only
some of the hydrodynamics phenomena that must
be taken into account. In addition, the presence of
the free surface can have a strong influence on the

The paper is organized as follow: first the governing equations for the unsteady motion of an incompressible viscous fluid in an inertial frame of
reference written in the arbitrary Lagrangian and
Eulerian (ALE) form (Anderson, 1995) are briefly
recalled; a detailed presentation of the proposed numerical algorithm follows. Then, numerical results
of the flow around a surface piercing hull with combined sway and yaw motion are reported; conclusion
and perspectives wind up the paper.
1

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

MATHEMATICAL MODEL

surface tension are neglected, the dynamic boundary


conditions read:

The governing equations for the unsteady motion of


an incompressible viscous fluid are briefly recalled
in this section. The equations are written in an inertial frame of reference. As the grid is moved and
deformed to fit the moving ship hull, the governing equations are written with respect to a moving
control volume. The equations in nondimensional
integral form (with respect to a reference velocity
U and a reference length L) are
I
U n dS = 0
S(V)

U dV +

p = ij ni nj +

ij ni t2j = 0 (3)

D F (x, y, z, t)
=0
Dt

(4)

Initial conditions have to be specified for the velocity field and for the free surface configuration:

(1)
(Fc Fd ) n dS = 0

S(V)

ui (x, y, z, 0) = u
i (x, y, z)
F (x, y, z, 0) = F (x, y, z)

(5)

NUMERICAL MODEL

Fc = pI + (U V) U


ij ni t1j = 0;

where ij is the stress tensor, whereas n, t1 and t2


are the surface normal and two tangential unit vectors, respectively. The actual position of the free
surface F (x, y, z, t) = 0 is computed by enforcing
the kinematic condition

V being a control volume, S(V) its boundary, and n


the outward unit normal. In equation (1), Fc and
Fd represent convective (inviscid and pressure) and
diffusive fluxes, respectively:

Fd =

z
;
F n2

Spatial discretization




1
T
+ t gradU + (gradU)
Rn

(2)

For the numerical solution of the equations (1), the


fluid domain D is partitioned into Nl structured
blocks Dl , each subdivided into Ni Nj Nk disjoint
l
l
hexahedrons Dijk
such that Dijk
= Dl . Conservation laws are then applied to each finite volume:

where V is the local velocity of the control volume


boundary, Rn = U L/ the Reynolds number,
the kinematic viscosity, whereas t denotes the non
dimensional turbulent viscosity. In the previous
equation and in what follows, ui is the ith Cartesian component of the velocity vector (the Cartesian
components of the velocity will be also denoted with
u, v, and w); p is a variable related to the pressure P
and the acceleration of gravity g (parallel to the vertical axis z,downward oriented) by p = P + z/F n2 ,
F n = U / gL being the Froude number.
In the present work, the turbulent viscosity has
been calculated by means of the one-equation model
by Spalart and Allmaras (1994).
The problem is closed by enforcing appropriate
conditions at physical and computational boundaries. On solid walls, relative velocity is set to zero
(whereas no condition on the pressure is required);
at the (fictitious) inflow boundary, velocity is set
to the undisturbed flow value, and the pressure is
extrapolated from inside; on the contrary, the pressure is set to zero at the outflow, whereas velocity
is extrapolated from inner points.
At the free surface, whose location is one of the
unknowns of the problem, the dynamic boundary
condition requires continuity of stresses across the
surface; if the presence of the air and the effects of

6 Z
X
s=1

U n dS = 0

Ss

U dV +

Vijk

6 Z
X
s=1

(Fc Fd ) n dS = 0

Ss

(6)
where Ss is the s-th face of the finite volume Dijk ,
whose measure is Vijk .
In order to obtain second order accuracy in
space, convective and viscous fluxes in the momentum equations, as well as surface integral of the velocity in the continuity equation, have to be computed by means of trapezoidal rule:
Z
U ndS = ul nl |0 As + O(2 )
Ss

Ss

Ss

Fc ndS = Fcs |0 As + O(2 )

(7)


Fd ndS = Fds 0 As + O(2 )

where the subscript 0 means that the quantities are


computed at the face center, As is the measure of
2

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Ss , is the diameter of Ss and:

u1 (ul vl )nl + pn1


Fcs = u2 (ul vl )nl + pn2
u3 (ul vl )nl + pn3

1l nl
Fds = 2l nl
3l nl

resulting scheme is only first order accurate. Higher


order accurate (up to any finite order) oscillation
free solutions can be obtained by modifying the evaluation of the right and left states of the Riemann
problem as explained in Harten et al. (1987). In the
particular case of second order accuracy, it can be
shown that left and right states have to be evaluated
as:
1
ql = qi,j,k
+ minmod(i1/2 , i+1/2 )
2

(8)

where vl , l = 1, 2, 3 are the components of the velocity of the control volume. The stress tensor at the
cell interface is computed as:


um
ul
lm |i+ 1 ,j,k = i+ 12 ,j,k
+
(9)
2
xl
xm i+ 1 ,j,k

qr = qi+1,j,k

(12)
where i+1/2 = qi+1,j,k qi,j,k and minmod is a
function that is applied to each vector component:

0
if xy 0
minmod(x, y) =
sign(x) min(|x|, |y|) if xy > 0
(13)
It is easy to prove that:

The derivatives of the velocity vector are obtained


by means of a finite volume approximation:

Z
1
um
=
um nl dS + O(2 ) (10)
xl i+ 1 ,j,k
Vi+ 12 ,j,k i+ 1 ,j,k
2

ql

where the integral is extended to the volume Vi+ 12 ,j,k


(whose boundary is i+ 12 ,j,k ) that includes the cell
face Si+ 21 ,j,k and is overlapped to half the cell (i, j, k)
and half the cell (i+1, j, k). In equation (9), i+ 12 ,j,k
denotes the sum of the kinematic and turbulent viscosity at the cell face.
The computation of the convective fluxes Fcs and
the surface integral of the velocity in the continuity
equation requires the evaluation of pressure and velocity at the face center. To this aim, a second order
ENO-type scheme has been adopted (Harten et al.,
1987). These schemes were originally developed for
compressible fluid flows, on the basis of the hyperbolic nature of the inviscid part of the Navier-Stokes
equations. The extension to incompressible flows is
possible when working in pseudo-compressible formulation.
The building block of this kind of algorithms is
the Godunovs scheme (1959), in which the flux vector at cell interface is computed as the solution of
a Riemann problem, whose right and left states are
given by the values of the numerical solution at two
neighboring cell centers. For example, at the cell
face i + 12 , j, k:
Fci+ 1 ,j,k = Fc (ql , qr ) = Fc (qi,j,k , qi+1,j,k )

1
minmod(i+1/2 , i+3/2 )
2

1
= qi,j,k + minmod(i1/2 , i+1/2 )
2

1 q
= qi,j,k +
x + O(x2 )
2 x i,j,k

(14)

= qi+1/2,j,k + O(x2 )
and:
qr

1
= qi+1,j,k minmod(i+1/2 , i+3/2 )
2

1 q
x + O(x2 )
= qi+1,j,k
2 x i+1,j,k
= qi+1/2,j,k + O(x2 )

(15)
and therefore, the Riemann flux being a Lipschitz
continuous function of its arguments:
Fc (ql , qr ) = Fc (qi+1/2,j,k ) + O(x2 )

(16)

The evaluation of the convective flux vector requires the solution of a Riemann problem at each
cell interface. In order to simplify the algorithm, a
second order accurate solution was used in place of
the exact one, which must be computed iteratively,
given the nonlinearity of the problem; details of the
algorithm can be found in Di Mascio et al. (2001).

(11)

Temporal integration

q being the vector of the state variables, which,


for pseudocompressible flows, take the form: q =
(p, u1 , u2 , u3 )T . This scheme yields oscillation-free
discrete solutions, also when the exact solutions are
discontinuous. However it can be shown that the

The semi-discrete system of equations can be rewritten in vector form as:



Vq

+ Rijk = 0
(17)
t
ijk

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

where = diag (0, 1, 1, 1) and:


Z
1
q=
(p/, u, v, w)T dV
Vijk Vijk

Turbulence model
For the sake of completeness the model by Spalart
and Allmaras (1994) is recalled here. The model is
based on the eddy viscosity concept, i.e. on the assumption that the Reynolds stress tensor (u0i u0j )
is related to the mean strain rate through an apparent turbulent viscosity (eddy viscosity) t :

(18)

being the volume average of the unknowns. Moreover, in the previous equation Rijk represents flux
balance on the current cell; is the pseudo compressibility factor (for more details about artificial
compressibility method, see Chorin (1967)).
In order to have a fully implicit scheme and to
obtain a divergence free velocity field, a dual (or
pseudo) timederivative (see Merkle and Athavale
(1987) for more details) is introduced in the discrete
system of equations


Vq
Vq
+
+ Rijk = 0
(19)
ijk
t ijk

u0i u0j

t = fv1 ();

m+1

Dt

2t

(22)

fv1 () =

3
(23)
3
3 + Cv1

= cb1 [1 ft2 ] S
 2
cb1 i
cw1 fw 2 ft2
(24)
k
d
+ ft1 U

1
+
(( + )
) + cb2 (
)2 ,

where S = S + [
/(k 2 d2 )]fv2 , S is the magnitude
of the vorticity vector, d the distance from the wall,
ft1 , ft2 , fw , fv2 are functions that depends only on
and the distance from the wall; finally, the c-s and k
are constants. The first two terms in the right hand
side represent production and destruction of , respectively; the third one is the so-called trip term,
that allows to specify the laminar-turbulent transition point location (in the results shown in the next
section, this term was always turned off); the last
part is a dissipation term, that contains also a non
conservative portion cb2 (
)2 which is responsible,
together with the non-linear part of the diffusion
term (

), for the advection of a turbulent
front into non-turbulent regions.

n1

Rm+1
ijk

The variable is computed from the solution of a


partial differential equation that reads

3 (Vq)ijk 4 (Vq)ijk + (Vq)ijk

u
j
u
i
+
xj
xi

t is computed by means of an intermediate variable


and the relation:

and the solution is iterated to steady state with respect to the pseudo time , for each physical time
step.
The time derivative in the previous equation is
approximated by means of a second order accurate
threepoints backward finite difference approximation formula, whereas the integration with respect
to the pseudo time is carried out by means of an
implicit Euler scheme, i.e.
m
(Vq)m+1
ijk (Vq)ijk

= t

=0

(20)
where the superscripts n and m denote the real and
dual time levels, is the pseudo time step and t
is the physical time step. The previous equation is
then solved with respect to qm+1
as in the Beam
ijk
and Warmings scheme (1978), i.e. the equation is
rewritten in delta form



Rijk
3
(Vq)m
(Vq)m
I+
ijk +
ijk =
2t
(Vq)
"
#
m
n
n1
3 (Vq)ijk 4 (Vq)ijk + (Vq)ijk
m

+ Rijk
2t
(21)
I being the identity matrix and (Vq)m
=
ijk
m+1
m
(Vq)ijk (Vq)ijk . The operator in the left hand
side of the previous equation is solved by an approximate factorization technique. The resulting
scheme is unconditionally stable to the linear analysis. Local dual time step ijk and a multi-grid
technique (Brandt, 1984; Favini et al., 1996) have
been used in order to improve the convergence rate
of the subiteration algorithm.

Free surface treatment and grid movement


The previous finite volume solver for the bulk flow
is coupled with a surface fitting approach. The free
surface is supposed to be a single value function
with respect to z, and therefore is explicitly written
in terms of a surface elevation function h(x, y, t) as
4

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

F (x, y, z, t) = z h(x, y, t). The kinematic boundary condition (4) can be rewritten as

Cb = 0.6 is considered for this test case. The numerical simulations are performed for Rn = 4 106 and
F n = 0.316 with a 4-block grid (the surface mesh
is reported in figure (1), where every other point is
shown); in the mesh used in the computation, 80
points are distributed along each side of the hull, 80
point normal to the wall (with y + = O(1) in the
boundary layer), 80 points girth-wise and 40 points
in the wake. The total number of cells is 768000 for
the fine mesh.

h(x, y, t)
h(x, y, t)
h(x, y, t)
+u
+v
=w
t
x
y
(25)
Like the Eulerian part of the RANS equations, (25)
is discretized by a ENO scheme. Once the shape
of the free surface is known, the grid is moved and
deformed to fit the actual elevation and the current
position of the ship hull. To this purpose, two grids
are used during the computation, a background grid
and the actual grid, which are used in the following
two stages

The prescribed motion of the hull consists in a


combination of a sinusoidal sway and yaw motions,
with a nondimensional period equal to 4. The amplitude of the oscillation is gradually increased, the
law of the prescribed motion being:

1. the background grid is moved rigidly with the


ship, in order to fit the current position of the
hull; this grid includes the region filled with
water and some space above the water plane.

t
yg (t) = y0 d(t) sin 2
4

2. The actual grid is generated by searching the


intersection of the the background grid with
the current free surface, and then by making
the grid points slide along the curvilinear coordinate lines of the background grid.


(26)

g (t) = 0 d(t) sin 2 +


4
4

where:
Z

d(t) = (1.0 e8.0t )4.0


Y

(27)

is a damping function.

yg (t/T)
g (t/T)

0.1

8
4

yg

0.05

-0.05

0.15

-4

-0.1

-8
0

t/T

Fig. 2: Ship in ZigZag manoeuvre: prescribed


PMM motion.

Fig. 1: S60 - PMM manoeuvre: mesh on the hull


and the water plane.

In the above equations, y0 = 0.1Lpp and 0 = 8.9


degrees are the amplitudes of the sway and the yaw
motions, respectively; the yaw motion has /4 phase
shift, so to have the hull always tangent to the trajectory at midship (see figure (2)). The time step
used in the simulation is t = T /200. The solution at steady state for the ship hull with uniform
forward speed is used as initial condition and the
integration is carried on until a regime solution (periodical solution in time) is attained.

NUMERICAL RESULTS
The algorithm described above is applied to the simulation of the unsteady flow around a ship hull with
a prescribed yaw and sway motion, which simulates
an experiment in the towing tank with the Planar
Motion Mechanism (PMM). A Series 60 hull with
5

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

0.01
14
0.008

12

0.006
8

Cx

V-Cycles

10

0.004

Fine
Medium
Coarse

4
0.002
2
0

0.25

0.5

t/T

0.75

0.25

0.25

0.5

0.75

0.5

0.75

t/T

0.1
0.075
1.0E-03
0.05

Cy

u 2

0.025
0

-0.025

5.0E-04

-0.05
-0.075

0.25

0.5

t/T

0.75

-0.1

Fig. 3: S60 - PMM motion; top: V-cycle per


sub-iteration; bottom: L2 norm of divergence

t/T

Fig. 4: S60 - PMM manoeuvre; time evolution of


drag for three grid levels. Top: drag coefficient;
bottom: side force coefficient

In the sub-iteration, the local pseudo time step


ijk is chosen to be four time larger than the stability limit of the corresponding explicit Euler scheme.
The implicit smoother is coupled with a four-level
multi-grid V-cycle with three smoothing steps per
level. The number of V-cycles required to reduce all
the residuals of three orders of magnitude is reported
in figure (3), together with the L2 norm of the divergence. It can be seen that the required convergence
level is attained within at most 10 subiterations.

Max
Min
Average
Amplitude

Coarse
0.0078
0.0053
0.0065
0.0025

Medium
0.0083
0.0045
0.0064
0.0038

Fine
0.0087
0.0041
0.0064
0.0046

Tab. 1: S60 - PMM: computed Cx on three grid levels

In figure (4) the solutions are reported for one


period of oscillation, after the transient has decayed.
The computation was performed
with three grid levels, the grid ratio being r ' 1/ 2 (and consequently
the coarsest was obtained from the finest one by removing every other point).

As expected, the lateral force has the same period of the enforced oscillation, whereas the frequency of the drag is two times larger. A phase
shift can be clearly observed from the plots for the
drag coefficient, whereas it is negligible for the lateral force coefficient.
6

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Acknowledgments

The results are summarized in table (1), where


the average, maximum and minimum values of the
drag coefficient are reported for the three solutions.
It can be seen that grid independence is almost attained with regard to the average value of the resistance; in particular, it is only slightly larger than the
measured value for the same ship in straight course
in the same condition of Froude and Reynolds numbers (see Toda et al. (1991)). Although the amplitude of the oscillation still depends sensibly on the
grid, its value is converging when refining the grid,
with an observed convergence order ' 1.42. Convergence was not observed, instead, with regard to the
lateral force coefficient, probably because the coarse
grid is too poor to give acceptable results.

This work was supported by the Italian Navy


through the project 6DoFRANSERP

References
Anderson, J. D. (1995). Computational Fluid Dynamics. McGraw-Hill, New York, USA.
Beam, R. and Warming, R. (1978). An Implicit Factored Scheme for the Compressible Navier-Stokes
Equations. AIAA Journal, 16:393402.
Brandt, A. (1984). Multi-grid Techniques: 1984
Guide with Application to Fluid Dynamics. The
Weizmann Institute of Science, Rehovot (Israel).

Free surface elevation during half a period of


the prescribed motion (after periodic solution is
achieved) is presented in figure (5). It can be seen
that the motion of the ship has a remarkable influence on the wave pattern.

Chorin, A. (1967). A Numerical Method for Solving


Incompressible Viscous Flow Problems. J. Comput. Phys., 2:1226.

The axial velocity contours at midship, stern and


two sections downstream are reported, during the
same half period of oscillation, in figure (6) and (7).
Again, the effect of the prescribed motion on the
boundary layer and the wake is clearly seen from
the plots.

Di Mascio, A., Broglia, R., and Favini, B. (2001).


A Second Order GodunovType Scheme for Naval Hydrodynamics, pages 253261. Kluwer Academic/Plenum Publishers.

CONCLUSION
TIVES

Godunov, S. (1959). A Finite Difference Method for


the Numerical Computation of Discontinuous Solutions of the Equations of Fluid Dynamics. Mat.
Sb., 47:271.

AND

Favini, B., Broglia, R., and Di Mascio, A. (1996).


Multigrid Acceleration of Second Order ENO
Schemes from Low Subsonic to High Supersonic
Flows. Int. J. Num. Meth. Fluids, 23:589606.

PERSPEC-

The computation of the flow around a Series 60 hull


with forced yaw and sway motion was performed by
means of the solution of the unsteady RANS equations. The numerical algorithm used for the simulation is a fully implicit finite volume scheme with
sub-iteration. The numerical solution for the PMM
test was computed on three grid levels and the convergence order estimated when possible. In particular, the solution was converging in terms of drag
coefficient, whereas unsatisfactory results were obtained with regard to the side force coefficient. This
must be ascribed to the grids used in the simulations, which were probably too coarse to yield grid
independent results.

Harten, A., Engquist, B., Osher, S., and


Chakravarthy, S. R. (1987). Uniformly High
Order Accurate Essentially NonOscillatory
Schemes. J. Comput. Phys., 71:231303.
Merkle, C. L. and Athavale, M. (1987). TimeAccurate Unsteady Incompressible Flow Algorithm
Based on Artificially Compressibility. AIAA paper, 871137.
Spalart, P. R. and Allmaras, S. R. (1994). A One
Equation Turbulence Model for Aerodynamic
Flows. La Recherche Aerospatiale, 1:521.
Toda, Y., Stern, F., and Longo, J. (1991). Mean
Flow Measuraments in the Boundary Layer and
Wake and Wave Field of a Serie 60 CB = 0.6 Ship
Model for Froude Numbers .16 and .316. IIHR
Report, No. 352.

Further developments will address the issue of


grid convergence and then validation of the results
by comparison with experimental data collected in
the towing tank.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

z
2.0E-02
1.8E-02
1.6E-02
1.4E-02
1.2E-02
1.0E-02
8.0E-03
6.0E-03
4.0E-03
2.0E-03
0.0E+00
-2.0E-03
-4.0E-03
-6.0E-03
-8.0E-03
-1.0E-02
-1.2E-02

-0.5

0.5

-0.5

0.5

2.0E-02
1.8E-02
1.6E-02
1.4E-02
1.2E-02
1.0E-02
8.0E-03
6.0E-03
4.0E-03
2.0E-03
0.0E+00
-2.0E-03
-4.0E-03
-6.0E-03
-8.0E-03
-1.0E-02
-1.2E-02

2.0E-02
1.8E-02
1.6E-02
1.4E-02
1.2E-02
1.0E-02
8.0E-03
6.0E-03
4.0E-03
2.0E-03
0.0E+00
-2.0E-03
-4.0E-03
-6.0E-03
-8.0E-03
-1.0E-02
-1.2E-02

t / T = 2/5

0.5

t / T = 1/10

0.5

-0.5
-0.5

-0.5

-0.5
0

0.5

t / T = 1/5

0.5

-0.5

0.5

2.0E-02
1.8E-02
1.6E-02
1.4E-02
1.2E-02
1.0E-02
8.0E-03
6.0E-03
4.0E-03
2.0E-03
0.0E+00
-2.0E-03
-4.0E-03
-6.0E-03
-8.0E-03
-1.0E-02
-1.2E-02

2.0E-02
1.8E-02
1.6E-02
1.4E-02
1.2E-02
1.0E-02
8.0E-03
6.0E-03
4.0E-03
2.0E-03
0.0E+00
-2.0E-03
-4.0E-03
-6.0E-03
-8.0E-03
-1.0E-02
-1.2E-02

t / T = 1/2

0.5

-0.5

/ = 3/10

0.5

t /T=0

0.5

z
2.0E-02
1.8E-02
1.6E-02
1.4E-02
1.2E-02
1.0E-02
8.0E-03
6.0E-03
4.0E-03
2.0E-03
0.0E+00
-2.0E-03
-4.0E-03
-6.0E-03
-8.0E-03
-1.0E-02
-1.2E-02

-0.5

-0.5
-0.5

Fig. 5:

0.5

-0.5

0.5

S60 - PMM: wave pattern during a half oscillation period

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

t / T =0

t / T =0

-0.1

-0.1

-0.1

0.1

-0.1

t / T =1/10

0.1

t / T =1/10

-0.1

-0.1

-0.1

0.1

-0.1

t / T =1/5

0.1

t / T =1/5

-0.1

-0.1

-0.1

0.1

-0.1

t / T =3/10

0.1

t / T =3/10

-0.1

-0.1

-0.1

0.1

-0.1

t / T =2/5

0.1

t / T =2/5

-0.1

-0.1

-0.1

0.1

-0.1

t / T =1/2

0.1

t / T =1/2

-0.1

-0.1

-0.1

0.1

-0.1

0.1

Fig. 6: S60 - PMM: axial velocity during a half oscillation period. Left: x/Lpp = 0.5 (midship),
right: x/Lpp = 1.0 (stern)

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

t / T =0

t/T=0

-0.1

-0.1

-0.1

0.1

-0.1

t / T =1/10

0.1

t / T = 1/10

-0.1

-0.1

-0.1

0.1

-0.1

t / T =1/5

0.1

t / T = 1/5

-0.1

-0.1

-0.1

0.1

-0.1

t / T =3/10

0.1

t / T = 3/10

-0.1

-0.1

-0.1

0.1

-0.1

t / T =2/5

0.1

t / T =2/5

-0.1

-0.1

-0.1

0.1

-0.1

t / T =1/2

0.1

t / T =1/2

-0.1

-0.1

-0.1

0.1

-0.1

0.1

Fig. 7: S60 - PMM: axial velocity during a half oscillation period. Left: x/Lpp = 1.1, right: x/Lpp = 1.2

10

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Takanori Hino
National Maritime Research Institute, Japan
Massive separation is expected to occur in
maneuvering conditions. Therefore, the standard
turbulence models may not have capabilities to
reproduce accurate flow fields. Do the authors plan
to employ advanced turbulence models in the future?
AUTHORS REPLY
We are aware of the problem of inaccurate
simulation of massive flow separation in ship
maneuvering by standard turbulence models.
Nevertheless, the simulation with simple models is
worth doing for verification purposes at least. When
addressing validation issues, the details of flow
separation must be considered with care to assess the
performances of the various turbulence models. This
will be the subject for the sequel of the present work.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Paper submitted to
Twenty-Fifth Symposium on Naval Hydrodynamics
August 8-13, 2004
St Johns, Newfoundland and Labrador , Canada

VALIDATION OF FORCES, MOMENTS AND STABILITY DERIVATIVES


OF A MANEUVERING SERIES 58 BARE HULL

Chao-Ho Sung, Bong Rhee and In-Young Koh


David Taylor Model Basin
NSWCCD, MD 20817-5700, U.S.A.

INTRODUCTION
Despite rapid progress in recent years in computer
memory and speed and numerical algorithms, the application of Computational Fluid Dynamics (CFD) to
submarine maneuvering problems falls far short of its
potential and expectations. CFD here is used to imply numerical techniques including the Euler solvers
and beyond. The two main reasons are first the long
computer run time required to obtain a viscous solution and second the lack of systematic validation efforts
to establish confidence in the accuracy of CFD results.
The purpose of this paper is to present the validation
results of a Series 58 bare hull both in straight flight
and in turning. The Reynolds-Averaged Navier-Stokes
(RANS) equations will be solved. This investigation
is the first stage of a systematic validation effort with
the ultimate goal of achieving reliable predictions of a
fully appended maneuvering submarine. In the future,
validations will successively progress from a submarine
with a sail, a submarine with stern appendages and finally a fully appended submarine.
The current status, practices and future research efforts related to submarine maneuvering problems will
be briefly discussed in order to illuminate the direction the present paper is leading to. The situation in
the aircraft community is very similar. The vehicle maneuvering simulation program (6 degrees of freedom of
motion) that currently applies to a submarine or an aircraft serves at least two purposes. One is to provide

ABSTRACT

A validation of forces, moments and stability


derivatives of a Series 58 bare hull in straight flight and
in rotation is presented. Convergence and grid independent solutions were investigated with three grid sizes.
It is concluded that a grid independent solution can be
obtained with a medium grid in the order of a half million grid cells in about 20 minutes in the straight flight
case and about 40 minutes in the rotating case. This is
not optimum. It is most likely that a grid independent
solution can be achieved with a much smaller grid size
therefore in a much shorter run time. The computer resources used are 8 CPUs of the IBM POWER 4 system.
The accuracy of the predicted normal force and pitch
moment coefficients and the longitudinal distribution of
the normal force is mostly within experimental errors in
the straight flight case. The range of pitching angles is
from 0 to 18. The accuracy in the rotating case can
not be definitely specified due to the scattering in the
measured data. The prediction of the stability derivatives of both the straight flight and rotating cases is in
good agreement with the measured values and potential flow predictions. This indicates that the predicted
slopes of forces and moments are in good agreement
with the measurement.
1

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

real time capability for man-in-the-loop trainers and the


other is to provide design analysis. The first application
has the very stringent time requirement that a simulation computation must be completed in real-time. The
other applications have a less stringent time requirement
but a simulation should still be completed within several
hours in order to have a quick return in practical applications. It is clear that beside solution accuracy, the
CPU time required to complete a simulation computation is an important consideration in developing simulation tools for a maneuvering vehicle.
It is important to have some ideas of the stateof-the-art CPU time requirement for the simulation of
a submarine maneuvering trajectory. For example, a
maneuvering submarine trajectory can be completed in
about two minutes in real time during a full scale trial.
For training purposes, the simulation of this trajectory
must be completed well within this real time. This is
very short but it can be accomplished by a simple potential flow method with viscous corrections at present.
The situation is quite different with the RANS simulation. A few years ago, Slomski (2004) made some
trajectory computations using RANS. A trajectory of
the fully appended ONR Body1 Radio Control Model
(RCM) took about 9 seconds real-time. Case 1 with the
propeller gridded required a total of 21,000 hours with
73 processors to simulate the trajectory while Case 2
with the propeller modeled by a disk force model required about 16,000 hours with 57 processors. The simulation times for both cases are several million times too
slow for the real-time (9 seconds) application in training
based on one PC. Allowing for parallel computations,
the wall clock times for Case 1 with 73 processors and
Case 2 with 57 processors are 287 and 281 hours, respectively. For design applications, these run times are
at least one to two orders of magnitude too long. If a
cluster of 10 PCs (compared to 57 and 73 processors)
is used, then the run time is 2 to 3 orders of magnitude
too long. From this argument, it can be concluded that:
1. The real-time application based on RANS is out of
the question for some time to come and the only feasible approach is based on potential flow methods with
appropriate viscous corrections. 2. The run time for design applications based on RANS is at least one to two
orders of magnitude too long using massive parallel supercomputers and 2 to 3 orders of magnitude too long
using a cluster of, say,10 PCs. Therefore the approach
based on RANS is not useful in practical design applications at present.
Because of its fast computation speed, the potential
flow method plays an important role in the maneuvering
problem at present. However, the potential flow method

has a fundamental flaw because vortices and flow features caused by viscous effects such as separation can
not be handled. Some ad-hoc approximations must be
made. Clearly, the approach based on RANS must be
adapted. In doing so, one emphasis is to establish confidence in accuracy through a systematic validation of
RANS solutions and the other emphasis is to speed up
RANS run times such that RANS solutions will be fast
enough for design applications.
To speed up conventional RANS computations by
2 to 3 orders of magnitude is a formidable task. Not
only the modern computer parallelization architecture
must be efficiently utilized but also numerical techniques must be significantly improved. For example,
the leveling-off of parallelization efficiency after about
30 processors can be corrected by the use of the hybrid MPI/OpenMP approach (Kiris et. al, 2000). Numerical techniques such as local refinement (Bai &
Brandt, 1987), higher-order accurate schemes (Kreiss &
Oliger, 1972 and Lele, 1992), new wall function method
(Shih, 2003) and vorticity confinement method (Steinhoff, 1994) etc. can be combined so that a much smaller
size of computational grid (thus shorter run time) can be
used to obtain high solution resolution. Recent research
efforts to achieve Textbook Multigrid Efficiency (TME)
in CFD (Brandt et al., 2002, Thomas et al., 2003) must
also be considered. The goal is to complete solutions
for general CFD problems in about 10 multigrid cycles.
This is a dramatic speed-up compared to the current
practice of requiring several hundred to several thousand multigrid cycles. Caughey and Jameson (2003)
are also collaborating to achieve the same goal. Another
new development in the Newton-Krylov method (Knoll
& Keyes, 2004) is also worthy of attention. The Newton
method is well known to be a very efficient matrix inversion method but it requires a relatively large computer
memory and therefore has not been popular. With the
rapid increase in memory size of modern computers, it
seems that increase in memory can be swapped for decrease in run time. Thus the use of the Newton method
for fast computation should be re-evaluated. Successful
implementation of the numerical techniques mentioned
above may provide the possibility to predict a submarine trajectory based on RANS within a few hours.
The major role of RANS in trajectory prediction is
to compute the six components of force and moment
coefficients. These six force and moment coefficients
are then used to solve the six degrees of freedom equations of motion (Feldman, 1979). Accurate prediction
of forces and moments is required not only at large angles of attack but also at small angles of attack. The
accuracy requirement at large angles of attack is for the
2

Copyright National Academy of Sciences. All rights reserved.

. . ..

Twenty-Fifth Symposium on Naval Hydrodynamics

obvious reason that the large forces and moments generated at large angles of attack can significantly affect the
ship motion. The accuracy requirement at small angles
of attack is essential for accurate computation of stability derivatives because they are normally defined at
small angles of attack and small angular velocity components (Feldman, 1995).
One of the main obstacle to achieve accuracy in
the prediction of forces and moments is the turbulence
modeling. In general accuracy begins to degrade as the
angle of attack increases beyond 8 or so mainly due
to inadequacy in Boussinesq-based turbulence models.
The computed eddy viscosity becomes too large at a
high angle of attack and the fluid velocity near the wall
becomes too fast resulting in underprediction of the
normal force and overprediction of the moment. It is
relatively easy to fix the problem in the zero-equation
type of turbulence models such as the Baldwin-Lomax
model. At a high angle of attack, the length scale is
overestimated leading to a large eddy viscosity. This
can be corrected by seeking the so-called second peak
of the wake function to reduce the length scale. Once
this correction has been made, the prediction of forces
and moments is quite reasonable (Sung et al., 1993). It
is harder to diagnose the two-equation turbulence models. One approach is to control the excessive generation of the turbulence kinetic energy near the wall or,
for the same effect, to promote the turbulent energy
dissipation. This can be accomplished by constructing some appropriate damping functions near the wall.
This is called the low Reynolds number correction. The
low Reynolds number correction of the k- turbulence
model constructed by Wilcox (second edition, 1998)
has been adopted in this paper. It appears to give accurate predictions.
The test case selected for this paper is the measured
data of forces and moments on a Series 58 bare hull
(Randwijck et al., 2000). Both the straight flight and
rotating cases were investigated at David Taylor Model
Basin experimental facilities. The straight flight tests
were performed on Towing Carriage 2 and the rotating tests were performed on the Rotating Arm Facility.
There are two unique features in this experiment. One
is that a set of the rotating data is the only one of its
kind available in the public domain. The other is that
not only were the total forces and moments measured,
but their distribution along the body was also measured.
This provides a more rigorous test of prediction accuracy. Since the total force and moment are integrated
quantities, it sometimes happens that correct prediction
is accidently obtained because of fortuitous cancellations of errors. The quality of the measured data for the

straight flight case appears to be good. This can be verified by the observation that the measured data at positive pitching angle are in good agreement with those
at negative pitching angle. Similar confidence does not
apply to the rotating case where data are scattered. The
scattering of data is most likely due to the effect of the
struts supporting the model. There are some concerns
whether the comparison between data and computations
should be presented for the rotating case because of this
scattering. Several reasons favor the presentation. Although there is an abundance of classified data for both
bare hulls and fully appended submarine models in rotation, the set of the rotating data of the Series 58 is the
only one of its kind available in the public domain as
mentioned earlier. Another reason is that accurate prediction of forces and moments of a body in rotation is
not only a challenging problem for RANS, it is also a
fundamental problem in the maneuvering prediction. It
is thought worthy of a presentation to encourage other
researchers to take the challenge. To ensure that the
RANS prediction of the Series 58 body in rotation is reliable, validations with several classified bodies of revolution have been performed to establish confidence. Another reason favoring the presentation is the following:
It is known that the potential flow method gives reasonable prediction of forces and moments except near the
stern where the viscous effect becomes dominant. This
observation is clearly illustrated in the comparisons of
the longitudinal distribution of the measured rotary normal force and pitching moment derivatives with the predictions by RANS and potential flow. Both RANS and
potential flow predictions are in good agreement with
measured data up to about 80% of the body. After 80%
of the body only RANS predictions are in good agreement with measured data.
The outline of the paper is as follows: first, the
experiment is described. The numerical methods and
the k- turbulence model used to solve the steady incompressible RANS equations in a rotating coordinate
system are briefly described omitting most of the details. For those interested in more details, references
are given. Convergence, grid independence and CPU
run time are discussed. The definition of stability and
the evaluation of its derivatives are given. Comparisons
between measured data and predictions by RANS, and
in some cases by potential flow, are discussed. Finally,
a conclusion is given.

DESCRIPTION OF EXPERIMENT
The experiments were performed with Series 58
Model 4621 reported by Randwijck et al. (February,
3

Copyright National Academy of Sciences. All rights reserved.

. . ..

Twenty-Fifth Symposium on Naval Hydrodynamics

NUMERICAL METHODS
The governing equations to be solved are the incompressible RANS equations in a steadily rotating coordinate system
ui
xi

 x uiu j  p i j 12 2x2  y2 i j
u
+2(  ri u j  x  x
i j 
ui
t

i
j

0.1

z/L

2000). The model length, L, is 15 feet with a model


length to maximum diameter ratio of 7.34. The center of gravity is located at x L = 0.4456. Both the
straight flight and rotating case were investigated at
David Taylor Model Basin experimental facilities. The
straight flight tests were performed on Towing Carriage
2 at speeds of 2, 4 and 5 knots corresponding to the
Reynolds number based on model length of 4.7, 9.4 and
11.7 million, respectively. The pitching angle was varied in increments of 1 from 0 to 6, and in increments
of 2 from 6 to 18. Above this point, only the angles of 30, 45, 60, 75, and 90 were tested. The
cases in the higher pitching angles range were not validated at this time. Normal forces and pitching moments
were measured. These values were then used to derive
experimental stability derivatives. In the straight flight
case, normal forces and pitching moments at both positive and negative pitching angles were measured. The
differences between them were small indicating good
quality of the data. The rotating tests were performed on
the Rotating Arm Facility at speeds of between 3 and 5
knots corresponding to the Reynolds number of 7.0 and
11.7 million, respectively. The non-dimensional pitching angular velocities tested were approximately 0.15,
0.2, 2.5 and 0.3. Tests were also performed with corresponding negative values of non-dimensional pitching angular velocities. Unlike the straight flight case
where the data with positive and negative pitching angles are very close, the rotating data scattered considerably. This is most likely due to the effects of the struts
supporting the model. Again, normal forces and pitching moments were measured and stability derivatives
obtained. To obtain a longitudinal distribution of the
normal force, the model was divided into 10 segments
of varying length as shown in Figure 1a. The normal
force on each segment was measured. Measurements
of the longitudinal distribution of the normal force were
made in both the straight flight and rotating case.

0.0
0

0.2

0.4

x/L

0.6

10

0.8

Figure 1a. Series 58 is divided into 10 segments. Force


measurement apparatus is installed in each segment for
independent measurement.

where x j and u j are the Cartesian coordinate and velocity component, respectively, p is the pressure p divided
by a constant density , r is a position vector, is the
steady rotation rate in the z-direction, i j is the Kronecker delta, is the kinematic viscosity and i j is the
Reynolds stress tensor.
The RANS code used in solving the governing
equations is called IFLOW (Sung, 1987 and Liu et al.,
1998) which has been developed as a production code
at David Taylor Model Basin. The numerical approach
is based on the artificial compressibility. The space
is discretized by a second-order accurate finite-volume
central difference scheme and the time is discretized
by a one-step 5-stage, explicit Runge-Kutta method.
A fourth-order artificial dissipation model is used for
stability. Various convergence acceleration techniques
such as multigrid, preconditioning, local time stepping,
implicit residual smoothing and bulk viscosity damping
have been implemented. Further detailed discussions of
some of the numerical methods used were given in a
previous paper (Sung et al., 2002). For parallel computing, a hybrid MPI/OpenMP has been implemented
to maintain good parallel efficiency even at very high
number of processors as mentioned earlier (Kiris et al.,
2000). Since standard two-equation turbulence models
have the tendency to overpredict the eddy viscosity at
high pitching angles, a low Reynolds number correction of the standard k- model is used. For a detailed
description of this model, see Wilcox (1998). Obtaining good convergence with a two-equation turbulence
model is a non-trivial problem. Some successful techniques have been described in a previous paper (Sung et
al., 2002). Discussions of how initial and boundary conditions were implemented for the flow in a rotating coordinate system were also given in that paper. Overall,
the emphasis throughout the development of IFLOW is
both in accuracy and speed.
4

Copyright National Academy of Sciences. All rights reserved.

. . ..

Twenty-Fifth Symposium on Naval Hydrodynamics

-1
X

48x 32x 32 grid


96x 64x 64 grid
192x128x128 grid

log10(RMS residual)

-2
-3
-4
-5
-6
-7

600

1200

1800

Multigrid cycles

Figure 1b. A typical computational grid for a Series 58 bare hull

CONVERGENCE AND GRID INDEPENDENT


SOLUTION
To ascertain that a computed solution is indeed a
solution, it is important to perform careful study in convergence and grid independent solution. For a practical
engineering problem it is almost impossible to reduce
the residual to machine zero. But at least three orders
of magnitude reduction in residual is desirable. Sometimes even this modest requirement can not be fulfilled
for some particularly difficult problems. In this case,
it is important to make sure that the solution will not
change much after calculation has been kept running for
a long time. It is also important to make sure that the
solution will not change much after a simple change in
the grid has been made by increasing the grid size.
A typical C-grid is shown in Figure 1b. The location of the inflow boundary is about one body length upstream of the bow and the location of the outflow boundary is approximately 5 body lengths from the tail. Because of symmetry along the centerline, for the straight
flight case only one-sided grids were used to reduce the
grid size but two-sided grids must be used for the rotating case. A total of three grid sizes were considered in
each case. In the straight flight case, there are a coarse
grid (48x32x32 0.05x106), medium grid (96x64x64
0.4x106) and fine grid (192x128x128
3.1x106).
The corresponding grid sizes for the rotating case
were: coarse grid (48x32x48 0.07x106), medium grid
(96x64x96 0.6x106) and fine grid (192x128x192
4.7x106). Note that due to more complex flow features
in the rotation case, the number of grid cells in the direction normal to the wall is significantly increased. It is

Figure 2a. Convergence history


of root-mean-square
of pressure
o
6
in the straight case at = 10 and Re = 11.7 x 10

well known that a smart distribution of grid cells is more


efficient than blindly increasing the grid size. Doubling
of the grid size in each direction is done here mainly for
the ease of demonstrating grid independent solution.
The convergence histories of the root-mean-square
residual of pressure for the straight flight and rotating
cases are shown in Figures 2a and 2b, respectively. The
convergence history of the multigrid starting procedure
is not included in the figures. This procedure obtained
the next-fine grid solution first and then interpolated it
to the fine grid to start the computation. This omission
explains why the convergence history starts somewhere
between -1 and -2 in the figures instead of zero. The
CFL (Courant-Friedrichs-Lewy) number used is 4.5 for
all cases at all time although CFL = 6.5 also works. The
distribution of the non-dimensional normal distance to
the wall of the first grid cell y is approximately 3-4 in
the coarse grid, 2-3 in the medium grid and 1-2 in the
fine grid. Although not shown here, y can be raised to
about 20 using the new wall function developed by Shih
(2003) without too much deterioration of accuracy.
In general the convergence is considered pretty
good in both the straight flight and rotating cases. It is
good to see that the convergence maintains monotonic
behavior even in the fine grid down to almost 6 orders
of magnitude in 1800 multigrid cycles. For the medium
grid, the convergence levels off after about 6 orders of
magnitude in 600 multigrid cycles. For the coarse grid,
the convergence levels off almost 7 orders of magnitude
in 500 multigrid cycles. The convergence for the rotating case is not as good because the flow is much more
complicated than the straight flight case. Convergence
5

Copyright National Academy of Sciences. All rights reserved.

. . ..

Twenty-Fifth Symposium on Naval Hydrodynamics

-1

186
48x 32x 48 grid
96x 64x 96 grid
192x128x192 grid

meas.: +10 o
meas.: -10 o
48x 32x 32 grid
96x 64x 64 grid
192x128x128 grid

4
6

0
2
4

-3

M
Z
1033
Z xx 10

log10 (RMS residual)

-2

-4
-5

0
2

-1

-2
0

-2-2
-4

meas.: +10
o
meas.: -10
48x 32x 32 grid
96x 64x 64 grid
192x128x128 grid

-6
-4

-3

-6

-8
-6

-7

600

1200

-10
-8
-4

1800

Multigrid cycles

histories of the normal force Z and pitching moment M


coefficients for the straight flight case at a pitch angle
of 10 are shown in Figures 3a and 3b, respectively.
Similarly, Z and M with a non-dimensional pitching
angular velocity q = 0.2 for the rotating case are shown
in Figures 4a and 4b, respectively. Since the computation starts from a uniform flow which is quite different
from the final converged steady solution, violent oscillations occur in the beginning of the computation. However, they quickly settle down and in the straight flight
case, both Z and M have reached the steady state solution in about 500 multigrid cycles. The convergence for
the fine grid is much slower, the steady state is reached
after about 1500 multigrid cycles. It is important to note
that all three grid sizes give approximately the same values of Z and M . The predictions given by the medium
and fine grids are consistently better than those obtained
by the coarse grid. Thus, it can be concluded that a grid
independent solution has been achieved by the medium
grid at about 0.5 million grid cells in 500 multigrid cycles. For economical computation, a reasonable solution probably can be obtained using a grid size somewhere between the coarse grid (0.05 million grid cells)
and the medium grid (0.5 million grid cells) in less than
500 multigrid cycles.

CPU RUN TIME


It is important to have an estimate of approximate
run time (wall-clock time) to achieve a grid independent solution. The computer used here is the IBM
POWER 4 system at the Department of Defense High

600
600
600

1200
1200
1200

1800
1800
1800

Multigrid cycles

Figure
Figure
3a.oConvergence
Convergence
history
history
of pitching
of normal
moment
force
Z
Figure 3b.
3b.
Convergence
history
Mcoefficient Z
o of
6
6 pitching moment
o
6
at
attack
at = 10 and
Re =of
and Re
Re == 12
12 xx 10
10
at angle
angle
of11.7
attackx4410and

506
meas.:
+10o o
meas.: +10
oo
meas.: -10
meas.:
-10
48x 32x
32x 32
48x
32grid
grid
96x 64x 64 grid
96x
64x 64 grid
192x128x128 grid
192x128x128 grid

40
2

30

MZ x 10
103 4

Figure 2b. Convergence history of root-mean-square


of pressure
in the rotating case at angular velocity q=0.2 and Re = 11.7 x 106

0
0

20-2
-4

10
-6

-8

-10
-10

600
600

1200
1200

1800
1800

Multigrid cycles
Multigrid
cycles
Figure3b.
3b. Convergence
of pitching
moment
Z
Figure
Convergencehistory
history
of pitching
moment
coefficient M
o
6
6
at angle
attack
4 and
Re = 12 x 10
at = 10o and
Re =of11.7
x 10

Performance Computing Modernization Office (DODHPCMC) at Naval Oceanographic Office (NAVO). Although this system has a total of 1328 CPUs, only one
node with 8 CPUs was used to avoid long waiting time.
Based on the computation with 8 CPUs in the medium
grid wtih 0.4 million grid cells in the straight flight case
and 0.7 million in the rotating case in 500 multigrid cycles, the run time is about 20 minutes for the straight
flight case and about 40 minutes in the rotating case.
This is not optimum. It is most likely that a grid independent solution can be achieved with a much smaller
grid size therefore in a much shorter run time. The run
time can be expected to decrease almost linearly as the
6

Copyright National Academy of Sciences. All rights reserved.

. . ..

Twenty-Fifth Symposium on Naval Hydrodynamics

1286

1286
o

meas.:
+10 48 grid
48x 32x
meas.:
-10 o 96 grid
96x 64x
48x 32x 32 grid
192x128x192
grid
96x 64x 64 grid
192x128x128 grid

4
6

8
2
4

2
4

0
2

33
Z xx 10
M
10 4

M
Z
10433
Z xx 10

0-20
o

meas.: +10
o
meas.: -10
48x 32x 32 grid
96x 64x 64 grid
192x128x128 grid

-4
-2

-4
-6
-4

-8

0
2

0-20
o

meas.: +10
o
meas.: -10
48x 32x 32 grid
96x 64x 64 grid
192x128x128 grid

-4
-2

-4
-6
-4

-8

-8
-6

-10
-8
-12

meas.:
+10 o48 grid
48x 32x
meas.:
-10 o 96 grid
96x 64x
48x 32x 32 grid
192x128x192
grid
96x 64x 64 grid
192x128x128 grid

4
6

-8
-6

0
0

400

600
600
800

1200

1200
1200
1600

2000

1800
1800
2400

-10
-8
-12

0
0

400

600
600
800

Multigrid cycles

Figure
Figure
4a.Convergence
Convergence
history
history
of pitching
of normal
moment
force
Z
Z
Figure 3b.
3b.
Convergence
history
Mcoefficient
6
o of pitching moment
6
o 0.2Re
6
at
of
== 12
at pitch angular
velocity
and
Rexx=10
at angle
angle
of attack
attackq44=and
and Re
12
1011.7 x 10

number of CPUs is increased. The run time estimates


here are most likely high because they are based on visual observations of the convergence histories shown in
Figures 3a, 3b, 4a and 4b as explained in the following.
In the straight flight case, it is reasonable to assume that
the experimental error is less than 10%. This level of accuracy definitely can be achieved when two significant
digits are obtained in the computed normal force coefficient Z and moment coefficient M . From the medium
grid computation, it was observed that three significant
digits had been obtained by Z soon after 100 multigrid
cycles. For M , two significant digits were obtained in
100 multigrid cycles but three significant digits had not
been obtained until about 400 cycles.

1200
1200
1600

1200

2000

1800
1800
2400

Multigrid cycles

Figure
Figure
4b.Convergence
Convergence
history
history
of pitching
of pitching
moment
moment
Z
coefficient M
Figure 3b.
3b.
Convergence
history
M
6
o of pitching moment
6
o 0.2Re
6
at
of
== 12
at pitch angular
velocity
and
Rexx=10
at angle
angle
of attack
attackq44=and
and Re
12
1011.7 x 10

Z
w

Zw

Z
w

 U1

Because the free stream velocity U is chosen as the


unit of velocity, U = 1. Following a similar derivation,
one obtains
M
w

Mw

 U1

If Z is taken to be the difference of Z at = 2


and = 0, the following approximations can be obtained

Zw 

Z 2o Z 0o 

2o 180
o

 Z2

2o 

180o

and
STABILITY DERIVATIVES AND STABILITY INDICES
Stability derivatives are essentially gradients of
forces and moments defined at an equilibrium state.
If the equilibrium state is defined as the state at zero
pitching angle and zero angular velocity, then stability
derivatives are defined at zero pitching angle and zero
angular velocity. This is the definition adopted here.
Consider the vertical plane stability in the coordinate
system fixed at the body, the velocity component in the
z-direction w is related to the free stream velocity U by

U sin

where is the pitching angle in radians. For small ,


w U cos U . Thus

Mw 

M 2o 

2o 180
o

because Z and M vary linearly with for small


values of .
The stability derivatives in the rotating case are also
defined for zero angular velocity q = 0. Unlike the
straight flight case, it is less certain to assume the gradients of forces and moments are constant in the neighborhood of q = 0. Nevertheless, it will be assumed that
the linear property still holds true. Then the stability
derivatives in the rotating case are computed according
to

Zq 

Z q 0 2Z q 0
q 0 2

Copyright National Academy of Sciences. All rights reserved.

 Z qq 0022




. . ..

Twenty-Fifth Symposium on Naval Hydrodynamics

and
-8

M q 0 2
q 0 2

In the following comparisons between computation and


measurement, the definitions given above are used to
compute stability derivatives.
It is well known that a body of revolution is dynamically unstable. Investigation of stability indices
becomes important only when appendages are added.
Nevertheless, the definition of the vertical plane stability index will be given below to illustrate how the stability derivatives defined above are applied in the stability
analysis. The vertical plane stability index is given as
(Feldman, 1995)

Gv

Mw

meas.: +
meas.: -
coarse: 48x32x32
medium: 96x64x64
fine: 192x128x128

-7

Zq m
Zw Mq

-6

Z x 10 3

Mq 

-4
-3
-2
-1
0

12

15

Pitching angle (deg)

18

21

Figure 5a. Comparison of calculated and measured normal force


coefficient on Series 58 barehull with k- turbulence model and
Re = 11.7 x 106

where the only variable which has not yet been defined
is the nondimensional mass m . Negative Gv indicates
instability and positive Gv implies stability. For a complete stability analysis, one needs to consider the horizontal plane stability index also. For detailed discussions, one is referred to Feldman (1995).

32
28

24
20
16

M x 10

FORCES AND MOMENTS IN THE STRAIGHT


FLIGHT CASE
The Reynolds number used in the computation is
11.7 million based on the body length. Although other
Reynolds numbers such as 4.7 and 9.4 million were also
tested, the differences are not significant. Figures 5a
and 5b show the comparison of calculated and measured
normal force coefficient Z and pitching moment coefficient M , respectively as the pitching angle increases
from 0 to 18. It can be seen that the computed Z
and M using the coarse, medium and fine grid are very
close to the measured data. The fine grid computation was not carried out for pitching angles of 16 and
18 because it is taking quite a long time to converge
at these high pitching angles. Also, the main interest
in the fine grid computation is to obtain very accurate
Z and M for the computation of stability derivatives
which are defined using small pitching angles. It is also
noted that the predicted Z by the coarse grid is underpredicted slightly and the predicted M is slightly overpredicted when the pitching angle increases beyond 8.
This symptom has been mentioned earlier in the discussion on the improvement of turbulence modeling. Apparently the current low Reynolds number correction on
the k- turbulence model needs improvement in coarse

-5

12
meas.: +
meas.: -
coarse: 48x32x32
medium: 96x64x64
fine: 192x128x128

8
4
0

12

Pitching angle (deg)

15

18

Figure 5b. Comparison of calculated and measured pitching moment


coefficient on Series
58 barehull with k- turbulence model and
6
Re = 11.7 x 10

grids. It is also possible that a coarse grid in the size


of 0.05 million grid cells should not be expected to obtain high accuracy because it is too coarse. Both the
calculated and measured Z and M are also given in
tabulated values in Tables 1a and 1b for the benefit of
those who wish to repeat the computations. Since the
uncertainty analysis was not performed in the measured
data, the error bands of data are not known. In addition, the measured values here were essentially digitized
from the figures in the report. For this reason, the measured data are mostly accurate to only one significant
digit and at most two significant digits. The computed
results are accurate to at least three significant digits.
8

Copyright National Academy of Sciences. All rights reserved.

. . ..

Twenty-Fifth Symposium on Naval Hydrodynamics

M x 10 4
1.0

TEST

0.5

Zx10 -3

-2.0

0.2

0.4

0.8

z/L

Segment location (x/L)


1

0.0
0

0.2

0.4

0.6

10

0.8

10.5

10.8

12.7

-8
-10

-16.0

12.7
16.0

-18.7

19.2

-22.0
-23.6

10
12

19.0

-12
-14

21.6

14

24.2

21.6
23.6

16

26.3

26.2

18

28.8

28.2

30

43.0

39.4

1
3

-0.136
-0.278
-0.433

2
4

-0.65

-0.607

-0.95

-0.806

-1.10
-1.75

-1.04

-6

1.10

-8
-10

1.70

2.50

-2.45

-2.41

-12
-14

3.50
4.65

10
12

-3.40

14

-4.60

-3.44
-4.68

-16

5.90

16

-5.80

-5.88

-18

7.30

18

-7.30

-7.40

30

-18.9

-1.62

8.0

4.0

0.0
-3

IFLOW

-0.15
-0.30
-0.50

-4.0

0.70

TEST

Zwx10

-4

0.25

-8.0

meas.
Potential calc.
RANS calc.:96x64x64 grid
RANS calc.:192x128x128

-12.0

-16.0

0.0

0.2

0.4

0.6

0.8

1.0

Segment location (x/L)

-18.9

Table 1a. Comparison of calculated and measured normal


force coefficient on Series 58 bare hull at pitching angle
and Re = 11.7 x 106

z/L

-2

TEST

16.2

Table 1b. Comparison of calculated and measured pitching


moment coefficient on Series 58 bare hull at pitching angle
and Re = 11.7 x 106

Z x 10 3

8.78

-12.2

-18

Figure 6. Longitudinal distribution of measured and calculated


normal force
coefficient in segment
on SERIES58 barehull
o
6
at = 10 and Re = 11.7 x 10

8.40

-6

-16

x/L

-8.50

-3

0.6

2.25
4.48
6.66

IFLOW

-4
test data = +10 (Ztotal=-2.45x10 )
o
-3
test data = -10 (Ztotal= -2.50x10 )
-3
96x64x64 grid (Ztotal = -2.41x10 )
-3
192x128x128 grid (Ztotal = -2.37x10 )

-1.5

2.00
4.40
6.50

-4.30

-0.5

TEST

1
-2

0.0

-1.0

0.0

0.2

0.4

0.6

0.8

10

x/L

Figure 7a. Longitudinal distribution of measured and calculated


normal force derivative
in segment on SERIES58 barehull
at Re = 11.7 x 106

FORCE DISTRIBUTION IN THE STRAIGHT


FLIGHT CASE
As mentioned earlier, the measurement of the longitudinal distribution of normal force (from which moment can then be obtained) is one of the unique features of this experiment. This can detect whether a good
prediction is accidental. Figure 6 shows the comparison of the normal force coefficients in 10 segments
along the body axis at a pitching angle of 10. The
prediction was made with the medium grid. Agreement is very good in every segment from the bow to the

stern. Note that the sum of the 10 calculated segmental


forces is 241  103 while the measured total Z s are
245  103 and 250  103 for pitching angles of
positive and negative 10, respectively. Comparisons at
a pitching angle of 4 were also made and similar good
agreement was obtained but are not presented here.
9

Copyright National Academy of Sciences. All rights reserved.

. . ..

Twenty-Fifth Symposium on Naval Hydrodynamics

6.0

5.0

4.0

0.7
0.6
0.5

2.0

1.0

0.0

0.2

0.4

0.6

0.8

1.0

Segment location (x/L)

Z x 10 3, M x 10 3

M wx10

0.4

3.0

0.0

test: Re = 11.7x106
6
test: Re = 10.5x10
test: Re = 9.4x106
Other tests
test: Sum of segments
6
calc.: Re = 12x10 , 96x64x96 grid
6
calc.: Re = 12x10 , 192x128x192 grid

0.8

meas.
Potential calc.
RANS calc.:96x64x64 grid
RANS calc.:192x128x128

0.3
0.2
0.1

0.0

-0.1
-0.2
-0.3
-0.4

z/L

-0.5
-0.6
1

0.0

3
0.2

0.4

7
0.6

8
0.8

10

-0.7
1

x/L

Figure 7b. Longitudinal distribution of measured and calculated


pitching moment derivative
in segment on SERIES58 barehull
at Re = 11.7 x 106

STABILITY DERIVATIVES IN THE STRAIGHT


FLIGHT CASE
Comparisons of the normal force and pitching moment stability derivatives are shown in Figures 7a and
7b, respectively. Potential flow results were also shown
for comparisons. These potential flow results were
taken from the report by Randwijck et al. (2000). With
the exception of the pitching moment derivative at the
first segment, the agreement between RANS results and
measured data are in general good at every segment. Potential flow results are in good agreement with the data
in the mid section of the bodies from segment 4 to segment 7. The agreement is not consistent in the nose region from segment 1 to segment 4. But the agreement is
poor in the stern region where viscous effects dominate
while RANS results are in very good agreement with
the data. This improvement is a very important support
for RANS to play a more important role in submarine
maneuvering problems.

FORCES AND MOMENTS IN THE ROTATING


CASE
As can be seen in Figure 8, there is a considerable
scattering of data among runs with the Reynolds number of 11.7, 10.5 and 9.4 million. The rotating arm experiment is by itself a more difficult experiment than
the straight flight case due to more complex physical
flow features. Part of the reason may also be because
the interfering effect from the struts which support the
model has not been properly minimized. To establish
some confidence in RANS results, validations of some
classified bodies of revolution in rotation have been per-

-0.8
-0.40

-0.30

-0.20

-0.10

0.00

0.10

0.20

0.30

0.40

Figure 8. Comparison of calculated and measured total normal


force and pitching moment coefficients at various angular
velocities at Re = 11.7 x 106 and center of gravity = 0.4456

formed and reasonable agreement has been obtained.


Most of the computations presented here are based
on the medium grid of 96x64x96 (0.6x106). The
Reynolds number considered is 11.7 million. The center of gravity is x L = 0.4456. Non-dimensional angular
velocity q at 0.13, 0.2, 0.25 and 0.32 were considered.
The comparison of calculated and measured total normal force and pitching moment coefficients is shown
in Figure 8. Because of the scattering in the measured
data, the accuracy of prediction cannot be conclusively
defined. However, the gradients (or slopes) of the computed Z and M appear to be about the same as those of
the sums of segments of test data as can be seen in the
figure. Medium grid (96x64x96 or 0.6x106) and fine
grid (192x128x192 or 4.7x106) results at q = 0.2 are
also shown in the figure to indicate grid independence.
The convergence of the fine grid has become very slow
requiring more than 2400 multigrid cycles but the convergence of the medium grid is quite fast. For example,
the accuracy of 2 significant figures is achieved in M after 200 multigrid cycles, 3 significant figures after 600
multigrid cycles.

STABILITY DERIVATIVES IN THE ROTATING


CASE
Comparisons of the rotary normal force and pitching moment stability derivatives are shown in Figures
9a and 9b, respectively. Potential flow results were also
shown for comparisons. The computed derivatives were
calculated with q = 0.2 as explained in an earlier section on stability derivatives and stability indices. In general, the agreement between RANS results and mea10

Copyright National Academy of Sciences. All rights reserved.

. . ..

Twenty-Fifth Symposium on Naval Hydrodynamics

2.0

0.13

1.16

M x10

meas.
potential calc.
RANS calc.:96x64x96
RANS calc.:192x128x192

4
1.0
-3

Z x 10

-1.32

0.0

M qx10

0.20

1.81

-1.97

0.25

2.47

-2.55

0.32

3.33

-3.27

-1.0

-2.0

0.0

0.2

0.4

0.6

0.8

1.0

Segment location (x/L)

z/L

Table 2. Calculated normal force and pitching moment


coefficients at various angular velocities (grid size:
96x64x96 and Re = 11.7 x 106)

0.0
0

0.2

0.4

0.6

0.8

10

x/L

Figure 9b. Longitudinal distribution of measured and calculated


rotating pitching
moment derivative in segment
on SERIES58

6
barehull at q = 0.2 and Re = 11.7 x 10

4.0
3.0

sured data is very good from the nose to the stern except an anomaly of computed Mq at segment 7. As in
the straight case, potential flow results deteriorated near
the stern where viscous effects become dominant. Despite the scattering of the measured data, the computed
and measured stability derivatives appear to have a good
agreement with each other. This is because the slopes
of Z and M of RANS are similar to those of measured
data as observed above. The good agreement with the
potential results except in the stern also provides support that the prediction is reliable.

and in rotation has been presented. Convergence and


grid independent solutions were investigated with three
grid sizes. In the straight flight case, there was a coarse
grid (48x32x32 0.05x106), medium grid (96x64x64
0.4x106) and fine grid (192x128x128
3.1x106).
The corresponding grid sizes for the rotating case
were: coarse grid (48x32x48 0.07x106), medium grid
(96x64x96 0.6x106) and fine grid (192x128x192
4.7x106). It is concluded that a grid independent solution can be obtained with a medium grid in the order
of a half million grid cells in about 20 minutes in the
straight flight case and about 40 minutes in the rotating case. This is not optimum. It is most likely a grid
independent solution can be achieved with a smaller
grid size therefore in a shorter run time. The computer
resources used in this paper are 8 CPUs of the IBM
POWER 4 system with a total of 1,328 CPUs. The
accuracy of the predicted normal force and pitch moment coefficients and the longitudinal distribution of the
normal force is in general within experimental errors in
the straight flight case. The range of pitching angles is
from 0 to 18. The accuracy in the rotating case can
not be definitely specified due to the scattering of measured data. The prediction of the stability derivatives
of both the straight flight and rotating cases is in good
agreement with the measured values. This indicates that
the predicted slopes of forces and moments are in good
agreement with the measurement.

CONCLUSION
A validation of forces, moments and stability
derivatives of a Series 58 bare hull in straight flight

ACKNOWLEDGMENT
This work is partially funded by the ILIR program at David Taylor Model Basin monitored by Dr.

2.0

0.0

Zqx10

1.0

-1.0
-2.0

meas.
potential calc.
RANS calc.:96x64x96
RANS calc.:192x128x192

-3.0
-4.0

0.0

0.2

0.4

0.6

0.8

1.0

z/L

Segment location (x/L)

0.0

0.2

0.4

0.6

0.8

10

x/L

Figure 9a. Longitudinal distribution of measured and calculated


rotating normal force derivative in segment on SERIES58 barehull
at q = 0.2 and Re = 11.7 x 106

11

Copyright National Academy of Sciences. All rights reserved.

. . ..

Twenty-Fifth Symposium on Naval Hydrodynamics

John Barkyoumb. Computer resources are provided by


The Department of Defense High Performance Computing Modernization Office (DOD-HPCMC) at Naval
Oceanographic Office (NAVO). After a long and distinguished career in the submarine maneuvering community, Mr. Tom Moran had the foresight to recognize
the importance of doing a simple thing, though not necessarily easy, correctly. His suggestion of this fundamental research topic is gratefully acknowledged. Discussions with Dr. Jerry Feldman are also gratefully acknowledged.

REFERENCES
Bai, D. and A. Brandt,Local Mesh Refinement
Multilevel Techniques, SIAM J. Sci. Stat. Comput., 8,
pp. 109-135, 1987.
Brandt, A., B. Diskin and J. L. Thomas,Recent
Advances in Achieving Textbook Multigrid Efficiency for Computational Fluid Dynamics Simulations, NASA/CR-2002-211656, ICASE Report No.
2002-16.
Caughey, D. A. and A. Jameson,Fast preconditioned multigrid solution of the Euler and Navier-Stokes
equations for steady compressible flows, International
Journal for Numerical Methods in Fluids, 43, pp.537553, 2003.
Feldman, J.,DTNSRDC Revised Standard Submarine Equations of Motion, DTNSRDC/SPD-0393-09,
June 1979.
Feldman, J.,Method of Performing Captive-Model
Experiments to Predict the Stability and Control Characteristics of Submarines, CRDKNSWC-HD-039325, June, 1995.
Knoll, D. A. and D. E. Keyes,Jacobian-free
Newton-Krylov methods: a survey of approaches and
applications:, Journal of Computational Physics, 193,
357-397, 2004
Kiris, C., W. Chan and D. Kwak,Parallel Unsteady
Turbopump Simulations for Liquid Rocket Engines,
2000, IEEE.
Kreiss, H. O. and J. Oliger,Comparison of Accurate Methods for the Integration of Hyperbolic Equations, Tellus 24, 1972.
Lele, S. K.,Compact Finite Difference Schemes
with Spectral-like Resolution, Journal of Computational Physics, vol. 103, No. 1, pp. 16-42, 1992.
Liu, C., X. Zheng and C. H. Sung, Preconditioned Multigrid Methods for Unsteady Incompressible
Flows, Journal of Computational Physics, vol. 139, 3557, 1998.
Randwijck, Eric F. and Jerome P. Feldman,Results

of Experiments with a Segmented Model to Investigate the Distribution of the Hydrodynamic Forces and
Moments on a Streamlined Body of Revolution at an
Angle of Attack or with a Pitching Angular Velocity,
NSWCCD-50-TR-2000/008, February 2000.
Shih, T. H., L. A. Povineli and N. S.
Liu,Application of Generalized Wall Function for
Complex Turbulent Flows, 2003.
Slomski, J. F., private communication, 2004.
Steinhoff, J.,Vorticity Confinement: A New Technique for Computing Vortex Dominated Flows, in D.
A. Caughey and M. M. Hafez, editors, Frontiers of
Computational Fluid Dynamics, John Wiley & Sons,
1994.
Sung, C. H.,An Explicit Runge-Kutta Method for
3D Incompressible Turbulent Flows, DTNSRDC/SH
-1244-01, July 1987.
Sung, C. H., M. J. Griffin, J. F. Tsai and T. T.
Huang,Incompressible Flow Computation of Forces
and Moments on Bodies of Revolution at incidence,
AIAA 93-0787, 31st Aerospace Siences Meeting & Exhibit, January 11-14, 1993, Reno, NV.
Sung, C. H. M. Y. Jiang, B. Rhee, S. Percival,
P. Atsavapranee and I. Y. Koh,Validation of the Flow
Around a Turning Submarine, The Twenty-Fourth
Symposium on Naval Hydrodynamics, July 8-13, 2002,
Fukuoka, Japan
Thomas, J. L., B. Diskin and A. Brandt,Textbook
Multigrid Efficiency for Fluid Simulations, Annual
Review of Fluid Mechanics 35, 317-340, 2003.
Wilcox, D. C., Turbulence Modeling for CFD,
DCW Industries, Inc. CA, second edition, 1998.

12

Copyright National Academy of Sciences. All rights reserved.

. . ..

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
George Watt
DRDC Atlantic, Canada
I agree with your response to one of the
referees that potential flow will not to the job:
- Normal force Z is zero in potential flow
- Pitching moment M is about 30% in error
in potential flow.
Your experiments breaking the forces up
into a distribution along the hull look very useful.
Are these data available in the open literature?
AUTHORS REPLY
All the test data such as normal force
coefficient and its longitudinal distribution presented
in this paper is readily available to those who wish to
repeat the computations.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Kevin L. Smith
Northrop Grumman Newport News, USA
First, I would like to thank the authors for
the opportunity to review and discuss their work.
The following comments are not solely my own, but
include those of Mr. Patrick Ryan and Dr. Wade
Miner, also of Northrop Grumman Newport News,
who are experienced users of Sungs IFLOW code.
As end users, rather than developers, we are
rarely afforded the opportunity or luxury of time to
thoroughly validate the capabilities of CFD tools, and
usually rely on the code developers for that function.
Unfortunately, since some of the codes that we use
were adapted from the aerospace industry, the
validation cases for those codes are frequently based
on geometry and conditions more appropriate for
aircraft.
Far fewer validation cases include
submarine-specific geometry or conditions (such as
very high Reynolds numbers, incompressibility, etc.)
much less the rigorous problems associated with
maneuvering. Therefore, it is greatly welcomed
when we see, in one place, a validation effort that
addresses both of these areas.
The authors correctly state the importance of
rapid convergence to industry, particularly in a
production environment that imposes tight schedules.
While real time solutions are unnecessary, typically
turn around times of several weeks or more cannot be
afforded. Our experience with the single-processor
version of their code has demonstrated that it does,
indeed, converge more quickly than many of our
other codes. Their efforts to validate IFLOW in a
multi-processor environment will further enhance our
ability to respond to time-urgent requests.
In addition, the subject of this paper (i.e. the
computational calculation of a submarines stability)
is also very timely. Future submarine capabilities
and mission requirements are driving us away from
conventional cylindrical hulls and toward nontraditional, non-circular hull shapes. One of the
primary concerns for these novel hull forms is their
vertical and horizontal plane stability. Unfortunately,
the further we deviate from circular hulls, the further
away we find ourselves from the vast compilation of
experimental force and moment data that has been
collected for submarine hulls over the past 50 or
more years. With no budget or time to support an
extensive test program, we are having to rely more on
computational methods to evaluate the stability of
new concepts. The results of this paper will increase
our confidence in those predictions.
We are also glad to see that the authors
plans include the analysis of a fully appended

submarine. However, as the solution time will


increase, so will the time required to generate the
computational grid, particularly if real-world details
are accurately modeled, such as the flap and rudder
gaps. To reduce the time associated with grid
generation, and further reduce the overall geometryto-solution time, we encourage the authors to include
the capability to handle unstructured or overset grids
into their code.
Finally, we must include a comment about
the actual geometry that was modeled. We are all
painfully aware of the pitfalls that can result from our
assumptions that the ideal geometry that we typically
develop for CFD models is actually representative of
the ships that we ultimately build and deploy. We
ignore small-scale features, more out of convenience
than necessity, with the assumption that modeling
them is not required to accurately predict the overall
flow field. However, recent experience has shown us
that small-scale features can, indeed, affect the flow
field around a submarine, especially when it is
maneuvering. Any asymmetry in the flow field can
manifest large forces, particularly since we are
integrating small pressures over large surface areas.
Because of this, we would recommend that
the authors repeat a limited portion of their analysis,
but with the CFD geometry modified to more
accurately represent the model that was tested, which
we would call modeling the model. In particular,
we would include the stings and struts that were
present in the experiment. Granted, these supports
should have minimal effect, as evidenced by the good
correlation between the predicted and measured
results, but including them in the CFD model simply
removes all doubt and ensures a fairer comparison to
the experiment. Likewise, the authors may want to
investigate whether the segment joints that were
added to measure the force distribution affected the
flow transition in the nose region during the
experiment something that was not mimicked in the
CFD model.
We would like to thank the authors for their
useful paper, especially their insights into grid size
reduction as well as the availability of the
maneuvering data. Their paper will become a
valuable addition to our library.

AUTHORS REPLY
As code developers, we always appreciate
very much comments coming from code users. These
comments are always taken seriously for future
improvement. From the CFD point of view,
modeling the model is the right thing to do in
general. In the case of the straight-run tests of the

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Series 58 body, the agreement with the data is so


consistent that modeling the model may not be
necessary. However, it is probably a meaningful
thing to do in the rotating case.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Thomas L. Moran
Computer Science Corporation, USA
It is with great pleasure that I comment on
the effort presented in this paper. I consider this work
to be a breakthrough in the application of RANS to
the maneuvering equations of motion development
process. But before I comment on that aspect, I
would first like to clear up a possible misconception
the authors may have given on the current simulation
code used at NSWCCD.
The Multi-Vortex simulation code is a semiempirical, geometry-based simulation utilizing twodimensional potential flow strip theory, discrete
vortex representation of hull separation, and lifting
line theory for the appendages. Vortices shed from
the hull and appendages are tracked and the resulting
forces and moments are also computed by strip
theory. Wakes from the hull and appendages also act
upon the stern appendages. The boundary layer is
modeled as a displaced body. The effect of the spanwise velocity distribution on an appendage is
modeled in the lifting line calculation for the
appendage.
The X-equation utilizes 4-quadrant
propulsor test data and coefficient-based terms to
represent propulsor forces and hull and appendage
drag. The simulation code computes the normal and
side forces and the rolling, pitching and yawing
moments. It contains parameters that allow the
forces and moments to be tuned to match captivemodel, RANS calculations, Radio Controlled Model
and full-scale data.
While it is true that the stability is computed
by potential flow, some viscous approximations have
been incorporated into the code, which have
accounted for, albeit semi-empirically, the
shortcomings of potential flow theory.
Secondly, this simulation code not only is
use for trainers and design support, but is primarily
used to develop recovery boundaries for a submarine
Submerged Operating Envelope (SOE) and flooding
recovery studies. These studies alone can require as
many as 10,000 simulations to develop the
boundaries. Clearly there is a need for an accurate
simulation that can run in real-time or faster.
Lastly, the simulation has the capability of
providing the shipyards with hull and appendage
loads needed for submarine structural design.
The US Navy is interested in developing a
simulation code for bodies with non-circular cross
section. Studies completed to date have shown that
the Multi-Vortex code can be successfully modified
for these hulls, provided test data or accurate RANS
calculations are available. This is the area where this

paper makes a breakthrough. As the authors point


out, potential flow theory does not do a credible job
computing the forces near the stern even for a body
of revolution. In the past, this was also true for
RANS computations. These results presented in this
paper are encouraging to me in that finally, perhaps
this shortcoming of RANS has been fixed and the US
Navy can proceed with the development and
validation of a simulation code for bodies with noncircular cross-section.
I would encourage the authors to perform
some additional correlation with not only the existing
database of circular cross-sections, but also the
extensive database for non-circular cross-sections.
Finally, I would like to comment on the
rotating arm data used in this paper. The Series-58
tests were conducted at a time when the rotating arm
testing techniques were in transition. It was a wellknown fact that vertical struts, while they were
acceptable for in-plane force and moment
measurements in straight-line test on carriage 2, were
not acceptable for the rotating arm due to interference
effects. It is believed that these Series-58 tests were
performed with a preliminary version of L-shaped
struts. One of the problems with these early L-shaped
struts was the tendency for the struts to deflect under
loads. I believe this is the main cause of the scatter
seen in these test results. I would encourage the
authors to perform some correlation with rotating arm
data conducted after 1985.
AUTHORS REPLY
We appreciate Toms valuable remarks. We
agree that it will be a useful effort to perform
validations on bodies with non-circular crosssections.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

not the same as the pressure predicted by the viscous


flow methods such as RANS. This is true near the
stern even at zero angle of attack.

DISCUSSION
Don Davis
Electric Boat, USA
In this paper, the authors present results for
normal force and pitching moment for a body of
revolution in straight flight and in steady rotation at a
range of pitching angles. The results that are shown
for both the normal force coefficient and the pitching
moment coefficient match the experimental data very
well for all three grid levels (slight variations are
evident in the coarse grid results. The convergence
behavior, as illustrated by the pressure residual and
the normal force and pitching moment iteration
histories, is also quite good. Based on these results,
the authors conclude that grid independent solutions
can be obtained with a medium grid in 20 minutes in
the straight flight case and in 40 minutes in the
rotating case. Although this conclusion may be valid
for pressure-based forces (i.e., normal force, pitching
moment), it is most likely not valid for shear-based
forces (e.g., drag). While the computed pressure may
be converged and grid independent, there is no
evidence here that the viscous part of the solution is
either converged or grid independent. If the reported
forces are dominated by pressure (even at the higher
angles), similar accuracy may be attainable with
lower order methods (e.g., potential flow). Since one
reason for using RANS over such lower-order
methods is its ability to directly capture physics in the
boundary layer, a RANS validation effort really
needs to include results that reflect such physics.
Thus, although the results shown here are quite good,
they exclude critical aspects of the flow, which may
affect the overall conclusions on grid requirements,
grid independence, and run time.
AUTHORS REPLY
The computations performed in this paper
are based on solving the Reynolds-averaged NavierStokes (RANS) equations which were supplemented
by a low-Reynolds version of the k- turbulence
model. The computed force and moment coefficients
are the sum of the effects due to the pressure and the
skin friction. Although the significant digit of the
measured data is only one and never higher than two,
the significant digit of the computed coefficients is at
least three. For this reason, it is certain that the
viscous effects are completely included within the
assumptions of RANS. The prediction based on
potential flow method alone cannot achieve the
desired accuracy at least near the stern as shown in
Figures 7a, 7b, 9a, and 9b. This is because the
pressure computed by the potential flow method is

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics


St. Johns, Newfoundland and Labrador, CANADA, 8-13 August 2004

Comparison and Validation of CFD Based


Local Optimization Methods for Surface Combatant Bow
E. F. Campana 1, D. Peri 1, Y. Tahara 2, F. Stern 3
( 1 INSEAN Italian Ship Model Basin, Rome, Italy, 2 OPU Osaka Prefecture
University, Japan, 3 IIHR Hydroscience and Engineering, University of Iowa, USA)
ABSTRACT

INTRODUCTION

Main objectives of the present paper are to describe the


development of two alternative Simulation Based
Design (SBD) environments based on complex CFD
analysis, to demonstrate their capability by applying
them to a real ship design problem and to give account
of the experimental campaign carried out to validate the
numerical results and assess the success of the
optimization.
The fundamental elements of a SBD framework are
analyzed and alternative components are described. A
derivative-free and a gradient-based local optimizers are
illustrated with emphasis on advanced strategies for the
use of CPU time expensive CFD solvers in the
optimization process, like the variable-fidelity/trust
region algorithm and a portable, multi-level parallel
framework. Alternatives in the geometry and mesh
manipulation techniques, to automatically adapt the
volume grid to the evolving shapes, are also illustrated.
As analysis tools for the evaluation of objective function
and functional constraints, two up-to-date free surface
fitting RANS solvers are used.
With these elements, two different SBD versions are
realized and demonstrated on a real design problem, the
optimization of the bow and sonar dome of the DTMB
5415, used as a typical example of a highly complex
ship redesign. Realistic functional and geometrical
constraints for preventing unfeasible results and to get a
final meaningful design are discussed and enforced. As
expected, being the two versions based on a local
strategy and constructed with different basic
components, their use resulted into two different final
designs, which however, beyond local differences,
clearly display common geometrical trends.
Finally, dedicated experimental campaigns for both
the optimized models have been carried out to validate
the computations and establish the success of the
optimization processes. The two optimized models
demonstrated improved characteristics beyond the
experimental uncertainty, confirming the validity of
both the SBD versions.

The paper describes the optimal shape redesign of the


bow of a naval surface combatant (the DTMB 5415)
and the experimental campaign carried out to assess the
success of the optimization, as a result of a close
interaction among IIHR - Hydroscience and Engineering
of the University of Iowa, the Osaka Prefecture
University (OPU), and the Italian Ship Model Basin
(INSEAN), in the framework of their respective NICOP
projects.
DTMB 5415 has been recently taken as a test case
by some optimization studies. The bow bulb redesign
was undertaken by Newman et al (2002) using
sensitivity analysis and complex variable finitedifference approach. The method used a RANS solver
to minimize sonar dome vortices neglecting free surface
effects. A finite-difference gradient-based approach was
followed by Tahara et al. (2000) for stern and sonar
dome optimization using RANS (with free surface
effects) and exploring the advantages of high
performance parallel FORTRAN. In a series of papers,
Peri and Campana (2001, 2003a,b) investigated a
variable-fidelity approach to speed up the optimization
process using free surface RANS in single- and multiobjective problems, while Peri and Campana (2003c,
2004) developed a global optimization algorithm,
applied to the solution of the same test.
Besides the few papers cited above, other recent
applications of CFD based optimization (Hino et al.,
1998, Minami et al., 2002) witness that optimal shape
design is receiving growing consideration in the naval
hydrodynamics community, filling the gap with other
fields (automotive, aeronautical, etc.) at a fast pace.
Nevertheless, previous works have been often
subject to criticisms raised by design engineers. The
most common criticism is attributed to the simplicity of
the optimization problem solved when compared to the
complexity of a real-life design problem. Additionally,
questions are often focused on the unrealistic final shape
obtained. This is most often a side effect of too simple
constraints adopted or too simple (low-fidelity) models
used as analysis tools, which may reveal themselves too
crude to successfully guide the optimizer toward a true

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

optimal design. Finally, the lack of experimental


evidence of the success of the optimization often shed a
shadow of uncertainty on the practical usefulness of the
whole procedure and on the maturity of the numerical
shape optimization.
Present work reports the development of two reliable
Simulation Based Design (SBD) environments that
involves a combination of simulation, modeling, and
design tools required for the design of complex systems.
The functional components needed to support SBD are
briefly analyzed and alternative components are
described. The developed tools are based on highfidelity analysis (free surface RANS solvers level) and
are applied to the optimization of an existing ship. The
Gothenburg 2000 workshop (Larsson et al., 2003) tested
up-to-date free surface RANS solvers and the DTMB
5415 was selected to represent navy designs. The two
analysis tools adopted in this paper, namely CFDSHIPIowa (Paterson et al., 2003, Wilson et al., 2004) and
MGShip (Di Mascio et al., 2000), classified in the
workshop as the best two codes on this test case,
demonstrating the accuracy of their prediction (Larsson
et al., 2003). The two SBD differ also for the
optimization strategy. One explores the capabilities of
the high performance computing, with a portable, multilevel parallelism for dynamic load balance, while in the
other the flow analysis solved on meshes of various
refinement provides a suite of integrated variablefidelity models used to reduce the computational effort
without loosing accuracy. Real-life constraints are
enforced during the optimization to provide realistic
optimal designs. Finally, results from the dedicated
experimental campaign are reported, demonstrating the
success of the optimization procedures and of the two
SBD environment developed.

CFD BASED LOCAL OPTIMIZATION


METHODS
To develop CFD-based optimization methods for shape
design, three main components must be built and are
common among many different applications (Fig. 1):
first, a method to solve the nonlinear optimization
problem formed by the objective function and the
constraints; second, a geometry modeling method to
provide a link between the design variables and a body
shape; and third, a CFD solver used as analysis tool to
return the value of the objective function and of
functional constraints. In the present study, two
approaches are investigated for each component.
Optimizers
In the numerical optimization of complicated systems
like ships, the use of low-fidelity analysis tools does not
guarantee real improvements in the detailed design
phase, especially when the margin for improvement is
small. Complex, high-fidelity CFD solvers such as free
surface RANS codes may then become necessary.

The drawback of this choice is that the use of these


expensive analysis tools in iterative procedures can be
prohibitively expensive, both in derivative-free and
derivative-based approaches.
This has motivated the development of two different
optimization strategies, described in the following
paragraphs, to reduce the computational effort.
Narrow Band Derivative-Free Approach
Genetic Algorithms (GA) are a stochastic search
technique that perform a multiple directional search by
maintaining a population of potential solutions. In the
present case a so-called Narrow Band flavor is used,
meaning with this that the population used is small and
that the bounds for the variables are narrow, so that the
search remains confined in a relatively small design
space, mimicking a derivative-free local optimization
technique.
The adopted algorithm proceeds as follows: (i)
generation of an initial population of individuals at
random manner; (ii) decoding and evaluation of some
predefined quality criterion, referred to as the fitness;
(iii) selection of individuals based on a probability
proportional to their relative fitness; (iv) crossover and
mutation. The steps (ii) through (iv) are repeated until
the generation achieves designated number. The
objective function is related to the fitness function, a
form of Sigmoid function in the present study, and
constraints are accounted for by using a penalty function
approach. GA is generally capable for finding global
minimum/maximum within the design space. In the
present study, the design space is relatively narrow
banded (i.e. narrow bounds are enforced on the design
variables), therefore, other advantages of GA are more
highlighted, namely the derivative-free nature and high
adaptableness for parallel computing.
In the present study, the conventional GA algorithm
has been extended for high performance optimization
method by introduction of parallel computing, i.e.,
Message Passing Interface (MPI) protocol (Tahara et al.
2004). Fig.2 shows the difference of GA architecture
between the present parallel and the conventional serial
computations. For the former case, processor 0 controls
overall GA procedure, and processors 1 through m,
where m is number of population, simultaneously
v
execute CFD method, i.e., evaluation of f ( i ) in the
figure. The parallel architecture offers advantage over
the serial architecture for considerably higher
computational efficiency, i.e., computational speed of
the former is nearly m times faster than that of the latter,
since most of CPU time is used for CFD method, and
communication
overhead
is
basically
small.
Furthermore, computational speed for GA in parallel
architecture does not depend on number of population
as well as number of design parameters.
Derivative-Based Variable-Fidelity Approach
When gradient-based optimization methods are
used, it is necessary to find a fast and accurate way of
calculating the gradient components, the most time
consuming portion of the design algorithm. The control

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

theory approach allows for dramatic computational cost


advantages over the finite-difference method of
calculating gradients, being effectively independent of
the number of design variables (Pironneau, 1984;
Jameson, 1988; for recent applications see Newman et
al., 2002 and Jameson and Kim, 2003). A comparison
among different sensitivity formulations and the finite
difference approach may be found in Valorani et al.
(2003). Automatic differentiation has been introduced
in the form of pre-compiler software tools, able to
generate new source code including the derivatives from
flow and grid generation codes (ADIFOR Automatic
DIfferentiation of FORtran, Bischof et al., 1992; TAF
Transformation of Algorithms in Fortran, Giering and
Kaminski, 1998). However, even if automatic
differencing compilers exists, their use has some
limitations: (i) they require an appropriate code
preparation (Thomas et al., 2003), that is a
development phase on the source code which is not
always a negligible effort and seldom it is openly
recognized; (ii) being necessary to modify the source
code only in-house CFD solvers can be used, and (iii)
any change in the number or type of design variables
and/or objective functions require a new adjoint code.
An alternative approach is explored in the present
paper, the variable-fidelity / first-order trust region
method proposed by Alexandrov (1996). The idea of
using low-fidelity, inexpensive models, together with
occasional recourse to high-fidelity, more expensive
models (for monitoring the progress of the algorithm)
has been used in engineering for long time. Recently
(Alexandrov 1996, Alexandrov et al. 2000) proposed a
general approach for controlling the use of variablefidelity methods by using a trust region control,
described in the following for completeness.
Lets suppose to have two models for computing our
objective function f and that they can both be used as
analysis tool: a simple low-fidelity model (flo),
computationally cheap, and a high-fidelity model (fhi)
more accurate but also more expensive. We can define
their ratio as a scale factor b(x), function of the design x,
the vector which contains all the design variables:
( x) = f hi ( x) f ( x)
lo
We can use this scale factor to improve the quality of
the lower approximation model. To this aim we need a
model for b(x) and this can be easily build by a firstorder local approximation bloc(x) of b(x) around the
current design xk:

loc ( x ) = ( x k ) + ( x k )T ( x x k )
This local approximation is then used to correct (to
the first-order) the value obtained with the low-fidelity
model flo:
f hi ( x) = ( x ) f lo ( x ) a ( x ) = loc ( x ) f lo ( x )
where a(x) is now the improved approximation which
satisfies the following consistency conditions:
a ( x k ) = f hi ( x k )

, a ( x k ) = f hi ( x k )

However, one has to keep in mind that the


approximation a(x) is accurate up to first order on the
base of local quantities and that cannot be applied in the

whole design space. Since we are progressively moving


away from xk during the optimization cycles, a trustregion approach must be therefore adopted in order to
determine the region (of radius r) in which we can still
trust the correction factor bloc(x) build at the design xk.
A general algorithm has the following general form: at
the k-step,
1. Compute the gradient of the corrected lowfidelity model a ( x k ) ;
2. Using a ( x k ) solve the optimization problem
and obtain the descent step sk under the
constraint s k r ;
3. The new design is xk+1 = xk + sk;
4. Measure the progress made: compute the ratio
between the expected and the real
improvement at the new design xk+1:
5. Analyze R:
f ( x k ) f hi ( x k +1)
R = hi
f hi ( x k ) a( x k +1)
a value close to 1 means that the model a(x)
predicts the actual behavior exceptionally well
and that the trust region radius r can be
increased;
- a value far from 1 means that the correction
factor bloc(x) is no more reliable and causes the
step to be rejected. The trust region radius r has
to be reduced (say, halved). Eventually the
radius may become too small and both the
local approximation loc(x) and the corrected
low-fidelity model a(x) have to be recomputed;
- the step is accepted otherwise, without
altering the r value.
A conjugate gradient algorithm is applied for the
descent direction computation, while during the line
search, limited by the trust-region radius, is performed
by means of the golden section method. The estimation
of the gradient with the expensive fhi model is only
required when the correction a(x) has to be newly
computed, since b(xk) is needed. The first formulation
for the correlation law (scale factor) is due to Chang et
al (1993) and is the one reported here. A flowchart of
the optimizer is reported in Fig. 5. Flexibility,
effectiveness and easiness of implementation are strong
points of this algorithm. Several versions exist
depending on the choice of the variable-fidelity models:
variable grid density, variable iterative accuracy,
variable physics. Finally, it should be underlined that
the final optimized shape obtained is the same as if a
high-fidelity model was used during all the process.
Indeed, the consistency condition ensures the quality of
the lower fidelity model (telling us when the local
approximation cannot be trusted anymore and has to be
re-build) and hence of the obtained minimum.

Geometry and grid manipulation


The geometry modeling is another important component
of CFD-based optimization methods. An efficient and
flexible way to modify the geometry of the body and the

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

computational body-fitted grid is necessary for a full


investigation of the design variables space and a
successful optimization. Once the vector with the new
design variable values is obtained by the optimization
algorithm, we have to spread the deformation over the
body surface and the volume grid. The most straight
forward method is superposition of several basic forms
or expanding/reducing basic geometry. For design of
aircraft, surface modeling based on B-Spline or other
similar methods were used relatively early. Another
capable method for geometry modeling is through
application of CAD systems. In the present study, the
two approaches are investigated: i.e., Geometry MethodA and B for (a) the CAD-based approach, and (b) the
additive perturbation method, respectively. Once the
ship surface is modified, the volume grid around the
hull should change accordingly with a simple adaptive
algorithm. The same grid manipulation is used in
association with Geometry Method-A and B.
CAD based
To modify the ship geometry, a CAD-Based hull form
modification method will be adopted. Two approaches
are possible, i.e., CAD direct control and CAD
emulation approaches. The implementations into
optimization environment are shown in Fig.3 and 4,
respectively. In the former, optimizer directly executes
CAD macro file in which procedures are pre-described
for hull form modification and evaluation of constraint
function. In the latter, a module is implemented in order
to emulate CAD operation based on NURBS surface
modeling. The latter approach offers an advantage over
the former for complete independence from CAD
system, i.e., designers are able to use any CAD system
and give/receive initial/optimized hull form geometry in
IGES format data. Displacement of the NURBS control
points can be design variables of the optimization
problem.
Additive Perturbation method
With this approach, the use of CAD is avoided and the
deformation of the shape is defined and controlled by
using a few control points, much less than the number
of nodes used for the discretization adopted for the flow
analysis. Mesh movement through the optimization
cycles is enforced in an explicit and simple way and
details can be found in Peri and Campana (2003b).
The original ship geometry, usually given in a CAD
format (IGES file), is easily translated into a standard
ascii format (e.g. plot3d) that constitute the input for the
optimization code. Final geometry (the output of the
optimization code) is easily translated back into the
IGES format, ready for the milling machine.
The first step of the procedure is to modify the hull
surface, according with the needs of the optimizer.
Geometry handling is simply performed by
superimposing one or several Bezier surfaces to the
original ships geometry. Each Bezier surface is
controlled by a given number of control points that are
used as design variables by the optimizer, and it is
acting on the whole geometry or on a part of it. Number

and position of the surfaces and number of control


points per surface can be changed in an easy and
flexible way, depending onto the details of the assigned
problem.
Grid manipulation
During the optimization, the grid is updated at every
optimization cycle as the hull form is modified. This is
accomplished by the use of an algebraic scheme to
increase the computational efficiency. A similar
approach was used by Tahara et al. (2000); however,
more simplified scheme was found effective in the
present study. The method is described in the following.
After an initial grid is generated, the geometrical
information is computed and stored in the memory, that
is as follows:
P = S 1 ( 1, 2 , 3 )

2 1 2 3
Q = S ( , , )
3
R = S ( 1 , 2 , 3 )

where P, Q, R are grid clustering and stretching


functions defined in the (1,2,3) directions,
respectively. More specifically, those are normalized
metric of (1,2,3) coordinates, such that 0 S i 1 ,

and S i = 0 and S i = 1 for i =1 and i =imax,


respectively. The grid points for the original geometry
are already defined in computational coordinates, i.e.,
x = x0 ( 1 , 2 , 3 )

1 2
3
y = y 0 ( , , )
z = z ( 1 , 2 , 3 )
0

and the hull surface is expressed as


x = x0 ( 1 ,1, 3 )

1
3
and
y = y 0 ( ,1, )
z = z ( 1 ,1, 3 )
0

x = xm ( 1 ,1, 3 )

1
3
y = y m ( ,1, )
1
3
z = z ( ,1, )
m

2
where is taken to be normal direction to the surface,
and values with subscript 0 and m correspond to the
original and modified hull surfaces. The grid points at
the outer boundary is fixed and given by
2
x = x0 ( 1 , max
, 3 )

1 2
3
y = y0 ( , max , )
z = z ( 1 , 2 , 3 )
0
max

In the optimization procedure, the hull surface is


modified but other computational boundaries. In the
work of Tahara et al., (2000), all grid points are
relocated using P, Q, and R when the surface is
modified, and an iterative manner is used to complete
the procedure. On the other hand, simpler grid
relocation method can be applied if the modification is
assumed to occur in local scale. That is, the method is
based on only Q and simply written as,

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

x =

y =

z =

x 0 ( 1 , 2 , 3 ) + ( x m ( 1 ,1, 3 ) +
x 0 ( 1 ,1, 3 ))( 1 S 2 ( 1 , 2 , 3 ))
y 0 ( 1 , 2 , 3 ) + ( y m ( 1 ,1, 3 ) +
y 0 ( 1 ,1, 3 ))( 1 S 2 ( 1 , 2 , 3 ))
z 0 ( 1 , 2 , 3 ) + ( z m ( 1 ,1, 3 ) +
z 0 ( 1 ,1, 3 ))( 1 S 2 ( 1 , 2 , 3 ))

Although the method is relatively simple and


straightforward, it was found that the grid quality is
nearly equal in the present application.
CFD
For an advanced fluid dynamic redesign of some part of
an existing shape, accurate analysis tools are necessary
for guiding the optimizer toward improved solutions.
This is true also for ship redesign and the most
advanced analysis tools available today to design
engineers are RANS solvers.
The degree of reliability of free-surface RANS code
has constantly matured during the last 10 years, and this
can be understood by looking at the series of the
international workshop on the numerical prediction of
the flow around a ship using viscous solvers started in
Gothenburg in 1980. The Gothenburg 2000 workshop
was focused on testing three modern hull forms, among
which the DTMB5415 was selected to represent navy
designs. Verification and validation procedures were
performed and the workshop showed that total
resistance, wakes and free-surface waves may now be
well predicted by some of the best codes. CFDSHIPIowa (Wilson et al., 2000) and MGShip (Di Mascio et
al., 2000), the two analysis tools adopted in this paper,
were identified as the best two codes on the DTMB5415
tests case as reported by Larson et al. (2003). Details of
the method were already provided in many references,
hence only overview is given in the following.
CFD-SHIP Iowa
The CFDSHIP-Iowa version 3.02 (for more details see
Paterson et al. 2003, Wilson et al., 2004) solves RANS
equations
for
unsteady,
three-dimensional
incompressible flow by using the higher-order upwind
difference method, a projection method for velocitypressure coupling, and the method of lines. The grid is
updated at each time step to conform to both the body
and free surfaces, where exact nonlinear kinematic and
approximate dynamic free-surface boundary conditions
are imposed. The k-omega turbulence model is used to
effect the closure.
MGShip
The free surface RANS solver used to perform the
simulations at INSEAN is referred to as MGShip (Di
Mascio et al., 2000). MGShip is a multigrid (FASFMG) multiblock structured grid code which solves the
Reynolds Averaged Navier-Stokes (RANS) equations.
The version adopted here uses a surface fitting
approach to compute the wave pattern. The

mathematical model is based on a pseudo-compressible


formulation of the RANS equations, approximated in
the discrete formulation by a finite volume technique. A
second order ENO-type scheme is adopted for nonviscous terms, while viscous fluxes are computed by a
standard centered finite volume approximation.
MGShip has been extensively validated and is
currently used by the It. Navy, as well as ship
(aeronautical) industries for hydro- (aero-) dynamic
simulations. More information on MGShip may be
found in the cited literature.
Integration of optimization components and
Simulation Based Design environment
Finally, the above-mentioned optimization components
are integrated to yield two optimization approaches, i.e.,
SBD-A and SBD-B. Optimizations using those
environments are demonstrated at OPU and INSEAN,
respectively. In the following, those approaches are
summarized in association with additional information
regarding the simulation based design environment.
SBD Version A
The SBD-A is composed by the Narrow-Band
Derivative-Free Approach, the CAD-Based and grid
manipulation method, and the CFD-SHIP Iowa RANS
solver version 3.02. For the present application, the
System-2 (CAD Emulation) is selected for the geometry
method.
The integrated system also involves the SMP code to
evaluate the Response Amplitude Operators (RAOs),
which are part of constraints described later. The
computations are performed on 64 CPU PC Cluster
(Intel Xeon 2.4 GHz x 64; Tahara et al., 2004) which is
recently designed by the authors for the present research
project.
SBD Version B
In the present version of the SBD-B, the optimizer adopt
the variable fidelity approach in the variable-grid
formulation. Indeed, being the RANS solver MGShip
based on a multigrid technique, we already have a suite
of meshes that could be used to this aim. The finest and
the second finest grid have been assumed as the highand the low-fidelity models.
The SBD-B is composed by the Derivative-Based
Variable-Fidelity Approach, the Additive Perturbation
Method with the grid manipulation described before,
and the MGShip RANS solver.
As for the SBD-A, the SMP code is used to evaluate
the heave and pitch RAOs. The computations are
performed on a Pentium IV (1.7GHz).

SURFACE COMBATANT BOW


OPTIMIZATION PROBLEM
For the complete definition of the design problem to be
solved, the following fundamental items must be
precisely addressed: (1) Selection of an initial design to
be optimized and of the extension of the modifiable

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

region; (2) Choice of the objective function to be


minimized plus number and position of the design
variables; and (3) Type and quantity of the constraints
of the problem. All these items are described in the
following.
As already stated, both the SBD versions are tested
on this problem. However, due to the multimodal
structure of the problem (which may contain many local
minima), and also due to the different parametrization
strategy (that changes the correspondence between ship
geometry and design parameters), two different final
shapes are expected.

design variables, i.e., dbt; dst and dsl, which control the
present hull modification.
For SBD-B, the control points of the patches used in
the Additive Perturbation method are the design
variables of the optimization problem. For the hull
shape parameterization 11 design variables have been
used: 2 for the side modification of the region above the
dome, 4 for the side modification of the dome, 3 for the
longitudinal modification of the dome and 2 for the
modification of the keel line below the dome.

Initial Design

To introduce elements of a realistic design problem,


seakeeping performances (peaks of the heave and pitch
RAOs for head seas) and a function of the axial vorticity
in a region behind the dome are chosen as functional
constraints. Geometrical constraints are imposed on the
design variables, the sonar dome volume, the bow entry
angle, the displacement and principal dimensions of the
ship. Complete definition of the problem, objective
function and constraints, is given below in tab. 1.

The objective function F to be minimized is the total


resistance RT at a speed of Fr=0.28 and Re=1.67x107 at
model scale. Design variables are used to explore the
design space, and changes in their values correspond to
different ship design. As a consequence, there are
closely connected with the specific technique adopted to
modify the geometry of the ship and the computational
mesh. As already described, those are closely related to
the geometry manipulation method which is
implemented in the overall optimization scheme.
For SBD-A, displacements of the NURBS control
points in Geometry Method-A (CAD-Based Approach)
can be the design variables of the optimization problem.
In the present study, design actions of widely used CAD
tools are emulated, so that control points define the bow
and sonar dome are moved in confined direction. That
is, control points for bow part move in only transverse
direction with dbt; and those for sonar dome in
transverse and longitudinal directions with dst and dsl,
respectively. Control points are adequately grouped in
order to avoid unrealistic shape modification. Control
points are adequately grouped in order to avoid
unrealistic shape modification. The bow and sonar
dome groups include 15x14 and 15x7 points,
respectively. All points in a group move with equal
displacement in the same direction. Therefore, the
above-mentioned design actions are related to three

Objective
function

Definition

min RT (Fr, R e)

Note
Bare hull, fixed model,
Fr = 0.28,
Re = 1.67*107

, peak heave RAO

x R Ndv

S C = 0.5 3 + 0.5 5 1 3

on
seakeeping
Functional constraints

Objective Function and design Variables Definition

Type

on sonar
dome
vortices

Bow entry
angle
Geometrical constraints

The initial design is Model 5415, which was conceived


as a preliminary design for a Navy surface combatant
ca. 1980. The hull geometry, available in IGES format,
includes both a sonar dome and transom stern.
Propulsion is provided through twin open-water
propellers driven by shafts supported by struts. There is
a large EFD database for Model 5415 due to an
international collaborative study on EFD/CFD and
uncertainty assessment between IIHR, INSEAN and the
Naval Surface Warfare Center, Carderock Division
(NSWC) see Stern et al. (2000). The validation data
includes boundary layer and wake, longitudinal wave
cuts, bow and transom wave fields, and wave breaking.
In the present study, the modification region is the bow
and the sonar dome, shown in Fig. 9. Bare hull is
considered.

Functional and Geometrical Constraints

Sonar dome
dimension
Sonar dome
position
Main
dimensions
Displacement

3p
5p
3o
HC=
1.02
3p
o
P C = 5 1.02
5p

1
N

( ox)i2

1
N

i
N

1,
p
( x )i2

5 , peak pitch RAO

o, optimized; p, parent
All the quantities
computed for * 0.4
RC is a circular region
placed at x = -0.30,
centered at y=0.02,
z = -0.07, with radius
r = 0.018.
o, optimized; p, parent

i RC
maximum
amplitude
variation of 5
A sonar array of
radius Rs and
height Hs should
fit inside the dome
Maximum forward
position
Lpp and depth
fixed
Maximum
variation 2%

2.5 per side


Hs=3m, Rs=2.5m
in ship length

Tab.1 - The nonlinear constrained optimization problem

GRIDS AND NUMERICAL


UNCERTAINTY ASSESSMENT
In use of CFD methods, uncertainty assessment must be
provided for the solutions and computational grid. CFD
uncertainty assessment consists of documentation,
verification, and validation. An overview of the

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

verification and validation procedure for the RANS


codes used in the present study is given in the
following.
Simulation uncertainty US is divided into two
components, one from numerics USN and the other from
modeling USM. The USN is estimated for both point and
integral quantities and is based upon grid and iteration
studies which determine grid USG and iterative USI
uncertainties. A root sum square (RSS) approach is used
to combine the components and to calculate USN, i.e.,
USN2=USG2 +USI2. CFD validation follows the method of
Stern et al. (2001) and Wilson et al. (2001), in which a
new approach is developed where uncertainties from
both the simulation (US) and EFD benchmark data (UD)
are considered. The first step is to calculate the
comparison error E which is defined as the difference
between the data D (benchmark) and the simulation
prediction S, i.e., E=D-S. The validation uncertainty UV
is defined as the combination of UD and the portion of
the uncertainties in the CFD simulation that are due to
numerics USN and which can be estimated through
verification analysis, i.e., UV2= UD2+USN2. UV sets the
level at which the validation can be achieved. The
criterion for validation is that |E| must be less than UV.
Note that for an analytical benchmark, UD is zero and
UV is equal to USN. Validation is critical for making
improvements and/or comparisons of different models
since USN is buried in UV.
For two optimization methods, i.e., SBD-A and B,
information of the uncertainty associated with
computational grid is provided in the following.
SBD-A
In order to reduce computational load during the
optimization cycles, a coarse grid is used to evaluate
objective and constraint function. Then, a fine grid is
used to verify and represent the optimal solutions. The
sizes of those grids are 238,760 and 1,779,648 for
coarse and fine grids, respectively. Using the same
structure of computational grids, Wilson et al. (2000)
had completed verification and validation results for
steady DTMB 5415 simulation for CFDSHIP-Iowa
version 3.02. Grid convergence was studied by
performing
steady
simulations
using
three
computational grids with refinement r = 2 in each
coordinate direction, i.e., of fine, medium, and coarse
grids. The grid sizes of those grids are 1,779,648,
654,192, and 238,760, respectively. Iterative uncertainty
USI was assessed through examination of the iteration
record of integral and point quantities and was taken as
one-half the range of the quantity of interest once initial
transients had died out. That is the preferred metric
because USI can be directly evaluated. Referring to the
values of Wilson et al. (2000), finally, UV for the present
computational grid is evaluated, i.e., those for total
resistance coefficient, wave profile, and wake filed are
about 2.8 % , 0.3 %, and 2.9 %, respectively.
SBD-B
For SBD-B, the same approach as that of SBD-A is
used, i.e., a coarse grid is used during the optimization

cycles, then a fine grid is used to verify and represent


the optimal solutions. The sizes of those grids are
253,952 and 1,712,128 for coarse and fine grids,
respectively. The uncertainties for the grid are evaluated
by Di Mascio et al., (2000) for steady DTMB 5415
simulation by using MGShip. The procedure is basically
same as that for SBD-A, i.e., three computational grids
were used. Being MGShip a multiblock-multigrid
solver, grids are composed by 5 different blocks, each
one allowing 4 grid sublevels with refinement r= 2
in each coordinate direction. Due to the requirements of
the adopted solver, the grids are block-structured, with
hexahedrical elements, and the transom region has been
properly modeled. On the faces common to two blocks
the cell distribution is unique. Computational domain
extends 1.0 LPP forward, 1.0 LPP backward and 1.25 LPP
aside. In the fine grid, 193 cells are placed along the
hull, allowing a correct description of the generated
wave train. Numerical beach is placed at the end of the
grid, in order to damp the solution and increase stability.
In the following table the relevant values of the error
components for the numerical and experimental data are
reported.
D*10-3 S*10-3 E% UE% UV% UD% Ur% USN% USI% USG% USN/UD
0.16260.16140.7381.0093.1950.6880.6883.120 0.0003.120 4.535

Numerical and experimental data are related to


model 5415. It is verified the validation of the numerical
data, since E% is smaller than UE%. Iterative error USI%
has been assumed to be 0, since the solution is
convergent. The error component related to the
numerical uncertainty USN% is 3.12%. Validation
uncertainty UV% is expressed as the sum of the squares
of the error in the data UD% plus the error in the
numerical uncertainty USN%: as a consequence,
validation uncertainty UV% is 3.2%. Since the
comparison error E, say the difference between the data
D and the simulation S, is 0.74%, and it is smaller than
the validation uncertainty UV%, the solution is
validated.

VALIDATION TEST
The experimental campaign used an existing model
of the DTMB 5415, the INSEAN 2340 model already
adopted in previous experiments (Stern et at., 2000).
Tank tests were carried out in the towing tank No. 2 at
INSEAN, and the new models will be tested in the same
basin. The tank No. 2 is 250 meters long, 9 meters wide
and 4.5 meters deep, equipped with a carriage capable
of a maximum speed of 10 m/s, recently renewed
allowing for a precision on the forward velocity of
about 0.1%.
To further reduce the uncertainty connected to the
model geometry, only the bow part of the new hulls has
been build while model 2340 was cut at station 15 and
prepared for the mounting of the two new bows (Fig.
15). The new bow shapes to be mounted on the 2340
body are indicated as 5415-A (SBD-A optimized bow)
and 5415-B (SBD-B optimized bow).

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Scale factor of model of 2340 is 24.824. This value


is compatible with the main dimensions of tank No. 2,
and avoids blockage effects. The model is 5.72 meters
long (between the perpendiculars), the breadth is 0.771
meters and the draft is 0.248 meters.
Displacement has been measured for all the models,
since the draught is assigned. Model displacements are
the following: for model 5415-A is 546.6 kg (8362 tons
in ship scale), for model 5415-B is 542.8 kg (8304 tons
in ship scale) and for the original model 2340 is 549 kg
(in ship scale, 8398 tons). Differences between original
and optimized hull displacements are -0.44% for model
5415-A and -1.13% for model 5415-B. In other words,
both the models satisfy the constraint on the
displacement (max. difference between original and
optimized less than 2%) enforced during the
optimization process.
During the tests a load cell with a reading range of
20 kgf has been used. The same load cell has been
applied for both the models and during the whole
measurement campaign. Towing force has been
registered during the experimental tests, together with
the bow and stern sinkage.
Uncertainty analysis has been performed for two
speeds, that is, Fr = 0.28 (cruise speed and also
optimization speed) and Fr = 0.41 (top speed). Ten
different profiles have been recorded for the interesting
speeds and for both the models. An admissibility test
has been performed on data, based on the variance of
the whole dataset. Bias errors have been assumed from
previous experimental data (Stern et al., 2000).

OPTIMIZED DESIGN AND


VALIDATION
SBD-A
Optimization was performed by OPU using the
scheme SBD-A. The system parameters of GA are as
follows: crossover rate=0.75, mutation rate=0.30, and
population size=60. As described earlier, the population
size coincides with number of available parallel
processors. The crossover and mutation rates are
determined through preliminary numerical tests. The
optimal system parameters may be found by further
investigation, but that is left for future work.
In the present study, the best solution in 50
generations is selected as a final solution (i.e., mode
5415-A), in which the solution was, in fact, obtained at
the 21st generation. For each generation, RANS code is
executed only once per processor, where 5,000 global
sweep iterations appeared to yield sufficient
convergence. CPU time to proceed 10 generations is
about 4 days by using PC-Cluster parallel computing
environment mentioned earlier. Advantage of the
present high performance parallel architecture is
evident, since the CPU time would be 60 times larger if
the conventional serial computation is adopted.
SBD-B
Optimization process is solved by INSEAN adopting
the scheme SBD-B. The variable fidelity approach

described before is used, in order to reduce the CPU


time needed (Peri and Campana, 2004).
The evaluation of the gradient of the objective
function requires 22 calls to the CFD solver. By adding
9 solutions necessary for the line search along the
descent direction, 31 solver calls are performed at each
optimization iteration. Some other CFD calls are needed
by the trust-region. A total of 10 iterations have been
required for the final convergence, giving a total
number of more than 310 function evaluations (with
CFD).
With the variable-fidelity approach, 36 calls to the
high-fidelity solver have been required (the finest grid),
while 314 calls to the low fidelity solver were used (the
coarsest grid). CPU time for the function evaluation
with the low and high fidelity solvers is about 900 s and
5400 s, respectively. As a consequence, the complete
optimization time has been roughly about 477,000 s
(time spent in grid and geometry manipulation is almost
negligible). On the other hand, only 304 objective
function values would have been necessary (after some
saving) by using the high-fidelity model alone, with a
total CPU time of 1,641,600 s. As a consequence, the
final CPU time saving has been 1,164,600 s, about the
71% of the total CPU time.
Optimized bows
The optimization processes carried out by the two
SBD versions ended with two different geometries
produced in fiberglass at INSEAN.
The two geometry (Figs. 8 and 9) shows some
similarities as for the general trend. The most important
is the reduction of the maximum breadth of the dome
which is reduced by about 20% in one case (5415-A)
and about 40% in the other (5415-B).
Volume reduction for the model 5415-A is rather
uniform along the dome axis and the overall shape of
the new dome is similar to the original but for the width
and the length. Perhaps another relevant difference is
the extension of the 5415-A sonar dome in the forward
(x) direction, about 10% ahead with respect to the total
length of the dome itself.
The 5415-B dome is less conventional than the
original and 5415-A models. The overall shape is more
triangular (see Fig. 6) and in some points very close to
the volume constraint posed on the dome. It is also
slightly longer than the original one, but the adopted
grid topology do not allows large variations on the bulb
length. Volume reduction of 5415-B is greater than that
for the 5415-A.
Differences in the region immediately above the
sonar dome are also visible as can be observed from
Fig. 8. The entry angle is reduced and the buttocks (Fig.
9) are more bended than the original ones. Differences
extend up to station 16 for model 5415-B, while for
model 5415-A remain confined between the extreme
bow and station 18.
Fig. 6 clearly shows that in both cases the constraint
on the dome dimension is satisfied and the sonar fits
inside the bulb. A close-up view of model 5415-B
seems to display a slight violation of this constraint near

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

the bottom corner. A careful analysis demonstrated that


the minor violation is located in between two grid
points. As a general comment it must be pointed out that
it is impossible to check geometrical details whose
dimension is below the grid resolution and that
constraints satisfaction can only be checked at grid
points. At all events, the minor violation is extremely
small and of the same order of the grid cells.
All the other geometrical constraints (displacements,
entry angle, etc. ) have satisfied.
Numerical results and verification
Results are summarized in Tab. 2. The optimization
are carried out with the model in fixed condition: SBDB assumes even keel, while SBD-A adopts the sinkage
and trim values measured for the original 5415 at
Fr=0.28.
Both the SBD are able to identify an optimal
solution. SBD-A produces the 5415-A model, with a
decrease of the objective function of about 2.85%
while SBD-B produces the 5415-B model that shows a
reduced value of the objective function of about
6.58%. Values are not comparable because the solvers
are far from convergence due to the grid density.
Influence of the solver on the solution has not been
evaluated because, due to the different requirements in
terms of grid topology, an exchange of the grid has been
not barely applicable.
For the numerical verification of these results, finer
grids (about 1.7Mcells) are used. The verification of
5415-B is initially performed at even keel and in a
second phase with the free model. For 5415-A,
verification is again carried out at the same sinkage and
trim measured for the original 5415 at Fr=0.28.
The two CFD solvers show opposite trends as for the
effect of the grid refinement on the computed Rt.
Indeed, from coarse to fine grid CFDSHIP-Iowa
prediction of Rt increases for both the original
(+18.38%) and the optimized (+15.38%) model. On the
contrary, MGShip predicts a decrease of the total
resistance from coarse to fine grid (original,-15.62%
and optimized, -13.20%).
Before carrying out the experimental campaign on
the two optimized models, the measured value of the
resistance Rt for the original 5415 model (INSEAN
model 2340) was used to estimate the new results (Rt at
Fr=0.28 was 4.6170 Kgf). A positive feature of both the
numerical results is that the total resistance reductions
computed by the two solvers with the fine grid are
higher than the differences between data and
computation for the original 5415. Indeed, for the 5415A, a reduction of 5.32% is predicted (from 4.6277 to
4.3817 Kgf, fine grid) against a difference between
numerical and experimental data (for the original 5415)
of 0.23% (4.6277 versus 4.6170 Kgf). Similarly, for the
5415-B an improvement of 3.90% is obtained (fine grid,
fixed model), while a difference of 2.25% with the
experimental data is observed. Coarse grid CFDSHIPIowa result underestimate the experimental data while
fine grid result overestimate it. The opposite for

MGShip. This means that the numerical solvers


approache the experimental value oscillating around it.
Still about the effects of grid refinement, changing
grid from coarse to fine, 5415-A appears to increase its
total resistance reduction (from 2.85% to 5.32%),
while the 5415-B display an opposite behavior (from
6.58% to 3.90% in even keel).
5415-B has been also computed in free condition.
The concern was about the fact that SBD-B
optimization was carried out at even keel, a condition
that can be far from the real condition at Fr=0.28. Free
condition results show a further decrease of the
reduction, from 3.90% down to 2.54%. (from 4.5133
down to 4.3988 Kgf). However this value is still higher
than the difference between data and computation for
the original 5415 (2.25%).
While 5415-A show substantially the same wetted
surface as of the parent hull, model 5415-B shows a
reduction of the wetted surface of about 1.2%. For the
fixed model, wetted surface variation may uniquely
come from hull shape modifications and from
differences in the wave pattern. Indeed, 5415-B dome is
smaller and the bow is narrower than 5415-A.
Improved resistance reflects on the computed free
surface too, as reported in Fig. 10 and 11. Wave pattern
caused by the optimized hulls is globally smoother, as
can be clearly seen by Fig. 12, that shows the wave
elevation near the hull. Both the optimized models
remarkably reduce the amplitude of the first crest. For
model 5415-B, the steepness of the first wave is also
appreciably smoothed, and also the first hollow is
moderately reduced.
The pressure on the hull, reported in terms of CP in
Fig. 13 shows improvements in the pressure distribution
with reduced low-values regions. Pressure varies more
gently along the hull and pressure gradients are reduced
in the modified region. Model 5415-A seems slightly
more effective in reducing the area of negative CP down
the tip of the dome.
Seakeeping computation have been also performed
on the optimized hulls to verify the functional
constraint. Comparisons of the RAOs in head seas are
reported in Fig. 5. Model 5415-B show slightly
improved seakeeping, larger for the heave than for the
pitch, as a consequence of the change in the bow
volume. Model 5415-A heave is improved while pitch is
slightly worse than the original hull (about 0.5%).
Anyway, both the hulls respect the constraints on the
vertical motions. In the following table, values of
maximum heave, pitch for original and optimized hulls
are reported. Merit function value is 0.989 for model
5415-A and 0.975 for model 5415-B. The constraint on
seakeeping is verified.
HC
PC
SC

5415
1
1
1

5415-A
0.983
0.996
0.989

5415-B
0.981
0.970
0.975

Constraint
1.020
1.020
1.000

Another functional constraint was imposed on the


vorticity shed by the dome. Immediately past the sonar
dome a control region was placed according to the

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

experimental data in which the axial vorticity should not


overcome the original value, as from the definition
given in Tab. 1. Numerical computations are reported in
Fig. 14. and clearly show that the constraint has been
satisfied. The control region, placed in a plane
orthogonal to the forward direction, is reported as a
black circle. The two final models reduce the core of the
main vortex, which appears to be confined near the
body surface.
Experimental validation
The success of the optimization processes is nicely
confirmed by the experimental measurements.
Reductions in total resistance with respect to the
original design are reported in Fig. 16 as a function of
the Froude number, so that values below 0% represents
improved
designs.
Additionally,
during
the
experimental campaign, uncertainty analysis has been
carried out at Fr=0.28 and 0.41 Ten (10) different
measurements have been taken and uncertainty on total
resistance is found to be about 1% at Fr=0.28 and 0.5%
at Fr=0.41. This is partly related with the precision error
of the load cell since it decreases near the maximum
measurable value, and it is also in line with previous
experiences.
At the design speed of Fr=0.28, the measured
reduction of the total resistance in model scale is about
3.80% for both the optimized models and the error bars
for the two experimental campaigns are much lower
than this improvement: this is a valuable results,
considering the small modifications allowed and the
good initial performances of the original model.
In the low Froude number region (lower than
Fr=0.2) performances of model 5415-B are better than
those of 5415-A. In the range of Fr = 0.2 - 0.4 the trend
of the two new models is similar and the reduction in
total resistance decreases from about 6% down to 0% at
the highest speed.

CONCLUSIONS

provide a meaningful design. As a direct consequence,


local optimizers may encounter convergence
difficulties, besides the impossibility of jumping from
one sub-set to another of the feasible set. Furthermore
typical objective functions of hydrodynamic character
display almost ever a multi-modal nature, so that local
optimizer are easily trapped in sub-optimal minima.
The adoption of Global Optimization (GO)
algorithms may greatly reduce the risk of stopping the
search at sub-optimal solutions. However, even if the
use of RANS codes is facilitated by the availability of
high performance computing platforms the cost of one
simulation, i.e. one objective function evaluation, can be
of the order of 1 to 10 CPU hours on current generation
of computers. Under these circumstances the
development of a SBD framework which combines
these highly costly analysis tools and Global
Optimization (GO) algorithms may appear as a paradox,
but design engineers of marine, aeronautical,
automotive transport systems, are very much tempted by
taking this direction. Indeed, the margin for design
improvements is continuously narrowing for the design
engineers are producing near optimal configurations in
many industrial fields, including ships, and probability
that performances breakthrough could come from local
optimization methods is likely small.
These open issues are motivating the development of
GO algorithms with reduced computational effort in
terms of function evaluations.

ACKNOWLEDGMENTS
This work has been partially supported by the U.S.
Office of Naval Research under the grants No.
000140210489, No. N000140210304 and No.
N00140210256, through Dr. Pat Purtell.
The authors would like to thank Angelo Olivieri for
the experimental data. EFC and DP thanks Andrea Di
Mascio for the use of MGShip.

REFERENCES

Two different basic SBD versions have been developed


and tested in a nonlinear constrained optimization
problem. Experimental tests have been carried out on
the two final optimized models. Improvements are
obtained while other important qualities of the ship, like
the seakeeping behavior, is preserved. This result prove
the applicability of RANS solvers inside an
optimization cycle and the reliability of the numerical
results. This application represent a first step into a
cooperative project among IIHR, OPU and INSEAN
which is now at the end of the second year. Multiobjective application is the final task.
Extension of the problem and future research
direction concerns the intrinsic limitation of a local
optimization algorithm. Typical feasible design spaces
of optimization problems, like the one proposed in this
paper, are non-convex and often non-connected due to
the nonlinear geometrical and functional constraints that
have to be enforced to prevent unrealistic results and

Alexandrov, N. M., 1996, A trust region framework


for managing approximation models in engineering
optimization, 6th AIAA / NASA / ISSMO Symp. on
Multidisciplinary Analysis and Opt., AIAA-96-4102
Alexandrov NM, Lewis RM, Gumbert CR, Green LL,
Newman PA, 2000, Optimization with variable-fidelity
models applied to wing design, 38-th Aerospace
Sciences Meeting & Exhibit, Reno, AIAA 2000-0841.
Bischof, C., Carle, A., Corliss, G., Grienwank, A. and
Hovland, P., 1992, ADIFOR: Generating Derivative
Codes
from
Fortran
Programs,
Scientific
Programming, Vol. 1, No. 1, pp. 11-29.
Chang, K. J., Haftka, R.T., Giles, G.L. and Kao, P. J.,
1993, Sensitivity-based scaling for approximating
structural response, J. of Aircraft, 30(2), pp.283288.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Di Mascio, A., Broglia R. and Favini B., 2000, A


second-order Godunov-type scheme for Naval
Hydrodynamics, in Godunov methods: theory and
application, Kluwer Academic publishers, Singapore.

Peri D., Campana E.F., 2003c, High Fidelity Models


in Simulation Based Design, 8th Int. Conf. on
Numerical Ship Hydro., Busan, South Korea

Giering, R., Kaminski, T., 1998, Recipes for Adjoint


Code Construction, ACM Trans. on Math. Software,
Vol. 24, No. 4;437-474.

Peri D., Campana E.F., 2004, High-fidelity Models


and Multiobjective Global Optimization Algorithms in
Simulation Based Design, to appear on J. Ship Res.

Hino, T., Kodama, Y.,Hirata,N.,1998,Hydrodynamic


shape optimization of ship hull forms using CFD, 3rd
Osaka Colloquium on Advanced CFD Applications to
Ship Flow and Hull Form Design, Osaka Prefecture
Univ. and Osaka Univ., Japan

Pironneau, O., 1984, Optimal Shape Design for


Elliptic Systems, Springer-Verlag, New York.

Jameson A., 1988, Aerodynamic design via control


theory, J. of Scientific Comp., 3, 233-260.
Jameson A., Kim S., 2003, Reduction of the Adjoint
Gradient Formula in the Continuous Limit, 41st
Aerospace Sciences Meeting and Exhibit, Reno (NE),
USA, AIAA 20030040
Larsson, L., Stern F. and Bertram V., 2003,
Benchmarking of Computational Fluid Dynamics for
Ship Flows: The Gothenburg 2000 Workshop, J. Ship
Res., Vol. 47, No. 1, pp. 63-81
Lhner R., Soto O., Yang C., 2003,An adjoint based
design methodology for CFD optimization problems,
41st Aerospace Sciences Meeting and Exhibit, Reno,
(USA), AIAA 2003299
Minami, Y., Hinatsu M.,2002, Multi Objective
Optimization of Ship Hull Form Design by Response
Surface Methodology, 24th Symposium on Naval
Hydrodynamics, Fukuoka, JAPAN
Newman III, J.C., Pankajakshan, R., Whitfield, D.L.
and Taylor L.K., 2002, Computational Design
Optimization Using RANS, 24th Symp. on Naval
Hydro., Fukuoka, Japan.
Paterson, E.G., Wilson, R.V., and Stern, F., 2003,
General-Purpose Parallel Unsteady RANS Ship
Hydrodynamics
Code:
CFDSHIP-Iowa,
The
University of Iowa, IIHR Report No.432.
Peri, D., Campana, E.F. and Di Mascio, A., 2001,
Development of CFD-based design optimization
architecture, 1st MIT Conf. on Fluid and Solid Mech.,
Cambridge, MA, USA.
Peri, D., Campana, E.F., 2003a, High fidelity
models in the Multi-disciplinary Optimization of a
frigate ship, 2nd MIT Conf. on Fluid and Solid Mech.,
Cambridge, MA, USA.
Peri, D., Campana, E.F., 2003b, Multidisciplinary
Design Optimization of a Naval Surface Combatant, J.
Ship Res. , Vol. 41, No. 1, pp 1-12

Stern, F., Longo, J., Penna, R. , Olivieri, A., Ratcliffe,


T. and Coleman H.W.,2000, International collaboration
on benchmark CFD validation data for surface
combatant DTMB model 5415, 23rd Symposium on
Naval Hydro., Val de Ruil, France
Stern F., Wilson R.V., Coleman H.W., Paterson E.G.
2001, Comprehensive Approach to Verification and
Validation of CFD SimulationsPart 1: Methodology
and Procedures, J. of Fluids Eng., Vol.123, 793-802
Tahara, Y., Paterson, E., Stern, F. and Himeno, Y.,
2000, Flow- and Wave Field Optimization of Surface
Combatants Using CFD-Based Optimization Methods,
23rd Symposium on Naval Hydro., Val de Ruil, France,
to appear J. Ship Res.
Tahara, Y., Katsui, T., Kawasaki, M., Kodama, K.,
Himeno, Y.,2004,Development of Large-Scale HighPerformance CFD Coding Method for PC-Cluster
Parallel Comp. Environment 1st Report: Setup and
Initial Evaluation of Coding Environment with MPI
Protocol, J. Kansai Society of Naval Arch., No. 241
Thomas, J.P., Hall, K.C. and Dowell, E.H., 2003, A
discrete adjoint approach for modeling unsteady
aerodynamic design sensitivities, 41st Aerospace
Science meeting and Exhibit, Reno (Nevada, USA).
Valorani, M., Peri, D., and Campana E.F., 2003,
Sensitivity analysis methods to design optimal ship
hulls, Optimization and Engineering, 4, 337-364, 2003
Wilson, R., Paterson, E. and Stern, F., 2000,
Verification and Validation for RANS Simulation of a
Naval Combatant, Workshop on Numerical Ship
Hydro. (Gothenburg 2000), Gothenburg, Sweden
Wilson R.V., Stern F., Coleman H.W., Paterson E.G.
2001, Comprehensive Approach to Verification and
Validation of CFD SimulationsPart 2: Application for
RANS Simulation of a Cargo/Container Ship, J. of
Fluids Eng., Vol.123, 803-810
Wilson, R., Carrica, P., Hyman, H., and Stern, F.,
2004, A steady and unsteady single-phase level set
method for large amplitude ship motions and
maneuvering, 25th Symposium on Naval Hydro., St
Johns, Canada

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

SBD B

SBD - A

Phase

5415
Surface Sinkage Trim 5415-A Surface Sinkage Trim
(mm) (deg) Rt(Kg_f) (m2)
(mm) (deg)
Rt(Kg_f) (m2)

Grid
size

Optimization
fixed model
Verification

3.9093

4.8692

8.29

0.05

3.7977

4.8652

8.29

0.05

CFDSHIP-Iowa

4.6277

4.9686

8.29

0.05

4.3817

4.9640

8.29

0.05 1,779,648

5.2164

4.9476

0.0

0.00

4.8730

4.8888

0.0

0.00

4.4016

4.8279

0.0

0.00

4.2298

4.7664

0.0

0.00 1,712,128

fixed model
Optimization
fixed model
Verification
MGShip

fixed model
Verification

238,760

253,952

MGShip
4.5133 4.9418
8.29
0.05 4.3988 4.8966
10.06
0.09 1,712,128
free model
Table 2 - Summary of results for the SBD A and B. All values are in model scale (Lpp=5.72m)

System 2: CAD Emulation

Optimizer

Optimizer
Design Var.

Geometry and grid


manipulation

Objective
Function

CAD

CFD

Constraint
Function

Fig. 1 Basic elements of a CFD-based optimization


environment.

IGES Control

Surface
Grid Generation

CFD

Fig. 4 Implementation of the CAD-Based hull form


modification into optimization environment: System 2
(CAD Emulation).
GA

Start with initial


design xo, trust
region radius

Manipulate
geometry and
volume grid

i = 1, , Ndv

Use Low-Fidelity
RANS to
compute F/xi

GA
Line search with L-F RANS.
Compute step s, new candidate
design xk + s

Fig. 2 Serial and parallel architecture of the Narrow


Band-GA optimizer.

Convergence?

System 1: CAD Direct Control

No

Reduce trust
region radius

too small?
Yes
Compute first
order corrections
and adjust trust
region radius

Constraint
Function
CAD

No

No

Objective
Function

Surface
Grid Generation

Stop

Compute (with L-F) objective


function and constraints at xk + s

Optimizer
Design Var.

Yes

Acceptable?
Yes
New design
xk+1 = xk + s

Compute High-Fidelity
RANS at xk+1

CFD

Fig. 3 Implementation of the CAD-Based hull form


modification into optimization environment: System 1
(CAD Direct control).

Fig. 5 Derivative based optimizer with variable


fidelity (dashed block) / trust region algorithm (dotted
block).

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Fig. 6 The constraint on the sonar dome is verified both the optimized models 5415-A (left) and 5415-B (right). The
two overlapping cylinders in blue represents the sonar array that fit inside the dome. The original shape of the dome is
also reported (solid lines) to put in evidence the common trend of the two solutions to have a smaller dome

Fig. 7 The behavior of the ship in waves was chosen as a functional inequality constraint. Heave (left) and pitch
(right) RAOs in head seas have been selected as performances to be monitored during the optimization process. Both
the optimized models satisfy the requirements

0.11

0.1

Original

0.1

Optimal

0.09

20

0.08
0.07
0.06

21

Optimal

20

0.08

19

18
17

0.07
0.06
0.05

19

18

17

21

0.05

0.04

0.04

0.03

0.03

0.02

0.02

0.01

0.01
0

-0.01

-0.01

-0.02

Original

0.09

-0.05

0.05

-0.02

-0.05

0.05

Fig 8 Comparison of bodyplan between the original and optimal hull forms. Original 5415 vs. 5415-A (left) and
original 5415 vs. 5415-B (right).

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Fig 9 Comparison of bow and sonar dome. Original 5415 (left), 5415-A (center) and 5415-B (right).

0.4

0.1

Optimal
0.05

0.2

-0.2
-0.4

0 .0

02

04
0.0 6
0
0.0

-0.05

Original

0.
0
0.0 06
04
0.
00

0
4
00
-0.

Optimal
Original
-0.
00
0 4

-0.1
0.5
0.75
1
0
0.1
0.2
x
x
Fig 10-SBD-A Comparison of wave contours between the original and optimal hull forms Original 5415 vs 5415-A
(computed with CFDSHIP-Iowa).
0

0.4

0.25

0.1

Optimal
0.05

0.2

-0.2

04
0.0 06
0
0.

0 .0
0.0 06
04

00
0.
2

-0.4

2
00
0.

-0.05

Original

0
00
-0.

Optimal
Original
-0 .
00
0 4

-0.1
0.5
0.75
1
0
0.1
0.2
x
x
Fig 11 -SBD-B Comparison of wave contours between the original and optimal hull forms. Original 5415 vs 5415-B
(computed with MGShip).
0

0.25

Copyright National Academy of Sciences. All rights reserved.

0.015

0.015

Twenty-Fifth Symposium on Naval Hydrodynamics

Original
Optimal

0.1

0.2

-0.005

-0.005

Z
0.005

Z
0.005

Original
Optimal

0.3

0.1

0.2

0.3

Fig- 12 - Comparison of wave profiles near bow between the original and optimal hull forms. Original 5415 vs 5415-A
computed with CFDSHIP-Iowa (left), original 5415 vs 5415-B computed with MGShip (right).
Cp contours
(Optimal)
0

Cp contours
(Original)

0.0

z
-0.025

z
-0.025

0.00

0.00

-0.
2

-0.40

-0.05

-0.20

-0.05

0.20

0.20

0.2

0.20

0.05

0.1

0.00
0

0.00

0.05

Cp contours
(Optimal)
0

Cp contours
(Original)

0.1

0.20

0.20

0
0.0

z
-0.025

z
-0.025

0.00

0.00

0.00
-0.2

-0.05

-0.40

0.05

0.1

-0.40

-0.05

-0.20

0.05

0.1

Fig. 13 Comparison of surface pressure (Cp) contours near the bow between the original and optimal hull forms. Top,
original 5415 (computed by CFDSHIP Iowa) versus 5415-A. Bottom, original 5415 (computed with MGShip) versus
5415-B

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

-0.02

-0.02
Axial Vorticity |x| Contours
(X=0.2: Behind Sonar Dome)

Optimal

Original

Optimal

Original

-0.04

Axial Vorticity |x| Contours


(X=0.2: Behind Sonar Dome)

-0.04
2

10

-0.06

10

-0.06
5

-0.08

-0.08
Control area

-0.025

Control area

0.025

-0.025

0.025

Fig 14 Comparison of axial vorticity contours between the original and optimal hull forms (X=0.2: behind sonar
dome). The control region, placed in a plane orthogonal to the forward direction, is reported as a black circle. Original
5415 vs 5415-A computed by CFDSHIP-Iowa (left), original 5415 vs 5415-B computed with MGShip (right).

Fig. 15 Experimental models. Original 5415 cut (left), 5415-A bow (center), 5415-B bow (right).

2
1
5415-A
5415-B

Percentage difference

-1
-2
-3
-4
-5
-6
-7
-8
-9
-10
-11

0.1

0.2

0.3

Froude number

0.4

0.5

Fig. 16 - Experimental validation of the numerical results. Drag reduction (%) as a function of the Froude number for
the two optimized models. Error bars are plotted for Fr=0.28 and Fr=0.41.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Luigi Martinelli
Princeton University, USA
This is a fine work; however I have a couple
of questions for the authors.
1) There exist well developed and documented
approaches for hydrodynamic optimization that
are more computationally efficient that the one
selected for this study, in that the cost of the
optimization does not depend on the number of
design variables, why were these approaches
discarded?
2) As a side note, the bibliography of some of the
competing approaches is scanty. In particular
that relative to the control theory based approach
of Jameson, which has been extended and
validated for marine applications as early as five
years ago. The authors should have been aware
of that and included in the bibliography.
3) The computational time reported in the paper, 110 CPU hours for a steady RANS calculation on
a 1.7 Million point grid,
revels a fairly
inefficient multigrid solver, at least in
comparison with the performance of other
multigrid solvers, with similar accuracy, that
were developed a decade ago, extensively
validated and used for analysis of ship flow.
What is the most computationally expensive
component of the solver used in the study?

local optima. Again, the framework would


accommodate gradient-based methods with
gradients of any origin.
2) By citing the pioneering papers by Pironneau
(1984) and Jameson (1988) and a couple of
recent applications (Newman et al., 2002 and
Jameson and Kim, 2003) we have had the feeling
that adjoint approaches were sufficiently
represented.
3) The authors believe that there must have been
some misunderstanding, since the computational
time is not reported in the paper. Likely the
comment makes reference to a general remark
made in the conclusion, hereafter reported:the
cost of one simulation, i.e. one objective function
evaluation, can be of the order of 1 to 10 CPU
hours on current generation of computers.
Quite obviously, this was intended to be a
reference to timing typically reported in
literature for free surface RANS solvers, e.g.
Larson et al. (2003).
REFERENCES
[1] Larsson, L., Stern F. and Bertram V., 2003,
Benchmarking of Computational Fluid Dynamics
for Ship Flows: The Gothenburg 2000 Workshop, J.
Ship Res., Vol. 47, No. 1, pp. 63-81

AUTHORS REPLY
We thank Prof. Martinelli for his discussion.
1) As already stated in the paper, adjoint methods
to which the author is referring are not portable
from one flow solver to another. Other
difficulties might come from the addition of
constraints. They require the availability of the
source code, which is not always the case.
Furthermore,
the
basic
variable-fidelity
framework we have adopted can be used with
any derivative-based method and set of models,
including cases where gradients are computed
via adjoints or via automatic differentiation (this
has been reported by other authors). In our case,
we are focusing on zero-order methods, because
we like the solver "portability" this implies,
because the problems in question have a
sufficiently small number of variables, and
because zero-order methods may have additional
smoothing effects that assist in not being stuck in

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Michel Visonneau
Ecole Centrale de Nantes, France
This interesting paper describes a
cooperative study between IIHR, INSEAN and OPU
aiming at comparing two different Simulation Based
Design environments based on RANS solvers for
ship hull optimization and at validating the respective
solutions by a dedicated experimental campaign.
Although the optimization algorithm used in SBD-A
has the possibility to carry out a global optimisation,
the domain of variation of the design variables is
restrained to maintain reasonable CPU times. It
would be interesting to evaluate the benefits from
neural algorithms or hybrid approaches as proposed
in [1] to improve the efficiency of a GA-based
algorithm, which would make possible a search of a
global optimum.
This study was performed at model scale.
Would it be possible, with the actual SBD
environments, to carry out the same computational
studies at full scale? Do you think that the bow
modifications proposed in this study would be
dramatically changed at full scale? Have you also
evaluated the role played by the turbulence modelling
errors?
REFERENCES
[1] R. Duvigneau & M. Visonneau, Hybrid genetic
algorithms and artificial neural networks for complex
design optimization in CFD, Int. J. for Numerical
Methods in Fluids, Vol. 44/11, pp. 1257-1278, 2004.

AUTHORS REPLY
We thank Dr. Visonneau for his discussion.
His first comment refers to the use of meta-models in
the optimization process, i.e. the use of surrogate
analytical models that, after an initial training, are
able to mimic the response of the expensive CFD
solver adopted for the analysis. The authors have
already some experience on that topic [1,2,3] using
many different approaches: Radial Basis Function
neural nets (RBF), kriging, and some others
techniques. In the present paper however, the authors
decided to focus on exploring the advantages of a
parallel cluster (SBD-A) and on those given by the
Variable Fidelity approach integrated with the trustregion method (SBD-B).
In principle there are no objections to extend
the numerical optimization to full scale problem, at
least from an algorithmic standpoint. The major

difficulties may come from the CFD tools, since


validation of the turbulence models at full scale is
still missing. Furthermore, another goal of this study
was the experimental verification of the numerical
results, which had to be performed at model scale.
Since we have never performed numerical
optimization at full scale using RANS solvers, we do
not have feeling of the differences in hull shapes due
to the different Reynolds number. It is quite clear
however that depending on the type of objective
function (local vs. integral quantities) the turbulence
closure might play a master role [4]. That will be a
good topic for future work.
The evaluation of the errors due to the
turbulence model adopted in the RANS solver has
been considered out of the scope of this paper.
However, the RANSE solvers adopted (CFDSHIPIowa and MGShip) have been validated during the
years, including the Gothenburg 2000 workshop on
the DTMB 5415 with encouraging results.
REFERENCES
[1] D Peri, EF Campana, High-fidelity Models and
Multiobjective Global Optimization Algorithms, J. of
Ship Research, in press, 2004
[2] D Peri, EF Campana, High-fidelity models in
Simulation Based Design, 8th International
Conference on Numerical Ship Hydrodynamics,
Busan (S. Korea), 2003.
[3] D Peri, EF Campana, High fidelity models in
global optimization, COCOS'03 2nd Int. Workshop
on Global Constrained Optimization and Constraint
Satisfaction, Lausanne, Switzerland, 2003, to be
published in Lecture Notes in Computer Science,
Springer.
[4] R. Duvigneau, M. Visonneau, G.B. Deng, On the
role played by turbulence closures in hull shape
optimization at model and full scale, J. Marine
Science and Tech., Vol. 8, 1, 2003

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics


St. Johns, Newfoundland and Labrador, CANADA, 8-13 August 2004

Parametric Optimization of SWAT-Hull Forms


by a Viscous-Inviscid Free Surface Method
Driven by a Differential Evolution Algorithm
Stefano Brizzolara
(University of Genova - Italy, EU)

ABSTRACT

INTRODUCTION

An optimization method capable to automatically


search for the optimal underwater hull form of
SWATH type vessels, with regards to total resistance,
is described in the paper. The optimization
environment is based on an original parametric
geometry definition module, capable to define a series
of underwater hull forms characterized by a double
hump and an intermediate hollow shape, and a set of
robust and well proven CFD codes, e.g. a linear free
surface panel method for the prediction of wave
resistance coupled with a thin boundary layer integral
method for the prediction of friction and form
resistance. The validity and good accuracy of the CFD
methods employed in the numerical optimization
procedure is proven in the paper by comparison of the
numerical results obtained in the case of a typical
SWAT demi-hull for which resistance test results were
conducted in towing tank.
The optimization strategy is driven by a differential
evolution algorithm which demonstrated its good
capability in finding the optimum solution, even in a
difficult design space, such as the one adopted for the
application examples presented in the paper. These
application examples regard always the same design
vessel (with fixed main dimensions and displacement),
whose underwater hull forms are optimized with
respect to different objective functions: pure wave
resistance or total resistance, at different given design
speeds. The substantial difference in the shape of the
optimized hull forms, obtained considering different
design objectives, highlights the importance of shaping
the hull form of SWATH ships on the base of a given
operating profile. The order of magnitude of the
reduction of total resistance achievable with an
optimized hull shape should convince the reader about
the importance of using automated CFD optimization
methods in the design of modern and highly efficient
SWATH ships.

Recently a revival of interest about SWATH


(Small Waterplane Twin Hull) ships has been
registered among naval architects, shipyards and
research centres. In fact, aside the usual and never
abandoned military applications, SWATH typologies
are being designed for passenger ships, fast ferries and
pleasure mega-yachts, rather globally. Also in Italy the
department of Naval Architecture and Marine
Engineering of the University of Genova has started an
independent research, which just recently turn out to be
of interest for a pre-competitive research project in
collaboration with an Italian shipyard, aimed to the
design of a foil assisted SWATH vessel. The qualities
which attract interest to a SWATH ship typology are
the potential lowest motions and acceleration in rough
seas in comparison with an equivalent mono hull or
catamaran and the potential low wash characteristics at
high speed. Nowadays both attributes are of primary
importance for the success of a passenger ship: the first
for assuring the best comfort to passengers and the
widest operability of the ship in rough seas; the second
to obtain the acceptance for the ship operation on a
route coming inside a closed bay or passing by sensible
littoral areas.
However, the scientific arguments that motivated
the author to initiate this independent research are
manifold and of different nature:
to integrate an inviscid and viscous CFD method to
be able to predict the total resistance of a SWATH
ship, in a time compatible with the usual design and
to validate it over a typical case
to experiment the possibility to set up an integrated
computational environment for the optimization of
hull forms, as automated as possible
to experiment the use of computer aided parametric
automated definition of (simple) ship hull geometries
to test the applicability of global optimization
algorithms to the design of efficient ship hull forms

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

to find, from the results of the application, a series of


guidelines for the design of underwater SWAT-Hull
forms, which can qualitatively drive the designer in a
new project.
The concept of computer assisted parametric
optimization of hull forms has been introduced in a
relatively recent time to naval architects and has
stimulated an immediate interest and fascination in the
community. This probably because, by tradition naval
architects have always been used to design the hull
forms through the definition of a restricted set of global
and local form parameters, determined on the basis of
available test results of systematic series of hulls. So,
the dream to derive (in a reasonable time) a complete
resistance database of a personalized series of hulls,
naturally captures the attention of the naval architect.
This kind of studies are done already as a routine job in
ship design, but relying almost always on a trial and
error procedure, which still use a lot of human
resources for low level tasks, such as the generation of
the modified ships geometry, the analysis of results
and relative decision on the new geometry adaptation.
The new fascinating idea, used also in this study, is
related to the use of an integrated computer assisted
(possibly fully automated) optimisation environment
based on global optimization algorithms. The idea,
here, is to shift the low added value and time
consuming work (i.e. generate the geometry, prepare
the mesh, run calculation, analyse the results) onto the
computer program, leaving the strategic part of the
work (i.e. choice of free parameters, definition of the
design constraints, definition of the optimization goal)
to the designer.
Some example of applications of these concepts in
the naval field were presented on papers (various,
2003) of different authors all involved in a joint
European research project dedicated to the
development new hull optimization methodologies,
involving different parametric approaches for the
variation of the hull shape and different CFD codes
(mainly panel methods) for the prediction of wave
resistance and/or ship motions. Other examples regards
the approach of Peri et al. (2001), who use a more
direct geometrical modification method that acts
directly on the control points of Bezier surface patches
which represent the hull surface. This methodology
requires the imposition of a set of not intuitive
continuity and fairing constraints to control the
behaviour of the modifying surface as well as other
design constraints, as for instance a fixed displacement,
to be included in the objective function. In this respect,
the smart approach of parametric geometry generation
used by Harries (1998, 1999) appears to overcome
these problems and to be based on a set of parameters
at least more meaningful and familiar to naval
architects.

SWATH types of hull, due to their modular and


geometrically simple constitutive elements, are an ideal
test case for the first application of an automatic
optimization procedure. In fact, already in the past,
several studies were dedicated to this problem.
Papanikolaou et al. (1991), present a very
comprehensive approach to the total design of fast
SWATH vessels, in which for the local form
optimization, he estimated the wave resistance with
Strettenskys integral method (developed for ships
moving in a canal) and integrated the flat plate local
skin friction coefficients along the hull panels and an
empirical form factor to estimate the viscous
resistance. A Lagrange multiplier technique was used
to drive the optimization, keeping as constant the given
displacement and longitudinal centre of buoyancy.
Another very comprehensive example is given by
Salvesen et al. (1985) who first present a well
structured SWATH design environment made up of a
series of modules, which would be serially used by an
operator to design an SWATH hull geometry and
estimate its resistance. The only optimization
procedure used by Salvesen is regarding the wave
resistance component only, estimated through a slender
body theory. A Lagrange multiplier method is used ans
it acts directly on the distribution of the transverse
section areas of the underwater hull forms as requested
in input by the adopted slender body theory.
The method presented in this paper, which extends
and completes the first work on SWAT underwater hull
optimisation, with regards to wave resistance only
(Brizzolara, 2003), uses a combined viscous-inviscid
method for the evaluation of total resistance of
SWATHs, and in this respect should be more general
and should give more accurate trends than the
previously cited works. Moreover, the automatic procedure based on the parametric geometry generation
driven by the optimization algorithm, and the use of the
coupled CFD methods, made possible to investigate a
huge number of cases (of the order of 1000) for each
optimization exercise, in a reasonable time, and to
easily single out (hopefully) the best hull form.
The marked difference in the shape of the
optimized hull forms obtained using the total resistance
as a criteria for optimization instead of just the wave
resistance, at the different considered speeds confirms
the importance of the significance of using a viscousinviscid CFD like the one used in our study.
NUMERICAL OPTIMIZATION ENVIRONMENT
A computer program has been developed
according to the flow chart presented in figure 1. We
define it optimization environment since the program
is complemented by a series of analysis and graphic
routines, which help the user to easily check the status

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

of the convergence, as well as to discuss the optimum


solution (against some other good candidates) and the
history of the free parameters and the objective
function, in the end. The core of the procedure is made
up by the parametric geometry generation module, the
CFD solvers and the optimization algorithm, but also
the routines that integrate various modules by
elaborating creating a valid input data for one from the
output of the other were required and time consuming.
With reference to the independent and automatic
character which the whole procedure must have, a very
important detail is also the recurrent check of error,
which must be done at all levels to prevent a break in
the optimization flow.
INPUT DATA

INPUT DATA

Design and Reference Data:


CP, L, Dmax. Ls, Bs
Fni, CWrif, CTrif, CVrif
Initial Range of Free Var.:
Min&Max{r0, r1, m1, m2, m3}
D.E. algorithm parameters:
CR, F, NP, VTR, iter_max

Free Variables Value:


r0, r1, m1, m2, m3

3D Panel Mesh:
Body with elliptic or circular sections

Interaction:

Transp. Vel. &


separated region

CFD

Input
file

Thin B.L. Integral Method:


Cf and Separation areas: Cvp

Calculate Objective Function:


CT, DCW for one or more weighted speed
YES:

VTR ?
ITERmax ?
NO:

STOP

MODULE

Convergence Criteria & Output Data:


Cw, Cf, Cvp, CT

OPTIMIZATION

Historic
Data

Free Surface B.E. Method:


Wave resistance,
wave pattern and streamlines

MODULES

Input file generation for


Free Surface B.E. method

Historic
Data

Historic
Data

MODULE

GEOMETRY

Determine the Generatrix Curve:


solve the linear system of eq.s

Generate New Parameter Vector


According to D.E. algorithm

Figure 1: Flow chart of the optimization program, with main


details of the major component modules.

The optimization environment has been created in


SCILAB, a platform-independent public domain scientific computation system in continuous evolution. This
environment offers a very complete and powerful set of
scientific functions for calculus, data manipulation and
graphics. A SCILAB macro can easily manage Fortran

and C subroutines together with system commands and


compiled executables. So apart from the CFD codes
written in Fortran, all the rest of the optimization
procedure is written in SCILAB, including the
optimization algorithm and the geometry generation
module. The structure in figure 1, reflects the hierarchy
of the highest level routines. After the definition of the
main input, the top level program, that drives the rest
of the modules, is the optimization algorithm, which
according to its strategy generates ceach iteration a
different set of free variables, which identify the
geometry of a SWAT-Hull, and successively starts the
serial launch of CFD modules and post processor
routines. The control of the various module is reserved
to the top level routine written in SCILAB, which
manages unwelcome crashes of the different modules.
The theoretical and numerical topics of the different
modules are briefly described in the next paragraphs.
PARAMETRIC GEOMETRY DEFINITION
Parametric representation of the hull surface is
essential for the automated optimization procedure.
SWAT-Hull forms are easy to be represented
parametrically. In our case we can assume: the struts
having parabolic or circular arc sections, defined by a
maximum thickness and mean length, a taper ratio and
an inward or outward cant; the underwater hulls either
as body of revolution, defined by one generatrix
profile, or as body with elliptic cross sections, having
the two symmetry axes defined by two profiles: one in
the longitudinal and the other in the horizontal
symmetry planes.
Following a previous work (Brizzolara, 2003), the
underwater hulls were kept as bodies of revolution with
a particular parameterization of the generatrix curve.
The intention is to generate unconventional underwater
hull forms that can maximize the wave cancellation
effects between the generated wave trains along the
length. For this reason a profile typology with at least
three relative extremes (maxima or minima), at given
longitudinal positions (Xm1, Xm2, Xm3), has been
defined. Its length, L, maximum diameter, D, and
therefore displacement or CP, are also fixed. The
curvature radii, RL.E. and RT.E., at the leading and
trailing edge of the curve are given and the thickness of
the profile in correspondence of the second and third
relative extremes.
Figure 2 represents these parameters on a nondimensional coordinate system assumed for the rest of
the paper: the transversal coordinates are nondimensionalized with respect to the maximum diameter
D, while the longitudinal coordinates with respect to
the maximum length L.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

In the affine non-dimensional plane {y,x}={Y/D,X/L},


the generatrix curve has been analytically defined
through the following polynomial of 8th degree:
8

y 2 (x ) = ak x k

(1)

k =1

The polynomial representation was inspired by


that used by Gertler (1950) for DTMB Series 58 of
underwater streamlined body of revolution, as well as
the way he used for determine the coefficients a k . In
fact, these coefficient are found by imposing the
following physically meaningful constraints:
- the non-dimensional length of the curve l = 1 ;
- the non-dimensional first maximum y1 = Y1 / D = 0.5
- the non-dimensional longitudinal position of the three
relative extremes: m1 , m 2 , m3
- the total displaced volume, imposed through the
longitudinal prismatic coefficient: C P = 4 LD 2
- the non-dimensional value of the curvature radii at
the leading and trailing edge of the curve,
respectively: r0 = RL.E . L D 2 ; r1 = RT .E . L D 2
The above conditions expressed in mathematical form,
result in the following system of equations:

y (m1 ) = 1 2
1
(
)
y
'
m
0
=

1
y 2 (x )dx = C 4
y ' (m ) = 0
P

2
& 0

y ' (m3 ) = 0
1 + y ' 2 (0) 3 2 y ' ' (0 ) = r
0
y (1) = 01

2
2
y ' ' (1) = r1
1 + y ' (1)

[
[

(2)

]
]

Substituting the polynomial representation of the curve


(1) into the system (2), one obtains the following
system of linear equations, with respect to the unknown
coefficients a k of the polynomial function (1):

8
k 1
8
k
k ak m1 = 0
a k m1 = 1 2

k =1

k =1
8
k 1
8
k ak m2 = 0 & a (k + 1) = C 4
k
P
k =1
k =1
8
a = r
k a m k 1 = 0
1
k
1 0

k =1
8
8
k a k = r1
a =0
k =1
k

k =1

(3)

In the optimization examples presented next in the


paper, the length, maximum diameter and volume of
the hull were fixed, so each iteration a reduced set of
free variables, FV={m1,m2,m3,r0,r1} is generated by
the differential evolution algorithm, the corresponding

generatrix curve is found by solution of the system (3)


and finally the hull panel mesh is generated. The
surface of the underwater body is intersected, without
any fillet radius, with the surface of the strut, that for
this study has been assumed vertically cylindrical with
parabolic section of the form:

Y (x ) =

t
1 tS
2 S3 X 2
2 LS
LS

(4)

in which tS is the maximum thickness of the strut and


LS is its length, and the reference point for local
coordinate are on the mid point of the strut length.

Figure 2: Type of generatrix curve of the underwater body of


revolution and relative parameters used to define it.

It is obvious that many other mathematical


representation of the generatrix curve are possible. For
instance, a convenient formulation would be a B-Spline
curve, defined by 7 vertexes, two fixed at the L.E. and
T.E., other two placed right vertically over the L.E. and
T.E. vertexes, to control the curvature radii, and the last
three free to move in any direction (within a certain
maximum range). This representation, though having
the advantage to depend on parameters defined over a
certainly continuous domain, looses the implicit
satisfaction of the required fixed volume, so it requires
an additional iterative optimization procedure to satisfy
this design constraint. On the contrary, the five free
variables FV employed to define the polynomial
generatrix curve of (1), result defined over a noncontinuous domain, which is made continuous, for the
sake of good convergence, returning an artificially high
object function value, where the system (3) has not
solution. This fact complicates the convergence toward
the minimum of the objective function. In fact, for an
objective function defined in such a complex domain
the usual analytical (deterministic) minimization
algorithms, would fail to find the absolute minimum.
Stochastic optimization algorithms, like genetic types
or differential evolution types, instead, have more
chance to explore the whole unconnected domain,
without getting stuck from a barrier of non-existent
solutions.
Examples of generated profiles and corresponding
demi-hull panel meshes are given in figure 3, which
compares the reference demi-hull form (with

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

underwater hull taken from model #40050165 of Series


58) with the four optimized hulls obtained at four
reference speeds. The corresponding generatrix curves
of the underwater body of revolution are represented in
figure 14. The body panel mesh is automatically
obtained from a net of cubic spline curves, defined
over several patches, which can have different panel
densities and longitudinal distributions to better define
the strut intersection with the hull body and especially
at the leading and trailing edges of both bodies.

S58

validated in the case of high speed mono- and multihull vessels (Brizzolara et al., 1998) also with the
possibility of including dynamic attitude prediction and
flow behind dry transom sterns (Brizzolara &
Bruzzone, 2000).
A Cartesian right-handed reference frame {XYZ}
travelling with the ship at U constant speed is centred
at an arbitrary point on the intersection of the
longitudinal symmetry plane with the undisturbed free
surface; the x axis oriented aft-wards, and the z axis
oriented upwards.
The total velocity potential = U x + , and the
unknown perturbation potential , with respect to the
uniform incident flow, must both satisfy the Laplace
equation in the complete domain:

= 0 , = 0

(5)

together with the following boundary conditions:


O30

r
n = 0

on the hulls

(6)

= 0

on the free surface

(7)

1
1
= U 2 on the free surface
2
2
U x , 0 for x

(8)

g +

O35

O41

O50

Figure 3: Perspective view of the panel mesh automatically


generated by the parametric geometry module. From top to
bottom: reference SWATH demi-hull form, with underwater
hull of Series 58, and the four demi-hulls optimized at
Fn=0.30, 0.35, 0.41, 0.40.

VISCOUS-INVISCID FREE SURFACE METHOD


An incompressible irrotational potential flow is
assumed by enforcement of the Laplace equation to the
total velocity potential in the fluid domain bounded by
the hull surface SB and the free surface SF. We use a
indirect boundary element method, linearized with
respect to the double model flow as developed by
Bruzzone (1994) and further adapted and successfully

(9)

namely, the Neumann condition on hulls surfaces (6),


the kinematic and dynamic condition on the free
surface (8,9) and the radiation condition for the
disturbance upstream (10). =(x,y) represents the
unknown free surface equation.
In our method the free surface boundary
conditions (7) and (8) are linearized using a small
perturbation theory, by which the total velocity
potential is considered as the sum of a main
contribution represented by the potential D of the
flow around a double model symmetrical with respect
to the undisturbed free surface, considered as deeply
immersed in the fluid, and the contribution of the
perturbation potential 0 due to presence of the wavy
free surface. Congruently the free surface is thought as
composed by the Bernoulli wave D calculated for the
double model, and a smaller order component o , i.e.:

= D + o

D =

U
D D

2g
2g
2

(10)
on z=0

(11)

Using this assumption, it is possible to combine


the free surface boundary conditions (7) and (8) in the
following analytical linear expression valid,
congruently with the linearization, on z=0:

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

2a x + 2b y + xx 2Dx + yy 2Dy +
+ 2 Dx Dy xy + g z = 2a ( Dx U ) + 2b Dy

a = Dx Dxx + Dy Dxy

where:

(12)

(13)

b = Dx Dxy + Dy Dyy

The double model and the linear free surface


potential flow problems are both solved by a boundary
element method which discretized the continuous
problem, approximating the hull surface (SH) and the
undisturbed free surface (SF), with a structured set of
quadrilateral planar panels each having constant
distribution of Rankine sources on it.
Defining the influence coefficient vector as the
velocity vector induced at the centroid of panel i by a
panel j having a uniform distribution of sources with
constant strength j:

quad j

1
)dS j ( X ij , Yij , Z ij )
rij

(14)

we discretized the boundary conditions (6) and (8),


imposing them on each panel centroid taking into
account for the contribution of any panel on the hull
(NH in number) and on the free surface (NF in
number). As a result, the following linear system of
equations in the unknown sources intensities is
obtained:
NH +NF

X
j =1

ij

n xi + Yij n yi + Z ij n zi = Ux

i = 1, N H

(15)

preferable, in general, in case of fast ship hulls with a


relevant dynamic trim and sinkage. The SWATH hull
attitude was kept fixed in the computation of this study.
The viscous effects are approximated by a bidimensional thin boundary layer integral method
applied on the three dimensional inviscid streamlines
distributed to cover the largest portion of the body
surface. This approach is widely used in several
viscous-inviscid panel methods developed in the
aeronautical field, for instance by Maskew (1968). The
methods have been validated with success on many
different airships fuselage forms and aircraft (Dvorak
et. al., 1977) at Reynolds numbers typical of those
reached during ships hulls towing tank tests. Usually
they give sufficiently accurate results when the cross
flow is negligible and when separation is absent or it is
confined in a limited portion of the body. The classical
bi-dimensional integral boundary layer equation is
solved for the laminar flow using Thwaites (1949)
method as revised by Curle (1967):
d
dU 1
+ (2 + H )
= Cf
d
U d 2

(17)

where is the b.l. momentum thickness at a point on


the curvilinear abscissa of the considered inviscid
streamline which has on any point a potential flow
velocity U=U{x(),y(),z()} interpolated on it from
the known solution of the inviscid free surface problem
(16), or from that of the double model problem.
C f = U 2 (u ) =0 is the local friction coefficient

and H= / is the shape factor; u() describes the


velocity profile across the b.l., being the local
coordinate perpendicular to the body surface.
Defining a parameter K proportional to the
gradient of the external velocity U, found from the
solution of the potential flow over the body, and the
other two correlated parameters l and L:
*

X ij

[2a i X ij + 2bi Yij + ( D ) i2


+

x
x
j =1

NH +NF

(16)
D 2 Yij

Yij
)i
+ 2( D )( D )
] j +
y
y
x
x x
D
D
+ 2g i = 2a i (
U ) + 2bi
i = 1, NF
x
y

+(

To compute the derivatives of the potential, a four


points differential operator is used, everywhere on the
free surface in both longitudinal and transversal
directions, except behind transom sterns. As known
this operator gives an implicit property of numerical
damping of the disturbance which otherwise should be
enforced by other means. The free surface waves are
found by substituting the total velocities calculated
over the free surface panels in (8).
The wave resistance is found, in this study, by the
classical integration of the dynamic pressure found at
the each panel centre, with Bernoulli equation over all
the hull panels. Other studies (Brizzolara et al., 1998)
used a (numerical) transverse cut method, to calculate
the wave resistance from the energy content of the
generated wave pattern. No attempt were made in this
study to compare the two methods, being the second

K=

2 U ,
u
l=

U

, L = 2[l K (H + 2)] (18)


=0

the b.l. equation (17) can be transformed in a new


differential equation:


U L
K
=

U

(19)

which can be solved using the empirical closing


relations between L and K found by Curle (1967):
L(K , ) = 0.45 6 K + g (K , )

(20)

where obeys to Curles differential equation:


=

4U d 2U
d 2

Copyright National Academy of Sciences. All rights reserved.

(21)

Twenty-Fifth Symposium on Naval Hydrodynamics

Equation (19) is integrated over the streamline length,


obtaining the solution for the b.l. momentum thickness:
2 ( ) =

0.45
[1 + 2.22 g (K , )]U 5 ( )d + 2 (0) U (0) (22)
U 2 ( ) 0
U ( )

which is integrated assuming an initial value of the


momentum thickness at the stagnation point =0 on the
streamline (see Maskew ,1981), with an iterative
procedure, since g(K,) is depending in turn by . The
first iteration g=0 may be assumed. Once (22) is solved
the other integral b.l. parameters are found by means of
(18) and using the Karman-Polhausen relations to
derive the b.l. thickness .
Transition is predicted by the Granville method,
which uses empirical analytical relations between the
local momentum thickness Reynolds number and the
local external velocity gradient parameter K, previously
defined. If laminar separation is predicted, an empirical
relation, based again on Rn, is used to check if the b.l.
will reattach as turbulent b.l. or will separate for the
rest of the streamline. After transition or reattachment,
the turbulent b.l. computation begins, assuming the
same momentum thickness found at the last calculated
laminar b.l. point and using an empirical relation for
the initial shape factor.
The turbulent form of the integral equation (17) is
solved using Nash and Hicks (1981) method, which is
based on the solution of three integro-differential
equations, obtained by integrating the continuity
equation along the boundary layer thickness, on every
point along the streamline. Assuming a Coles velocity
profile across the boundary layer:
u ( , ) =

u
u u

1 cos
+ C +
ln

K

2

(23)
VALIDATION OF THE CFD METHOD

K=0.41 and C=2.05, u is the friction velocity and u , a


parameter with dimension of a velocity, a set of three
ordinary differential equation is obtained of the form:

1 d
1 du
1 du
+ , 2
+ ,3
=
d
d
d
1 dU
= ,3
+
d

,1

(24)

with the coefficients ,i and (=1..3, i=1..4) that


are depend only on the following unknown parameters:
, u , u , and C . This last parameter is found
introducing another equation:

dC 0.15
=
C C
d

1
where C =
d

2 U 2 0

(25)
1

C = 0.0251
H

A standard first order finite difference scheme is used


to integrate the three eq. (24) with (25), by refreshing
the values of the coefficients ,i and along the
streamline, at each iteration. The relations valid for the
Coles profiles are used at the end to calculate the
boundary layer momentum and displacement thickness
as a function of the standard thickness, found as part of
the solution.
At this level the two methods are simply related.
The real iterative coupling scheme between the two
flows may be performed (see Cebeci, 1998), for
instance, using the transpiration velocity concept of
Lighthill. Virtual normal velocity components on the
body (named transpiration velocities) are needed in the
potential flow problem to impose that the calculated
edge of the boundary layer is a streamline also of the
potential flow problem. The transpiration velocities,
depend, as from Lighthill, from the boundary layer
growth on the body. So, interpolating the transpiration
velocities found on each streamline back at each panel
centroid, it is possible to solve the potential flow
problem, with the BEM, assigning a finite normal
velocity instead of the usual Neumann condition in (7).
The new solution of the inviscid method, with
transpiration velocities, is in turn used to solve a new
b.l. problem and the procedure is iterated until the
convergence is found. For the optimization procedure
though, no iterative coupling was used between the tow
methods. Only a rough estimation of viscous pressure
resistance is made, by excluding the contribution of
those panels which belong to a region of separated
flow, from the integration of the dynamic pressure
which gives the total (dynamic) pressure resistance.

The viscous-inviscid method outlined in previous


section, has been applied in the case of a SWATH like
monohull which was tested in towing tank for the
validation study of CFD methods to be used in an
ongoing research project on similar hulls. The test case
is a mono-hull of the SWATH type having elliptic
cross sections and a symmetric strut having a circular
arc section. The horizontal and longitudinal profiles of
the hull are represented in figure 4, while the main
characteristics of the model are reported in table 1.
Table 1: Main Characteristics of the SWATH demi-hull
model tested in towing tank for validation of CFD methods
LOS [m]

Bmax [m]

3.625

0.463

1.72

0.12

T [m]
Hmax
[m]
Underwater Hull
0.350
0.475
Strut
0.125
-

Copyright National Academy of Sciences. All rights reserved.

S [m2]

[m3]

3.06

0.258

0.542

0.021

Twenty-Fifth Symposium on Naval Hydrodynamics

Anyhow, the thin boundary layer method used did


not predict any flow separation up to the body
truncated end, so frictional resistance were the only
component of viscous resistance. The numerical
frictional resistance coefficient is compared in figure 6
with the reference value obtained for the whole body
using the correlation curves of the turbulent flat plate
of Schoenherr and of the ITTC57.
These reference curves are named composed
since the total friction resistance coefficient is obtained
by summation of the two partial coefficients of the strut
and of the hull, each at its characteristic (length)
Reynolds number, and weighted by the correspondding wetted surface, i.e.:

C FTOT = C F Rn LSTRUT

)S

STRUT

TOT

+ C F Rn LHULL

)S

HULL

S TOT

(26)

4.5

1.2

Numerical

4.3

Figure 4: Longitudinal and horizontal profile of the SWATH


demi-hull model tested in towing tank for resistance
measurements.

1.1

ITTC'57 composed
4.1

1.0

Schoennher composed
3.9

0.9

Rn-Fn

0.8

3.5

0.7

3.3

0.6

3.1

0.5

2.9

0.4

2.7

0.3

Bare hull resistance tests were conducted without


the use of any turbulence stimulator, so also in the
numerical calculations the natural transition criteria of
Granville was used.
Figure 5 presents the comparison between the
numerically predicted and experimentally measured
total resistance. Evidently, the agreement is excellent in
the whole speed range, also near the peak due to wave
resistance; poorer, on the contrary, for Fn<0.28, where
probably the interactions between viscous-inviscid
flows are more pronounced and highly non linear
(large separated regions also in the laminar flow).
Probably in this regime, direct viscous-inviscid
interaction method, with a proper description of the
separated flow regions and those with laminar bubbles,
would lead to better correlations.
14

12

CT*103

10

Numerical
Experimental
6

2
0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.0

Fn Los

Figure 5: Comparison of predicted and measured total


resistrance for the test SWATH demihull

CF*10

3.7

2.5

FnLos

Free surface inviscid calculations were performed


in the complete range of Fn tested and the thin
boundary layer, based on the inviscid pressure
distribution, was calculated for the corresponding
Reynolds number in model scale. A number of about
700 panels was used with about 30 streamlines to
describe the (half of the) body. The free surface was
discretized with about 3000-4000 panels depending on
the Froude number, for an extension of about 3 hull
length by one hull length aside.

0.2
2

10

RnLos *10

12

14

16

18

-6

Figure 6: Comparison of numerical total frictional resistance


predicted for the test case at model scale Reynolds numbers,
with total frictional resistance calculated on the basis of
classical correlation curves.

For highest Reynolds number the numerical curve is


practically identical to the turbulent flat plate curve of
Schoenherr, while for Rn < 12 10 6 , it results even
lower, in spite of any form factor. In fact at these low
Rn, a considerable portion of the hull and of the strut
are interested by laminar flow, according to the
transition criteria used. This large laminar portion of
flow, clearly visible from figure 8 which presents the
plot of the local friction coefficient at a typical model
scale Rn, is also due to the very fine entrance body of
the underwater hull. Laminar flow region on the
streamlines extend up to the magenta color.
When numerical calculations for the full scale
Reynolds number, Rn [100 : 1200] 10 6 , are compared
(figure 7), then a certain form factor re-appears in the
numerical calculations with respect to considered
friction lines. In this respect the results obtained for full
scale were judge realistic, keeping in mind that the
scope of the optimization, for our optimization scope,
has a comparative more than absolute meaning.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

R is the resistance (or a component of it) evaluated for


the hull defined by the currently generated FVj vector;
the optimization procedure is naturally an iterative
procedure in which a number of successive tentative
solutions are evaluated and the convergence is
controlled by a deterministic or stochastic optimization
strategy.
The constraints used for the optimisation were,
first of all, the range limits of the free parameters. In
our case, allowing the larger variation:

2.4

Numerical
2.2

ITTC'57 composed
Schoennher composed

CF*10

2.0

1.8

1.6

1.4

0 < m1 , m2 , m3 < 1

1.2
0

200

400

600

800

RnLos *10

1000

1200

1400

-6

Figurre 7: Same comparison as in figure 6, but at typical full


scale Reynolds numbers.

Figure 8: Plot of streamlines coloured in relation to the local


friction coefficient (colour scale on the right); one of the
lowest model scale numbers RnLos=4.84106.

OPTIMIZATION PROBLEM DEFINITION

The problem of optimizing the shape of the underwater


hull in our case is easily identified and consists in a
minimization problem of the numerical resistance
experienced by the a parametric hull shape which is let
free to change in a certain design space bounded and
limited by a number of constraints. In general, also
other less obvious optimization problems can anyhow
be reduced to a minimization problem of a certain
function named objective function, which (also
indirectly) depends from the values of the free
parameters of the problem. In our case, in the most
general sense, the objective function can be assumed as
the weighted average of the resistance deviation with
respect to that of a reference solution, calculated at one
or more speeds NV, i.e.:

Depending on the method adopted, these


constraints may explicitly be part of the formulation of
the numerical algorithm or can be implicitly included
inside the evaluation of the objective function, leaving
the parameter vector unbounded.
Often the objective function is non-linear, with
respect to the free variables and may not behave well in
some regions of the design space, i.e. can be singular or
non definite. Our case has all these characteristics,
since for instance there will be some combination of
the free parameter vector FV={m1,m2,m3,r0,r1}, which
lead to non real solution of system (3). There are other
solutions which lead to unfeasible curve shapes, in
which the underwater hull surface is slimmer than the
strut local breadth at some intersection point; or in
which the generatrix curve of the underwater hull has
more than three relative extremes over its length due to
the eight degree polynomial form (1) used for its
definition.
When any of the above mentioned situation
happens, the optimization procedure must not stop and
the optimization algorithm must always have in return
a real value of objective function. So one keeps these
unfeasible individuals alive, but practically discard
them from being possible solutions adding a high
penalty figure to the objective function value. In our
case in addition to the F.V. bounds of (11), we checked
these conditions:

L. E .
T .E.
x : y HULL (x ) > y STRUT ( x ), x x STRUT
; x STRUT

x i , i = 1, ni & ni > 3 :

dy HULL ( x i )
= 0, xi
dx

a k , k = 1,8 : m(a k ) 0

(29)
(30)
(31)

and we assigned them the following penalty values:

NV

[r0 , r1 , m1 ,m 2 , m3 ] j = p i R% (Vi )
R% (Vi ) = 10 2

(28)

0 < r0 , r1 <

i =1

R (Vi ) RRe f . Hull (Vi )


RRe f . Hull (Vi )

(27)
NV

p
i =1

=1

Condition
Penalty
Function

F.V.out of range (28)


10 (m j ), m j > 1

(29)

(30)

(31)

100

500

1000

102 (1 m j ), m j < 0

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

OPTIMIZATION ALGORITHM

From all the above reasons, adding that the possibility


that the CFD solvers, though adapted, may crash or
give unrealistic results, it is clear that the minimization
problem is not as easy task to be solved. The highly
non-linearity and bad behaviour of the objective
functions, which are analytically non-differentiable (or
numerically differentiable at a great cost in terms of
calculation time) and characterized by many possible
local extremes and peaks (a penalty value for an
unfeasible solution), convinced us to opt for a nondeterministic stochastic type algorithm. Among them,
the Differential Evolution is a relatively recent and
efficient direct search method first developed by Storn
& Price (1995), which is having good resonance in the
scientific world, being among all inherently parallel.
This method uses a rather simple population based,
stochastic approach and its global convergence
properties have been proven over several test cases
with real valued multi-modal non-linear and non
differentiable objective functions. Its efficiency
secured it the third ranking in the First International
Contest on Evolutionary Computation (1stICEO), after
two non-genetic algorithms, which are not universally
applicable, though. A short description of the main
principles of the adopted scheme is given below but
more details can be found in Price (1996) and Storn
(1996) for the practical applications.
The D.E. algorithm create successive generations
each one having NP population individuals, each one
characterized by a F.V. vector. The generation
individuals do not change in number during the
evolution minimization process. The first generation
has individuals uniformly (in probabilistic sense)
distributed within a certain guessed initial domain, or
around a known solution. Up to this point, like genetic
algorithms. The idea behind D.E. is a new scheme to
generate the trial free variables vectors: D.E. generates
new F.V. vectors by adding the weighted difference
vector between two population members to a third
member. If the resulting vector scores a lower objective
function value, then the new vector replaces the
individual with which it was compared. The
comparison vector can, but not need to be, a member of
the population. To keep track of the progress of the
minimization procedure, the best individual is saved
for each generation. Several variation of the algorithm
exist, depending on the way of calculating the random
deviations on the base of the vector difference of two
generation members.
The scheme DE1 uses the following relation:
r
r
r r
(32)
xnew = x1 + F ( x2 x3 )

in which xr1 , xr 2 , xr3 are three vectors taken randomly


among the individuals of the same generation. F is a
constant factor which controls the amplification of the
differential variation. In order to include a certain
mutation among individuals of the same species, as in
genetic algorithms, a direct cross-over (direct substitution of each vector component) is performed on the
newly generated vectors. The factors which control the
number of substituted vectors components is the crossover factor CR [0;1] , which is related to the
probability of that component to be substituted. The
new mutated vector will be part of the next generation
if its objective function will be lower than that obtained
with the original base vector.
The scheme DE2, differs from the DE1 basically
in the generation of the new trial vectors:
r
r
r
r
r
r
(33)
x new = x1 + ( xbest x1 ) + F ( x 2 x3 )
The additional control variable regulates the
greediness of the scheme to converge towards a local
minimum by adding a vector deviation which depends
also by the distance to the best member of the current
r
generation x best . Other variations may be introduced by
modifying the crossover function (exponential or
r
binomial) and the base vector x1 in for the new vector
differential generation formula (15) that instead of
being randomly selected it may be fixed to the best
vector of the current generation.
In any case no attempt of personalizing the
algorithm has been done for this study, and the
standard DE/rand/1/exp algorithm has been used with
the following constant factors: F=0.6, CR=0.8, NP=50.
Usually the minimum value is reached after 20-30
generations, after that no more improvement is found.
Being the minimum value not known, the value to
reach has been set to an extreme value of -100% so the
procedure always stops when the maximum number of
iteration is reached. In such a difficult and irregular
free variables design space, it is difficult to ascertain if
the absolute best vector is really the optimum. A weak
confirmation may come from the comparison of the
shape determined for the last three or four best cases: if
these forms are similar, and were found in the last
generations, hopefully they represent the optimum hull
forms. This is comparison is shown, in the examples of
figure 14.
OPTIMIZATION EXAMPLES

In Brizzolara (2003) a first work on the optimization of


the underwater hull form of SWATH was presented
and already showed promising potential possibilities
for an eventual automated intelligent procedure. In
the previous study, in fact, the same parametric hull

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

forms, with the same given displacement, were used


and a systematical evaluation of wave resistance was
performed, exploring a certain range of variation of the
five free variables FV, without any optimization
strategy. The best hull shape was found, a posteriori,
by ordering the solutions with respect to the calculated
wave resistance. The calculation were done for two
different design speeds, Fn=0.30 and Fn=0.50, and
already some interesting and marked trend of the shape
of the best hull were found.
The first set of new optimization activities is
intended to complete and extend those first calculations, introducing, as a novelty, the optimization
algorithm and a couple of new design speeds, to better
resolve the effect of forward ships velocity.
The object function (27) has been calculated
assuming, as reference, the calculated resistance for a
conventional SWATH design, which features two
underwater bodies of revolution in similitude from
model #40050165 of Series 58 (Gertler, 1950) and a
parabolic strut, whose form was kept constant during
the parametric optimization. Main characteristics of the
reference hull are summarized in table 2, while the 3D
representation of panel meshes are illustrated in figure
3, together with those of the optimum hulls found.

length with respect to the reference hull), as well as the


best position of the first maximum diameter (m1) is
more advanced or the best of LCB at this speed is
towards midship and the T.E. curvature radius is higher
than the reference hull.
In addition, from the analysis of the best generatrix
curves, it emerges that when the speed is increased, the
through position advance from the most rear
corresponding to Fn=0.30 to a more central one at the
highest speeds (Fn=0.41,0.40). This shift can be
correlated to wave interference effect. In table 3 which
summarizes the main characteristics of the inviscid
optimized hull forms, it has been reported also the
distance between two peaks (m3-m1) of the optimized
hull, and that between the two peaks and the
intermediate through (m3-m1 or m2-m3), which seems to
be correlated to the half or full (/L) fundamental wave
length. At Fn=0.35 and 0.41, for instance, the hull peak
to through distance is close to half the fundamental
wave number. In fact the scatter plots of figure 9, it
may be noted that the best four solutions are laying in
the area of the graph where the non-dimensional
coordinate of the first maxima is close to the leading
edge (m10.1) and second relative maxima is close to
trailing edge (m20.8).

Table 2: Main geometric characteristics of the reference


SWATH design with underwater hull from Series 58.

Table 3: Main characteristics of the optimized hull forms


with inviscid method

Underwater Hull
LH [m]
80.0
DH/LH
0.075
CP
0.65
S/2/3
L/1/3

Strut
LS/LH
BS/LS
XLE/LH
Global
11.9
T/L
4.89
LCB% (+ aft)

0.70
0.07
0.15
0.131
-4.6%

At each design speed, the optimization algorithm


created some 2000 set of parameters with an effective
total calculations of about 900 (only the feasible
geometries were calculated). The considered speeds
were Fn=0.30, 0.35, 0.41, 0.50, so ranging from
medium to high speed. The four best geometries found
at the end of the optimization are represented in figure
14, which shows the 2D plot of the hull generatrix
curves together with the strut profile and the reference
S58 profile. It may be noted that the shape of the four
best candidates (the ones which scored the 4 minimum
values of the wave resistance) are very close together,
although their FV vectors are distant in the domain
space. The corresponding generated 3D geometries are
represented in figure 3.
Figure 13 presents an example of correlation plots
between the objective function and the values assumed
by the geometrical parameters found during the
optimization, at Fn=0.41. It is clear the distinct increase
of wetted surface (S) needed to obtain a lower
resistance (the wavy form of the curve increase its

Des
Fn
0.30
0.35
0.41
0.50

/
(2L)
0.28
0.38
0.53
0.78

R0
2.02
1.74
2.60
3.22

r1
0.30
1.19
1.95
2.18

m2 m1
0.60
0.68
0.73
0.77

m3 m1
0.44
0.40
0.23
0.28

m2 m3
0.16
0.28
0.50
0.49

LCB

%
-6.1
-1.6
-0.8
-0.6

S/
2/3
12.1
12.3
12.4
12.4

Of course, the positive interference effects are


valid around the design speed only: in general at other
speeds the optimized hull will, on the contrary,
experience negative interference effects. This fact is
clearly shown by the comparison of the wave
resistance coefficients of the four optimized hulls over
a wide speed range, represented in figure 11. It turns
out, in fact, that the hull which is optimized for
Fn=0.30 is worse at highest Fn, with respect to the hull
optimized for these highest speeds. The contrary holds
for hulls optimized at high speed, which have the
highest wave resistance at low Froude numbers. In the
same figure it is reported also the wave resistance
calculated for the reference hull, which turns to be the
worst in absolute in the whole high Fn numbers range.
The above correlation is confirmed by the
examination of the generated wave pattern at different
speeds. This comparison is shown in figure 16.
Evidently, each optimized hull creates the lowest
waves at its optimum speed, while in general the
conventional SWATH produced always the highest
waves at high speeds. The wave cancellation effects are

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

visible mainly in the transversal wave system induced


between the demi-hulls and in the wake, but also in the
divergent wave components, important at the highest
Froude numbers.
If the optimization criteria incorporates a weighted
average of the resistance at different speeds, as
formulated in (27), then of course, the optimized hull
forms will appear different, but in any case always with
a double peak shape. In this case the resistance curve
will also show a compromise trend with respect to the
ones shown in figure 11, depending on the weight
given at the wave resistance at each considered speed.
From the designer point of view, however, it is
more interesting the comparison of the results obtained
with the viscous-inviscid automatic optimization. The
calculations were made for a full scale length of 80m,
keeping the same maximum diameter, draft and
prismatic coefficient of the reference hull (table 2) as in
the previous optimization cases.
Table 4 present the characteristics of the optimum
hull forms found at the same target speed of the
inviscid design case. Figure 15 shows the generatrix
curves of the optimized hull forms obtained with
viscous-inviscid method (plain thick line) in
comparison with those obtained at the same speed with
the inviscid method (dashdot line).

part is dominant with respect to the viscous part of the


total resistance. This concept of dominance of the
various components of the object function is well
rendered by the two Pareto plots of figure 10, which
represent the viscous resistance reduction vs. the wave
resistance reduction, for all the calculated cases during
the optimization at Fn=0.35 and Fn=0.41. The four best
cases are positioned all on the Pareto frontier which
defines the limit curve on which it is not possible to
decrease one component of the object function without
increasing also the other. From all the above reasons it
is natural that the optimum solution moves towards
higher viscous resistance (with respect to the reference
hull form) which are more than compensated by the
considerable decrease obtained on the wave resistance.
Interesting, at last, is the comparison between the
predicted total resistance for the optimized hulls. As
expected the optimized hull for Fn=0.30 (O-30) has the
lowest predicted total resistance in the lower speed
range, while already above 18 knots the other three
hulls (O-35, O-40, O-50) experiment the lower
resistance. Also in this case the conventional hull form
with S-68 hull, is the worst at the highest speeds, due to
its high wave resistance properties.

Table 4: Main characteristics of the optimized hull forms


with viscous-inviscid method

Without repeating main conclusions already drawn in


the previous sections, we summarize the followings.
The validity of our viscous-inviscid CFD method
for the prediction of the total resistance of SWATH
like hull shapes has been verified against experimental
results, with very satisfactory, as even unexpected,
correlations. The possibility to use an automated
optimization procedure involving parametric geometry
generation methods, a viscous-inviscid CFD method
and an efficient global optimization algorithm, is
proven to be effective and brought in general to
acceptable and feasible solutions. The importance in
considering the total (wave+friction+viscous_pressure)
resistance for the optimization of the SWATH hull
forms is also hopefully demonstrated in the paper and
obviously depends on the design speed. The
advantages in using a non-conventional underwater
hull forms, shaped in such a way to reduce total
resistance, is also demonstrated in the paper. The
maximum gains in terms of wave resistance, respect to
a (non optimized) conventional solution can be up to
70%, while in terms of total resistance the gain is more
limited but still considerable (order of 30%). These
gains, naturally, vary with the relative Froude number.
Some topics for future developments regard:
The thin boundary layer method used could be
improved and made more robust, especially in the
prediction of the transition location and separation
point. The parametric geometry generation method,

Des
Fn
0.30
0.35
0.41
0.50

/
(2L)
0.28
0.38
0.53
0.78

R0
0.05
2.87
1.79
1.62

r1
0.41
1.15
1.14
1.75

m2 m1
0.15
0.64
0.69
0.63

m3 m1
0.08
0.58
0.35
0.25

m2 m3
0.07
0.06
0.34
0.38

LCB

%
-4.4
-4.6
-2.0
-0.8

S/
2/3
11.8
12.3
12.3
12.3

Being the total resistance used as objective function for


this new series of optimizations, the best hull form will
be the one which realizes the best compromise between
the (usually contrasting) optimum shapes to minimize
each individual component of the total resistance
(friction, viscous-pressure and wave resistance, in our
scheme). The compromise implicitly takes into account
the share which each resistance component has in the
total resistance at each considered design speed. For
instance, the hull optimized at Fn=0.30, where the
wave resistance is not the main part of total resistance,
shows, with respect to the inviscid case, a reduced
gradient of the generatrix curve at the L.E., and a lower
peak to through distance. This modifications tends to
reduce the resistance due to friction and b.l. separation,
which result to be the main components of the total
predicted resistance, at this speed.
Same type of trend in shape modification is noted
increasing the speed up to Fn=0.50, at which, on the
contrary, the optimum hull form is nearly the same as
for the inviscid case. At this speed, in fact, the wave

CONCLUSIONS AND FUTURE PROSPECTS

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

though very efficient and inherently respecting the


design volume constraint, demonstrated some
weakness in global application, giving from half to one
third of unfeasible solutions, which in general slower
the convergence of the optimization algorithm. Alternative formulations are being tested using splines.
More optimization cases will be run, considering
different design scenarios, including also the effect of
separation of the demi-hulls, the dimension and shape
of the struts, the number of underwater hulls, etc.
AKNOWLEDGEMENTS

Author wishes to thank Rodriquez Engineering for


according to publish towing tank test results of the
SWATH like demi-hull together with the validation
study made for the an ongoing research project with
them dedicated to the design of a hydrofoil hybrid
SWAMH. Special thanks to Claudia for her continuous
support.

REFERENCES

Brizzolara S., Bruzzone D., Cassella P., Scamardella I.,


Zotti I., "Wave Resistance and Wave Pattern for High
Speed Crafts; Validation of Numerical Results by
Model Tests", Proc. XXII Symposium on Naval
Hydrodynamics, Washington D.C., 1998, pp. 69-83.
Brizzolara S., Bruzzone D. (2000), "Numerical Wave
Resistance and Dynamic Trim for High Speed Crafts",
Proc. NAV 2000 Int. Conference, Venice, Vol.I, pp.
4.2.1-4.2.13.
Brizzolara S., Parametric Optimisation of SWATH
Hull Forms by Systematic Use of CFD Methods,
Proc. ISOPE 2003 Int. Conference on Offshore and
Polar Engineering, Honolulu, Hawaii, 2003.
Bruzzone, D., "Numerical Evaluation of the Steady
Free Surface Waves", Proc. CFD Workshop Tokyo,
Ship Res. Inst. Tokyo, Vol.I, 1994, pp. 126-134.
Cebeci T., An Engineering Approach to the
Calculation of Aerodynamic Flows, Springer Verlag,
1998, ISBN 3-540-65010-5.
Curle H., A Two-Parameter Method for Calculating
the Two-Dimensional Incompressible Laminar
Boundary Layer, Journal of Royal Aeronautical
Society, Vol 71, 1967.
Dvorak F.A., Maskew, B., Woodward, F.A., Investigation of Three-Dimensional Flow Separation on
Fuselage Configurations, Analytical Methods Inc.,
USAAMRDL-TR-77-4, March 1977.
Gertler M., Resistance Experiments on a Systematic
Series of Streamlined Body of Revolution for
Applications to the Design of High Speed
Submarines, DTMB report C-297, 1950.
Harries S., Parametric Design and Optimization of
Ship Hull Forms, PhD Thesis, TU Berlin, Mensch &
Buch Verlag, 1998.

Maskew B.: Program VSAERO, A Computer


Program for Calculating the Nonlinear Characteristics
of Arbitrary Configurations, User's Manual, NASA
CR-4023, Sept. 1987.
Nash J.F.; Hicks, J.G., An Integral Method Including
the Effect of Upstream History on the turbulent Shear
Stress, Computation of Turbulent Boundary Layers,
Vol. 1, Proceed. AFOSR-IFP-Stanford Conference,
Stanford University Press, 1969.
Papanikolaou A., Zaraphonitis G., Androulakakis M.,
Preliminary Design of a High-Speed SWATH
Passenger/Car Ferry, Marine Technology, Vol. 28,
no.3, May 1991, pp. 129-141.
Peri D., Rossetti M., Campana E., Design
Optimization of Ship Hulls via CFD Techniques,
Journal of Ship Research, Vol.45, 2001, n.2, pp.140149
Price K., Differential Evolution: A Fast and Simple
Numerical Optimizer, Proceedings NAFIPS, 1996,
Berkeley, California, pp. 524-527.
Salvesen N., Von Kerczek C.H., Scragg C.A., Cressy
C.P., Meinhold, M.J., Hydro-numeric Design of
SWATH Ships, SNAME Transactions, Vol. 93, 1985,
pp.325-346.
Storn R., Price K., Differential Evolution A Simple
and Efficient Adaptive Scheme for Global
Optimization over Continuous Spaces, TR-95-012,
March 1995, Int. Computer Science Institute, Berkeley,
California.
Storn R., On the usage of Differential Evolution
Algorithms for Function Optimization, Proceedings
NAFIPS, 1996, Berkeley, California, pp. 519-523.
Thwaites B. Approximate Calculation of the Laminar
Boundary Layer, Aero Quarterly, Vol. 1, 1949.
Various, papers of Session III Hull Modelling and
hydrodynamic Optimisation, NAV 2003 Coonference
Proceedings, Palermo, 2003.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Figure 9: Correlation plots between the position of the two


relative extremes m1 and m2 (top) on the generatrix curve of
the underwater body of revolution, obtained from all the
cases calculated for the optimization at Fn=0.41. Bottom:
same correlation but for Fn=0.35.

Figure 10: Pareto plot of the wave resistance gains vs.


viscous resistance gains during the optimization of the hull
forms at two different speeds: Fn=0.35 (top) and Fn=0.41
(bottom). Reference resistance calculated for the S58 hull.
2.5

0.5

2.0

Fn

RT [MN]

Series-58
Optimum Hull for Fn=0.30
Optimum Hull for Fn=0.35
Optimum Hull for Fn=0.41
Optimum Hull for Fn=0.50
Fn

0.45

1.5

0.4

1.0

0.35

0.5

0.3

0.0

0.25
10

Figure 11: Comparison of the predicted wave resistance


coefficients for the four optimized hulls (with respect to wave
resistance only) over a large speed range. The wave
resistance coefficient of the reference hull is also represented.

15

20

25

V [knt] 30

Figure 12: Comparison of the predicted total resistance for


the four optimized hull (with respect to full scale total
resistance) over the complete speed range. Predicted
resistance of the reference hull S58 is also represented.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Figure 13: Correlation plots between geometry characteristics and calculated wave resistance for all the about 900
cases analysed during the automatic optimization procedure of SWATH hull, at Fn=0.41.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Fn=0.30

Fn=0.41

Fn=0.35

Fn=0.50

Figure 14: Best underwater hull generatrix curves found at the end of the optimization procedure for wave resistance alone, at four different design Froude numbers.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Fn=0.30

Fn=0.41

Fn=0.35

Fn=0.50

Figure 15: Comparison between hull generatrix curve optimized with the viscous-inviscid method (plain thick curve) and those optimized with the inviscid method
only (dash-dot curve), at four different design Froude numbers. The generatrix curve of the reference Series 58 hull (plain thin curve)is also plotted.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

S-58,Fn=0.30

O-35,Fn=0.30

O-30,Fn=0.30

O-50,Fn=0.30

S-58,Fn=0.35

O-35,Fn=0.35

O-30,Fn=0.35

O-50,Fn=0.35

S-58,Fn=0.50

O-35,Fn=0.50

O-30,Fn=0.50

O-50,Fn=0.50

Figure 16: Contour plots (/LH*100) of the wave patterns generated by the original hull (S58) and by the three
optimized hulls (O-30, O-35, O-50), each optimized at a different optimization speed (Fn=0.30, 0.35, 0.50).

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Theoretical Hull Form Optimization


for Fine Higher-Speed Ships
Keh-Sik Min (Hyundai Heavy Industries, Ulsan, Korea)
Yeon-Seung Lee, Seon-Hyung Kang and Bum-Woo Han
(Hyundai Maritime Research Institute, HHI, Ulsan, Korea)
ABSTRACT
Systematic theoretical and experimental studies have
been performed to establish fuel-economic hull form
design method for fine higher-speed ships like
container carriers or naval ships. In this study,
extensive fundamental studies on the relation between
hull form characteristics and ship resistance have been
conducted in advance, and the results have been
utilized in the preparation of the hull form design
method.
Seven actual ships, that is, five best container carriers
in different size and two destroyers recently built at
Hyundai Heavy Industries were selected and their hull
forms were re-designed by the prepared hull form
design method maintaining the main dimensions and
the underwater volume the same. Model tests were
conducted for both existing and re-designed hull form
ships, and the test results were compared. It has been
proved by the comparison of the test results that
performance characteristics of the re-designed hull
form ships are superior to those of the existing ships.

INTRODUCTION
Hull form design technology is one of the most
important core-technologies in shipyard operation.
The importance of hull form design may not be too
much emphasized. Hull form design is particularly
important in two aspects, that is, as the first or
beginning stage of new building project and as the
index of determining the economy of ship
operation. For the effective operation of ship yard,
therefore, it is extremely important to possess the
capability of designing superior hull forms in short
time. Recognizing the importance of hull form, the
authors have long carried out the research works on the
theoretical hull form design method (Min, K-S et al.,
1983, 1984, 1987, 1992, 1998, 2003).

Prior to establishing hull form design method,


extensive fundamental studies on the relation between
hull form characteristics and ship resistance have been
performed systematically utilizing the linearized thinship
theory
and
mathematical
hull
form
equations. Furthermore, studies on the optimization of
main hull form characteristics were numerically
conducted to investigate more overall aspect of ship
resistance characteristics. The results from the above
studies have been utilized in the preparation of the hull
form design method.
The detail subjects included in the fundamental studies
are as follows:
- Theoretical resistance calculation method
- Effect of the waterline equation
- Effect of the underwater section shape
- Effect of the waterline optimization
- Effect of the overall optimization
In fact, each of the above items is a large-scale R & D
subject. The authors have carried out thorough
theoretical studies first for each of the above subjects,
and tried to verify the validity of the theoretical studies
by comparing the results from theory with those from
experiment as much as possible.
All the tests for the fundamental study were conducted
in the deep-water towing tank of Maritime Research
Institute, Hyundai Heavy Industries. The tests for this
study are different from the ordinary model test, since
the ships manufactured according to the designed hull
form are not model ships. They are small ships for the
verification of the results of theoretical study.
As mentioned, this study has been performed for the
purpose of establishing fuel-economic hull form design
method for fine higher-speed ships with superior
resistance and propulsion characteristics. The contents
and the results of this study shall be presented and
discussed in this paper.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

THEORY

S = 2

Resistance calculation
For the theoretical calculation of ship resistance, the
traditional resistance theory has been utilized together
with the residual resistance expression derived from the
linearized thin-ship theory. In order to represent hull
form and flow characteristics, the coordinate system
shown in Figure 1 has been adopted.
z

b=B/2

y = f(x,z)
= L/2

-b

-T

Figure 1: Coordinate system for the presentation of


hull form and flow characteristics around a ship hull

(1)

(2)

In equation (2), RT, RF and RR represent total resistance,


frictional resistance and residual resistance,
respectively. Frictional resistance is expressed as
follows:

RF =

1
CF v 2 S
2

(5)

RR = P nx dS

(6)

In equation (6), P and nx represent pressure and xcomponent of inward unit normal vector n . If the
hull source obtained by distributing Havelock Sources
on ship's central plane are substituted into equation (6)
with the application of thin-ship theory and if noncontributing terms are eliminated, then famous
Michell's Integral for residual resistance is obtained as
follows (Wehausen) :

RR =

4 g 2
v2

represents the hull form, and L, B, T and represent


ship length, beam, draft and half of the ship length,
respectively. According to the traditional resistance
theory, the total resistance of a ship is assumed to be
composed of the frictional resistance and the residual
resistance as follows:

RT = RF + RR

1 + f x2 + f z2 dxdz

In equation (5), f x and f z represent partial derivatives


of f ( x, z ) in x-direction and z-direction, respectively.
The resistance on the ship hull due to pressure
distribution, that is, residual resistance, could
symbolically be expressed as follows:

In Figure 1,

y = f ( x, z )

d sec e
3

2
0

g
v2

( z + )sec2

cos 2 ( x )sec
v

(7)

For the sake of convenience, replace sec by cosh u in


equation (7) and adopt the following symbols, then, the
total resistance is expressed as follows:

RT = C F v 2 1 + ( f x ) 2 + ( f z ) 2 dxdz

+ A du cosh 2u[ P 2 (cosh u ) + Q 2 (cosh u )]


0

(8)

{ }

P (cosh u )
cos
( x )
= dxdzf x ( x, z ) e z
Q (cosh u )
sin
A=

(3)

dxdz d d f ( x, z) f ( , )

g
g
4 g
, = 2 cosh u , = 2 cosh 2 u
2
v
v
v
2

Hull form equation


In equation (3), , CF, RN, v and S represent water
density, frictional resistance coefficient, Reynolds
Number, ship speed and wetted-surface area of a ship
hull, respectively. Frictional resistance coefficient, CF,
is found from the ITTC - 57 correlation line, that is,

CF =

0.075
( Log10 RN 2) 2

(4)

In equation (3), the exact mathematical expression of


the wetted-surface area is as follows:

In order to find out the hull form with minimum


resistance, it is necessary to introduce a hull form
equation that describes the hull shape mathematically.
The hull form equation should not only properly
describe the geometric shape of a ship hull, but also be
as simple as possible so that mathematical treatment is
possible when substituted into the resistance equation.
However, it is by no means easy to describe the
complicated hull shape satisfactorily by a simple
equation.
As the early stage study for the

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

mathematical hull form equation, the following


elementary type hull form equation has been adopted:

y = f ( x, z ) =

B
X ( x) Z ( z )
2

(9)

In equation (9), X(x) and Z(z) represent waterline plan


and body plan, respectively. In equation (9), it is
recognized that section shape has no relation with the
longitudinal direction. In other words, waterline shape
and section shape are independent of each other. In
actual ships, however, section shape varies according to
the longitudinal position. Therefore, the authors have
devised realistic hull form equations. They are
symbolically expressed as follows:

y = f ( x, z ) =

B
X ( x ) Z ( x, z )
2

(10)

The hull forms mainly used in this study shall be


discussed.

simple, yet well describes actual ship's section shapes


and is able to change the shapes along the longitudinal
position as designer's desire.
There are two section equations created by the authors
as follows:
- No. 1 section equation:
Z(x, z) = [ 1 + (z/T) ]n(x) , - T z 0

(12)

- No. 2 section equation:


Z(x, z) = 1 - | z/T |n(x) ,

(13)

-T z 0

If the form exponent, n(x), in equations (12) or (13) is


systematically varied, the shape of the curve is also
systematically varied as shown in Figures 2 and 3.
The actual ship's section shapes could be well
described utilizing the shapes represented by equation
(12) or (13).
z

Equation of waterline plan

n(x)

, -T z 0 b
1

Z(x,z) = [1 + (z/T)]

0.0

Z(x,z) = 1 - |z/T|

0.6

1.67

+ a3 x /

+ a4 ( x /

+ a6 ( x /

(11)

In equation (11), ai (a2, a3, a4 and a6) are the unknown


hull form coefficients to be so determined as to
generate the waterline shape with minimum resistance.
Equation of body plan
It is much more difficult in practice to create the
equation of body plan which sufficiently well expresses
the actual ship's section shapes, and yet is able to vary
the shapes along the longitudinal position according to
designer's desire. In fact, there has never been
successful single equation until now in the world which
sufficiently well describes actual ship's section shapes.
However, the authors have succeeded in representing
the body plan by a single equation that is extremely

0.4

2.5

X ( x) = 1 + a2 ( x /

0.3

3.3

A vast amount of the fundamental study has been


performed on the effect of the equation of design load
waterline for all the possible combinations from the 4thorder to the 10th-order polynomial equations. The
results shall be briefly discussed later. In this study,
however, the following 6-degree polynomial equation
has been mainly used.

, -T z 0

n(x) = 0.2

n(x) = 5.0

The authors have long tried numerous different


equations to represent waterline shape for fine higher
speed ships and realized that polynomial type equation
generally gives good results.

n(x)

0.0

0.8
1.0

1.25
1.0
0.67

1.5
2

0.5
0.33

3
5

0.2

10

0.1

25

0.04

50

-T

Figure 2:

0.02

No. 1 Section

-T

Figure 3:

No. 2 Section

For the fundamental studies, however, the section


shape is not important and it was decided to use No. 2
section equation.
Optimization of waterline shape
Traditional resistance theory may not be much useful
quantitatively, yet could be effectively utilized for the
qualitative purpose. The optimization of hull form,
particularly the optimization of waterline shape is one
of successful examples of the qualitative use of the
resistance theory.
In order to calculate the frictional resistance, the
wetted-surface area expressed in equation (5) should be
calculated first.
In the course of hull form
optimization, however, the exact wetted-surface area
equation expressed in equation (5) cannot be handled
mathematically.
Furthermore, linearized thin-ship
theory has been utilized to derive residual resistance.
For the sake of consistency and to make mathematical
treatment possible, therefore, the exact expression of
wetted-surface area is also linearized and the linearized
total resistance equation is obtained as follows:

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

1
1
RT = CF v 2 [1 + ( f x ) 2 + ( f z )2 ]dxdz
2
2

(14)

+ A du cosh 2u[ P 2 (cosh u ) + Q 2 (cosh u )]


0

In order to find the optimum waterline shape, that is,


the waterline shape with minimum resistance, it is
necessary to express the resistance equation by the
unknown hull form coefficients. In order to do that,
the hull form equation is substituted into the total
resistance equation. After a long complicated and
tedious mathematical treatment, an expression for the
total resistance in terms of the unknown hull form
coefficients is obtained as follows:
RT = R0 + R2 a 2 + R3 a3 + R4 a 4 + R6 a 6
1
+ R22 ( a 22 ) + R23 ( a 2 a3 ) + R24 ( a 2 a 4 ) + R26 ( a 2 a 6 )
(15)
2
1 2
+ R33 ( a3 ) + R34 ( a3 a 4 ) + R36 ( a3 a 6 )
2
1
+ R44 ( a 42 ) + R46 ( a 4 a 6 )
2
1 2
+ R66 ( a 6 )
2

In equation (15), R0 , R2 , ... , R66 are the resistance


coefficients of the unknown hull form coefficients, a2,
a3, a4 and a6 to be optimized.
Lagrangian multiplier method has been selected for the
optimization.
In order to apply the method,
Lagrangian equation should be constructed first
expressed by the unknown hull form coefficients ai and
Lagrangian multipliers and . The following
object function and boundary conditions have been
utilized to construct Lagrangian equation:
- Object function:

Total resistance

f1 = RT = RF + RR

(16)

- Boundary conditions

B
(1 +
2

)=0

(17)

i=2

Constant volume:

f3 =

T
= 0,
ai

f ( x, z )dxdz V = 0

(18)

Lagrangian equation is constructed as follows:

T ( a i , , ) = f 1 + f 2 + f 3

(19)

T
= 0,

T
=0

(20)

Then, a linear system is obtained with respect to the


unknown hull form coefficients and Lagrangian
constants in the following form:

AX = B

(21)

By solving the linear system, the hull form equation


that leads the least resistance is obtained.

FUNDAMENTAL RESEARCH
Theoretical resistance calculation
In order to investigate the accuracy and the practical
applicability of the theoretical resistance prediction, a
simple elementary hull form was selected for
verification.
For the sake of convenience, the selected hull form
shall be called the simple hull form and denoted by
SHF. The equation of the simple hull form is as
follows:

y = f ( x, z ) = b X ( x ) Z ( z )
= b[1 ( x / )2 ] [1 | z / T |2 ]

(22)

In equation (22), b represents the half beam, i.e.,


b = B/2.
The main characteristics of the simple hull form ship
have been summarized in Table 1.
Table 1:

Closure condition:

f 2 = f ( ,0 ) =

If Lagrangian equation (19) is partially differentiated


with respect to the unknown hull form coefficients ai,
and Lagrangian constants and , and set to be
zero, that is,

Main characteristics of the simple hull form

Length P.P.(LPP) or
Waterline length(LWL)
Beam (B)
Draft (T)
Volume (V)
Block coeff. (CB)
Length / Beam ratio (L/B)
Beam / Draft ratio (B/T)
Length / Draft ratio (L/T)

Copyright National Academy of Sciences. All rights reserved.

8.000 m
0.800 m
0.320 m
0.910 m3
0.444
10.0
2.5
25.0

Twenty-Fifth Symposium on Naval Hydrodynamics

Total Resistance Coeff. (CT) x 10

7.0
6.5
6.0

Simple Hull Form


: Computed
: Measured

5.5
5.0
4.5
4.0

0.2

0.3
0.4
0.5
Froude Number (F N)

0.6

0.7

Figure 4: Comparison of total resistance coefficient between computed and test results
700

Total Resistance (R T, N)

500

Simple Hull Form


: Computed
: Measured

400
300
200
100
0
0.1

: (R R

100

0.2

0.3
0.4
0.5
Froude Number (FN)

0.6

COMPUTED

-R R MEASURED) / R R MEASURED x 100

75
50
25
0
-25
-50
-75
0.1

0.2

0.3

0.4
0.5
Froude Number (F N)

0.6

0.7

Figure 6: Difference between computed and measured total resistance for the simple hull form

Effect of the waterline equation


To optimize the waterline shape, the waterline is
represented by an equation with the unknown hull form
coefficients, and the coefficients are so determined that
ship's total resistance is to be minimum. Before the
general application of the idea to actual hull form
design, the authors had decided to investigate the effect
of the waterline equation on the optimum waterline
shape and resistance itself. A wide variety of different
types of equations had been tried such as polynomial
type, trigonometric type, elliptic type, hyperbolic type,
etc. In fact, this is another serious field of research
and cannot be discussed here in detail due to time and
space. In this paper, however, a brief discussion shall
be made only for the case of polynomial type waterline
equation. First of all, the following basic type of
polynomial equation was introduced.

+ a5 x /

3.0

600

Simple Hull Form

125

X ( x) = 1 + a2 ( x /

3.5

2.5
0.1

150

Difference (%)

Figures 4 and 5 show the comparisons between


calculated and measured results of total resistance
coefficient and total resistance itself for the simple hull
form ship. When the quantitative aspect of the
theoretical resistance calculation is examined from
Figures 4 and 5, it is known that theory overpredicts
resistance in the low Froude Number region and
slightly underpredicts in the high Froude Number
region. In qualitative aspect, theory tends to
exaggerate the interaction phenomenon between bow
and stern waves in the low Froude Number region. As
ship speed increases, the general trend of the
theoretical prediction agrees well with the measured
result. Particularly, the qualitative trend for the hump,
hollow and the final hump at Froude Numbers around
0.3, 0.35 and 0.5 agrees remarkably well with the test
result. Figure 6 shows the difference between the
theoretical calculation and test results. In general, the
difference is decreasing as Froude Number
increases. In other words, the result of theoretical
calculation and the test result are in general
approaching each other as Froude Number increases
and the theoretical calculation becomes to have
sufficient practicability even quantitatively as well as
qualitatively.

+ a3 x /

+ a6 ( x /

)
6
10
) + + a10 ( x / )
3

+ a4 ( x /

(23)

With the waterline equation (23), the optimum


waterline shape and the total resistance for that
particular optimized case were investigated for all the
possible combinations of terms in equation (23) from
the 4th-order to the 10th-order with the systematic
variation of design variables such as 2 different section
shapes with 5 different form exponents, 5 different
block coefficients and 5 different Froude
Numbers. This work is really an enormous amount of
the actual designs and calculations for 116,500
different cases of combination. Table 2 clearly shows
the above mentioned number of cases. Figure 7 shows
an example of the optimum waterline shape according
to the waterline equation when waterplane area
coefficient and Froude Number are constant.

0.7

Figure 5: Comparison of total resistance between


computed and test results

For the study on the effect of the waterline equation,


however, no definite conclusion could be derived in
spite of the enormous amount of investigations. In this

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

study, therefore, it was decided to select the equation


(11) of 5-term 6-degree polynomial for the
representation of waterline shape due to simplicity and
effectiveness.

Nondimensional HB

1.0

FN = 0.2

C W = 0.667

0.8

HF2-3-4
HF2-3-4-6
HF5-7-8-9-10

0.6
0.4
0.2
0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

Nondimensional Entrance Length

Figure 7: Variation of the optimum waterline shape


according to different waterline equations
Table 2: Total number of optimum waterline shape
design and resistance calculation
Design
Variable
Section
Shape

Description

Number
n

No. 1 Section : Z(z) = (1 + z)

No. 2 Section : Z(z) = 1 - | z |n


No. 1 Section : 1 , 1 , 1 , 1 , 1
3 4 5 6 7

Form
Exponent
(n)
No. 2 Section : 3, 4, 5, 6, 7

Block
0.400, 0.450, 0.500, 0.525,
5
Coeff.(CB) 0.550
Froude 0.200, 0.225, 0.250, 0.275,
5
Number(FN) 0.300
4 - term : 9C3
84

Terms of
Polynomial

5 - term :

9C4

126

6 - term :

9C5

126

7 - term :

9C6

84 466

8 - term :

9C7

36

9 - term :

9C8

10 - term : 9C9
Total Number of Cases :
2 5 5 5 466

(12) and (13) were used as the section equations.


Theoretical investigations for total 350 combinations in
5 block coefficients, 5 Froude Numbers and 14
different form exponents were made. In order to
validate the theoretical investigations, 3 hull forms
with No. 2 section shape having form exponents 5, 9
and 11 were designed according to the main
characteristics summarized in Table 1, and 3 small
ships were manufactured.
Table 3 shows the typical theoretical and experimental
examples for the variation of resistance characteristics
according to the systematic variation of the underwater
section shape. Table 3 also shows the comparison (%
difference) of resistance in terms of that for the section
shape with form exponent 5, i.e., n = 5. Considering
the test accuracy, the experimental result sufficiently
confirms the result of the theoretical study.
Table 3:
Variation of resistance characteristics
according to the variation of underwater section shape
and comparison

- No. 2 Section : Z ( z ) = 1 | z / T |n
- FN = 0.25
Form Exponent (n)
Characteristics
Theoretical
Calculation

RT (N)

11

52.12

51.66

51.79

52.11

52.53

% Diff. 100.89 100.00 100.25 100.87 101.68

Experiment RT (N)
(Measurement) % Diff.

53.42

54.15

54.07

100.00

101.37 101.22

From Table 3, a very important conclusion could be


derived that variation of the underwater section shape
within the practical range has almost no effect on the
resistance characteristics as long as the waterline shape
is optimized with respect to that particular section
shape.

1
116,500

Effect of the underwater section shape

Another important investigation was made on the effect


of the underwater section shape, that is, variations of
the optimum waterline shape and resistance
characteristics according to the systematic variation of
the underwater section shape were carefully
investigated when the main dimensions and the
underwater volume were maintained the same for the
given ship speed. For this investigation, equations

Effect of the waterline optimization

In this study, the effect of the waterline optimization on


the
resistance
characteristics
was
seriously
investigated. In order to do that, 3 waterlineoptimized hull forms were designed at Froude
Numbers of 0.2, 0.3 and 0.5 with the same main
characteristics summarized in Table 1, and it was
decided to denote the 3 hull forms by WHF-1, WHF-2
and WHF-3, respectively. Therefore, the main
characteristics for the simple hull form ship and 3
waterline-optimized hull form ships are all the same
even if their hull shapes are different.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Non-Dimensional HB

1.0

SHF
WHF-1
WHF-2
WHF-3

CW = 0.667

0.8
0.6

8.0

Total Resistance Coeff. (CT) x 10 3

Figure 8 shows the comparison of the waterline shapes


between the simple hull form and 3 waterlineoptimized hull forms and Figure 9 shows 4 actual small
ships for validation.

0.4

Computed
: SHF
: Optimized Hull Form 3 for all FN
: WHF-3

6.0

4.0

0.1

0.2

0.3

0.2
0.0

: Measured (WHF-3, Free Trim)


: Point of Optimization

0.4

0.5

0.6

0.7

Froude Number (FN)


0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

Non-Dimensional Entrance Length

Figure 8: Comparison of the waterline shapes


between the simple hull form and 3 waterlineoptimized hull forms

Figure 10: Comparison of total resistance coefficient between the simple hull form and the optimized
hull forms
600

Computed
: SHF
: Optimized Hull Form 3 for all FN
: WHF-3

Total Resistance (R T, N)

500
400

: Measured (WHF-3, Free Trim)


: Point of Optimization

300
200
100
0
0.1

0.2

0.3

0.4

0.5

0.6

0.7

Froude Number (FN)

Figures 10 and 11 show the resistance characteristics


for the ship whose design load waterline is optimized at
Froude Number 0.5. Theoretical resistance calculation
shown in the Figures 10 and 11 was made with the
assumption that there was no trim. However, test was
conducted at the free-trim condition. As ship speed
increases the situation becomes more and more
different.
Therefore, it was decided to conduct
another resistance test with the ship fixed so that trim
might not occur. Figures 12 and 13 show the
comparison between the result of the theoretical
calculation and the test result for the case without trim.
Here, it should be noted that Figures 10 and 11 also
show the theoretical prediction of the resistance
characteristics for the ship whose waterline shape is
optimized for all Froude Numbers. However, this is
fictitious because an optimized hull form for all Froude
Numbers does not exist.

Total Resistance Coeff. (CT) x 10 3

Due to limited space, it is difficult to discuss resistance


characteristics for all 3 waterline-optimized hull form
ships. Therefore, a brief discussion shall be made only
for the case of Froude Number 0.5 as a typical
example.

8.0

: Computed (WHF-3)
Measured (WHF-3)
: Free Trim
: Fixed
6.0

: Point of Optimization

4.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

Froude Number (FN)

Figure 12: Total resistance coefficients of the optimized hull forms


600

: Computed (WHF-3)
500

Total Resistance (R T, N)

Figure 9: The simple and 3 waterline-optimized


hull form ships (SHF, WHF-1, WHF-2 and WHF-3)

Figure 11: Comparison of total resistance between


the simple hull form and the optimized hull forms

400

Measured (WHF-3)
: Free Trim
: Fixed
: Point of Optimization

300
200
100
0
0.1

0.2

0.3

0.4

0.5

0.6

0.7

Froude Number (FN)

Figure 13:
forms

Total resistance of the optimized hull

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Overall optimization

The effect of the waterline optimization on the


resistance characteristics has been investigated for the
given ship speed and for the same hull form
characteristics. However, it may be useful to examine
the resistance characteristics from more overall
aspect. In other words, it is necessary to investigate
not only the effect of the waterline optimization, but
also the effect of optimization of the main hull form
characteristics such as block coefficient and main
dimensions for the given ship speed and for the same
underwater volume.

In this way, 3 overall-optimized hull forms were


prepared and named OHF-1, OHF-2 and OHF-3,
respectively. Two small ships, that is, OHF-1 ship and
OHF-2 ship were manufactured with the same main
characteristics shown in Table 4. For OHF-3 ship,
however, a model ship with 1/2 scale ratio was
manufactured, since the ship length of 13.737 m was
too long. Figure 14 shows the comparison of the
design load waterline shapes between the simple hull
form and 3 overall-optimized hull forms, and Figure 15
shows 3 overall-optimized hull form ships.
0.7

Previously, it has been discussed that the optimum


waterline shape could be obtained for the given
underwater volume, main dimensions and ship
speed. In the opposite way, resistance could be
estimated
once
the
waterline
shape
is
known. Therefore, the simultaneous optimization of
the waterline shape and the main hull form
characteristics could be numerically accomplished by
optimizing the waterline shape and calculating the
resistance simultaneously for the systematic change of
design variables with constant underwater volume and
ship speed.
In this study, the overall optimization was performed
for the cases of 3 ship speeds, that is, 1.771 m/sec,
2.657 m/sec and 4.429 m/sec, respectively. These
speeds are equivalent to those of Froude Numbers 0.2,
0.3 and 0.5 for the simple hull form ship or for the
waterline-optimized hull form ships. Table 4 shows
the main characteristics of 3 overall-optimized hull
form ships determined through the above discussed
procedure.
Table 4: Main characteristics of the simple and the
optimized hull forms

- Constant underwater volume (0.910 m3)


Main
Simple Waterline-Optimized Overall-Optimized
Hull
Hull Forms
Hull Forms
Hull
Form
Form
Charact(SHF) WHF-1 WHF-2 WHF-3 OHF-1 OHF-2 OHF-3
eristics

Half Beam (m)

0.6

SHF
OH F-1
OH F-2
OH F-3

0.5
0.4
0.3
0.2
0.1
0.0
0 .0

1 .0

7 .0

Table 5: Definition of hull form symbols

Basic Hull Form


Speed (Simple Math.
of Opt. Hull Form)
(m/sec)
FN Symbol

B (m)

0.8

0.8

1.168 0.969 0.611

T (m)

0.32

0.32

0.334 0.323 0.244

1.771

0.2

L/B

10.0

10.0

5.0

7.5

22.5

2.657

0.3

B/T

2.5

2.5

3.5

3.0

2.5

4.429

0.5

4/9

4/9

0.4

0.4

4/9

7.186 7.116 7.080 6.205 6.828 9.297

6.0

At this moment, it is considered to be necessary to


clarify the hull form symbols to avoid confusion. For
this purpose, Table 5 has been prepared to clearly show
the definition of the hull form symbols.

5.839 7.269 13.737

7.130

5 .0

Figure 15: Three overall-optimized hull form ships


(OHF-1, OHF-2 and OHF-3)

8.0

S (m )

4 .0

Figure 14: Comparison of design load waterline


shape between the simple hull form and 3 overalloptimized hull forms

8.0

3 .0

E ntrance Length (m)

LPP (m)

CB

2 .0

WaterlineOptimized
Hull Form
FN Symbol

OverallOptimized
Hull Form
FN Symbol

0.2 WHF-1 0.234 OHF-1


SHF

0.3 WHF-2 0.315 OHF-2


0.5 WHF-3 0.382 OHF-3

The results of the overall optimization for all 3 hull


form ships are very interesting and dramatic. Due to

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

limited space, however, only the result for OHF-3 ship


shall be presented here. Figures 16 and 17 show the
dramatic change and the overall trend of the resistance
characteristics according to the optimization.
10

Overall review of the optimization

To have an overall view on the relation between the


optimization and the resistance characteristics, Table 6
has been prepared. Table 6 may be regarded as the
summary of the results of the optimization study.

Total Resistance Coeff. (C T) x 10 3

Computed
: SHF
: WHF-3
: OHF-3

In order to extrapolate model test result into that of full


scale ship, either one of the residual resistance method
or the form factor method should be used
nowadays. For the sake of consistency, residual
resistance method has been used in this study to
extrapolate model test result of OHF-3 ship into fullscale data.

Measured

: SHF
: WHF-3
: OHF-3

Point of Optimization : VS = 4.429 m/sec

: WHF-3
: OHF-3

5
4
3
2

VS (m/sec)

Figure 16: Comparison of CT between SHF and the


3rd optimized hull forms (WHF-3 & OHF-3)

It is considered that the difference between the result of


the theoretical calculation and the test result in Table 6
mainly comes from the occurrence of trim during tests
and also from the use of the linearized theory. The
error in model test may be another minor reason to
cause the difference.
HULL FORM DESIGN

600

Total Resistance (R T, N)

Computed
500

: SHF
: WHF-3
: OHF-3

400

Measured : SHF
: WHF-3
: OHF-3

Point of Optimization : V S = 4.429 m/sec

300

As discussed already, extensive fundamental studies on


the relation between hull form and ship resistance
characteristics have been performed in advance, and
the results have been utilized in the preparation of the
hull form design method.

: WHF-3
: OHF-3

200

100

VS (m/sec)

Figure 17: Comparison of RT between SHF and the


3rd optimized hull forms (WHF-3 & OHF-3)

For commercial ships, hull form is generally composed


of three parts, that is, fore-body part, aft-body part and
bow bulb part. The fore-body hull form is directly
optimized utilizing the linearized thin-ship theory
together with the mathematical hull form
equation. For the aft-body hull form design, however,
traditional potential theory is useless due to thick
boundary layer and possible separation along the ship
hull afterwards. Furthermore, special arrangement of
propulsion system should be considered. Therefore,

Table 6: Improvements in resistance characteristics according to the optimization of main hull form characteristics

- Constant underwater volume (0.910 m3)


Froude Number
VS
(m/sec)

WHF OHF
SHF
Series Series

SHF
Theory

1.771 0.200 0.200 0.234

Comparison of Resistance*
(Resistance Reduction Ratio, %)

Total Resistance(RT, N)

Test

42.07 37.32

WHF
Series
Theory

Test

37.50 35.59

2.657 0.300 0.300 0.315 109.86 102.54 101.44 104.94

OHF
Series
Theory

Test

SHF
and
WHF Series

SHF
and
OHF Series

WHF Series
and
OHF Series

Theory Test Theory Test Theory Test

32.49 36.77 -10.86 -4.62 -22.77

-1.46 -13.36 +3.31

78.81 85.76

-7.66 +2.34 -28.26 -16.36 -22.31 -18.27

4.429 0.500 0.500 0.382 385.70 388.65 361.36 360.75 231.22 258.71

-6.31 -7.18 -40.05 -33.43 -36.01 -28.28

* - : Decrease, + : Increase

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

no direct theoretical method of designing aft-body hull


form is available yet in the world.
In this study, therefore, the authors have developed an
indirect theoretical method of generating aft-body hull
form geometry by hydrodynamic based interactive
curve fitting with the prime effort to reduce the
resistance related to viscosity and to improve
propulsion characteristics. The main procedure for
both fore-body and aft-body hull form design for
commercial ships is in general as follows:
1) Design of the optimum load waterline.
2) Design of the bottom tangent line
3) Design of the side tangent line
4) Design of the body plan
5) Iteration if necessary
6) Calculation of hydrostatic and hydrodynamic
characteristics
For naval ships and cruise-liners, it is not always
necessary to go through all the steps of the above
procedure. In many cases, one or two steps are
enough with proper mathematical hull form equation to
generate the complete hull form. Two typical
examples for a commercial ship and a naval ship shall
be discussed in the performance evaluation section.
In order to design the most suitable bulb from the
overall operational viewpoint, ship's operational data
were collected from ship operators, and the operation
ratio between full load and ballast load was
analyzed. It would be much simpler to design the
optimum bulb for one particular loading
condition. From the viewpoint of the overall economy,
however, several loading conditions have to be
considered, and the performance for some particular
loading conditions may be somewhat sacrificed. The
bottom shape of the bulb has been modified so that the
added resistance, added propulsion power and impact
by slamming due to wave and ship motion in heavy
weather may be minimized.
The important items in the bow bulb design are as
follows:
- Bulb profile shape
- Bulb section shape
- Height of bulb center
- Bulb length
- Height of bulb upper part
- Size (width or thickness) of bulb
For the actual design of bow bulb, the standard shape
has been prepared from the authors long experience
and from the result of the study.

PERFORMANCE EVALUATION

When major part of this study was almost completed, it


was decided to evaluate the result of the study, that is,
the quality of performance comprehensively and
systematically. For this purpose, five best container
carriers and two naval vessels recently built at Hyundai
Heavy Industries were selected. The 5 selected
container carriers are 2,500 TEU, 4,800 TEU, 5,600
TEU, 7,200 TEU and 8,600 TEU carriers and 2 navel
vessels are 3,500 tonne-class and 5,000 tonne-class
destroyers. Due to limited space, however, a brief
discussion shall be made only for the 5,600 TEU
container carrier and 5,000 tonne-class destroyer, and
they will be denoted by CC5 and CDX,
respectively. The hull forms of these two vessels have
the best transportation efficiency in our database.
Hull forms of the selected vessels were re-designed by
the established hull form design method maintaining
the main characteristics the same.
The main
characteristics of 2 hull forms are summarized in Table
7. Figures 18 and 19 show the comparisons of hull
forms (body plans) for 2 selected vessels, respectively.
More than 6.0 m long model ships were manufactured
for all container carrier tests to evaluate complete
performance characteristics including propulsion
properties. However, 2.0 m long model ships were
manufactured for destroyers only for the resistance
tests, since propulsion characteristics are rather similar
for open-shaft system ships.
Main characteristics of the selected vessels

Table 7:

Characteristics

CC5

CDX

Length P.P (LPP)

(m)

271.0

138.0

Beam (B)

(m)

40.0

15.787

Draft (T)

(m)

12.5

4.924

81,860

5,461

Block Coeff.(CB)

0.5894

0.4967

Max. Section Area Coeff.(CM)

0.9691

(m)

6.0

not applicable

Length

(m)

7.0

Height from B.L.

(m)

11.5

Center Height from B.L (m)

8.0

25.0

0.250

Displacement ()

(tonnes)

Bilge Radius (RB)


Bulb

Design Speed (VS)


(knots)
Froude Number at the Design
Speed (FN)

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Existing

Re-designed

Figure 18: Comparison of hull form (body plan) for


5,600 TEU container carrier

Existing

Re-designed
Figure 22: Self - propulsion test of the re-designed
hull form ship for 5,600 TEU container carrier (CC5)

Figure 19: Comparison of hull form (body plan) for


5,000 tonne-class destroyer

Figures 20 and 21 show model ships of the existing and


the re-designed hull forms for 2 selected vessels,
respectively. All the model tests for the container
carriers were carried out in the deep-water towing tank
of Maritime Research Institute (HMRI), Hyundai
Heavy Industries. Figure 22 shows a typical selfpropulsion test of the re-designed hull form ship for
5,600 TEU container carrier. For the naval vessels,
resistance tests were conducted in the circulating water
channel of HMRI. Figure 23 shows a typical
resistance test of the re-designed hull form ship for
CDX. Comparisons of the test results of the redesigned hull form ships for each of 2 selected vessel
types are summarized in Table 8 in terms of percent
(%) of those for the existing hull form ships.

Figure 20:

Figure 21:

Figure 23: Resistance test of the re-designed hull


form ship for CDX
Table 8: Comparison of the performance characteristics of the re-designed hull form ships

- % of the performance characteristics for the existing


hull form ship
CC5
VS
(kts)
23.0
24.0
25.0
26.0
27.0

CDX

FN

EHP

BHP

0.230
0.240
0.250
0.260
0.270

98.60
98.60
98.29
97.59
96.62

97.52
97.48
96.71
95.84
94.82

VS
(kts)
25.0
27.0
29.0
31.0
33.0

FN

EHP

0.350
0.378
0.406
0.434
0.461

99.1
99.0
98.7
98.2
97.6

Two model ships of CC5

DISCUSSIONS AND CONCLUSIONS

Two model ships of CDX

This R & D work has been performed for the purpose


of establishing fuel-economic hull form design method
for fine higher-speed ships. As mentioned already, a
vast amount of the fundamental studies on the relation
between the hull form characteristics and the resistance
characteristics have been carried out in advance and the
results have been incorporated in the preparation of
hull form design method. From the results of the
whole study, the following discussions and conclusions
could be made:

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

- The resistance theory used in this study is


considered to be lack of practical applicability for the
quantitative purpose in overall viewpoint. However, the
theory is quite satisfactory in predicting the qualitative
trend, and hence, could be well utilized for the hull
form design of fine higher-speed ships. Particularly, it
could be well utilized for both quantitative and
qualitative purposes for fine ships with length-beam
ratio greater than 7.5 and operational speed above
medium speed range.
- The effect of the waterline equation is not
simple. An enormous amount of the optimum
waterline designs and resistance calculations for
116,500 combinations was made and analyzed to find
out the best polynomial equation for the representation
of the design load waterline shape. However, it was
not possible to derive generally applicable
conclusions. This subject is going to be presented
separately in the near future.
- For the effect of the underwater section shape, a
practically important conclusion could be derived that
variation of the section shape within the practical range
does not have much effect on the resistance
characteristics as long as the waterline shape is
optimized with respect to that particular section shape.
- The waterline optimization has definite
effect. This process greatly improves the resistance
characteristics, that is, ship resistance is greatly
reduced by the waterline optimization. The waterlineoptimized hull forms produce much lower waves,
possibly due to the favorable interaction between bow
waves and stern waves.
- In order to further improve the resistance
characteristics, it is necessary to optimize the main hull
form characteristics. In fact, the optimization of the
main hull form characteristics is more important than
that of the waterline shape. To achieve the greatest
possible improvements in the resistance characteristics,
it is concluded that the overall optimization, that is, the
simultaneous optimization of the waterline shape and
the main hull form characteristics should be performed.
- The established hull form design method has the
following practical advantages:
i

The optimum hull form configuration could


easily be derived by the overall optimization
process.
Fuel-economic hull form with superior resistance
and propulsion characteristics can be prepared
within an extremely short time.

The prepared hull form always has consistent


performance characteristics for the various
different design conditions.

This R & D project has been completed successfully


and the prepared hull form design program is now
being actively utilized as the efficient tool for the fueleconomic hull form design of fine higher-speed ships.

REFERENCES

Min, K-S. and Hong, S-K., Systematic Study for


New Ship Series, The Proceedings of the Second
International Symposium on Practical Design in
Shipbuilding(PRADS '83), Tokyo and Seoul, OCT. 17
22, 1983.
Min, K-S. and Kim, K-J., A Study on the Optimum
Hull Form Design Based on the Minimum Resistance
Theory, Journal of the Society of Naval Architects of
Korea, Vol. 21, 1984.
Min, K-S., A Study on the Effect of Ship Speed and
Fullness on the Optimum Waterline Shape , The
Proceedings of the Third International Symposium on
Practical Design of Ships and Mobile Units(PRADS
'87), Trondheim, Norway, 1987.
Min, K-S., Experimental Study on the Optimum
Ratio of Entrance and Run Parts according to Ship
Speed for Full Slow-Speed Ships, The Proceedings
of the Fifth International Symposium on Practical
Design of Ships and Mobil Units(PRADS '92),
Newcastle upon Tyne, England, 1992.
Min, K-S. and Kang, S-H., Systematic Study on the
Hull Form Design and Resistance Prediction of
Displacement-Type Super-High-Speed Ships ,
Journal of Marine Science and Technology, Vol. 3, No.
2, The Society of Naval Architects of Japan, 1998.
Min, K-S., Lee, Y-S., Kang, S-H. and Han., B-W.,
Study on the Fuel-Economic Hull Form Design
Method for Fine Higher-Speed Ships , The
Proceedings of the Eighth International Marine Design
Conference (IMDC03), Athens, Greece, 2003.
Wehausen, J.V., Unpublished
University of California, Berkeley.

Copyright National Academy of Sciences. All rights reserved.

Lecture

Note,

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Stefano Brizzolara
University of Genoa, Italy
The method you use to obtain the final body
plan from the actual body plan is not clear to me. In
particular, the method used to pass from several
optimized base waves (WL, tg at bottom, etc.) rely on
designers experience and decision and in this sense
could in principle invalidate the trends found by the
parametrical study.
AUTHORS REPLY
In the fundamental studies, we have
discussed that variation of the underwater section
shape, that is, the body plan within the practical range
has almost no effect on the resistance characteristics
as long as the design load waterline (DLWL) is
optimized with respect to that particular section
shape. This is the proved fact theoretically and
experimentally.
In practice, therefore, we first obtain the
optimum design load waterline (DLWL) with the
initial section shape through the discussed
procedure. The section shape may be slightly
modified for the final shape-particularly at the
maximum section area.
However, original
characteristics of the initial section such as section
area, and hence, the form exponent are maintained
the same in average sense.
As well known, the maximum section for
commercial ships is generated with the bilge radius
and about 2 neighboring sections are modified
accordingly. However, the original characteristics
are maintained for the most of parts.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Hoyte C. Raven
MARIN, Netherlands
As you use thin-ship theory to predict the
wave resistance, can you explain how you deal with
bulbous bows? At the bulb, the assumptions on
which this theory is based break down, and source
strength may go to infinity.
AUTHORS REPLY
I do not understand the meaning of the
question exactly, because such a problem could be
easily removed by a proper manipulation of the
equation.
We have mainly utilized the residual
resistance equation expressed as follows:
RR =

4 g 2
v2

dxdz d d f
s

du cosh 2 u e v

( z + ) cosh u
2

( x, z ) f ( , )
g

cos 2 ( x ) cosh u
v

However, the above expression may give


divergent integrals for some kind of bulbous bows,
submarines or swath ships. If the above equation is
integrated once by parts with respect to x and apply
the boundary condition, then another expression for
the residual resistance equation could be obtained as
follows:
RR =

4 g 4
v6

dxdz d d f ( x, z ) f ( , )
s

du cosh u e v
0

( z + ) cosh 2 u

cos 2 ( x ) cosh u
v

The above expression could be well utilized


for the calculation of the residual resistance for any
kind of hull geometry - bulbous bows, submarines,
swath-ships, etc. Only the matter is the accuracy, not
the calculation itself.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

!"# $ %

()

*
/-

-) $

+) ,
,) - .
)
0
.
1
2
/(
1
3) 4 5 /6!(
1
2

&&'

/ 7$-$A 2
8

7
%

4$

) 7

%
(

8
% ) 7

9
9

)
5

:
)

8
:

%
8
8

) 7
%

)1
8

*
5

:
)

5
5

%
8

%
!

7
8

;B@ (C %
8
%=
5

)1

:
8

1,)

! "
7

%
*

<

% )(%
-

>

8
%

%
>

8 F) ;

%
)
%
) =
!

8
!

) $
- 8
%

%) 7

%
%

:
5
(
; $4(
(4;(

4$

) 4

8
8
9

%
)7

%
%

>
%

8
#,
/@4( $ 2

) <

% %
)7

8 %

%
%
%

8
%

Copyright National Academy of Sciences. All rights reserved.

%
)

Twenty-Fifth Symposium on Naval Hydrodynamics

@))2

>

%
%

/
%

%
%

*
%

%
%

:
8

%>

B
!

;
z

%>
#

>
$

A
z7
z
z

>

A :
z7
z

>
%

1%

"> A

8
! # '

9 %

% &&&

>
%

1%

)))

>

>> ,C

%
%
/

8
%

2)

>

zzG
z

)))

,H

?
)))

%
8

%% !

%%

' )

%
8

9
)

1%
,3

#>-

1%

' > B
/

>

2)

7
5

$,

%
/
9
%
?2 7
%
/- .
&&#2)

!
/-$7$A 2

- 8
%
9

>
? 7

7
%
%
2 /1 %

5
)
B
!

>
2)

%
5

%
/7

; (B
2)

);
5
) 1%

1%

8
> ,C

,3

,H
)
8

1%
>

'
1%

> (5

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

1
; (B

)$

4
9

%
9

/[ " [

)@

% /

7 2

*
%

);
%

#
/ 2

/ 2
/ 7 #2

% %

8
%

[ #2
) 1

)1

$
%
%
/5" 5 5#2

(*

>

/

"
7

2 3

3

,
,

&

%:
%
%
) ;

/%

*
B
1%

I>G

>

&
*

<
;

%
4
9
>

1%



/Q


Y

/Q

J > G
:
;

'

'

2  /Q  V MQ 2

/

2  /Q  VQ 2

 V MQ 2
 VQ 2

 @  E MY  &

%
%
) /$

%
9!Z
$ ) "KK' "KK 2

9 %

Q J M

 JY @ L  EY  &

) 1

/

N V &) N J
E
'&
K

E M &)&K N V M &) N J M "

(
B ,
! 9

)7

>

%*

; $4(
% 1
$
%

5 /"K 2

Copyright National Academy of Sciences. All rights reserved.

V MQ
VQ

2
2Y

Twenty-Fifth Symposium on Naval Hydrodynamics

*
'

*
%

1
9

%
)

D
/ = 7$- P ; D B

%) 7
%
5
%

>

);


2/

2^"

`

&

) 7

2
/ 2

""

""
&

&

%
) @
>

UQ

%
/D2
/@2

!
8

2
5

/

)
%

# #

&

""

""

&

""
""

""

&


Q

u


""

&
&
&

""

7
7
= 7$-

J
/
2
7 %

)
%
>

%
) 1
= 7$9 %
/
%

%
8

!
) ;

%
)

' )

B
2

) ,

2
-;! = 7$! "

%) 7
%

&

> ; $4( /
2 ; (B
@4( $ / 9
%2 7$-$A /
A
:
4
9
%

9
%
%

) @

)
:
) ,

)
%
"#
),

%
5
/,A(2) 7

) 7
%

J
5

"

"KKK2

O,

8
%

=
=

%
O,

) 7

,
9

% A (5
%

:
5

/$

>,

) =
) 7

=
%

:
;B@ (C

)
%

:
*

%
"KKK2 /$

%2
?

5
) (
%
O=

)7
;B@ (C %

%
)

/$
8 &&&2

Copyright National Academy of Sciences. All rights reserved.

5
=

Twenty-Fifth Symposium on Naval Hydrodynamics

- ! %

'

%%

()

; $4(

"&

5
8
9

/01

.
%
9

&
&)&&

&)"&

&) &

&)#&

&)'&

&) &

&)I&

) ;
!

) <
5

;
(5 )

!"&

!"

'

1%

) 1

"& > 7
9)

; $4(
(4;() 1 %
K

%
7

)
) 7

%
) (
)

"'

1%

>

/ % 2

"

(5 )

"&

/21

&
&I
&'

)
;
(5 )

&

&&
&"
!&

& &

&

& #&
'

1%

&#

& '&

&'

& &

"" >
)

&)&&

&)"&

&) &

&)#&

'

1%

K >
/

&)'&

/*

&)I&

"'

/4$
/

&) &

(5 )

"

"&

4$

/21

&

<
(

&'

% 1
5

) 1%

4$ (
9 %) 7

4$

&)

& &

&

& #&
'

1
5

% 1

&&
&"
!&

9) 7

&

"&

%
%

&I

1%

" >
)

&)#&)

Copyright National Academy of Sciences. All rights reserved.

&#
,

& '&

&'

& &

Twenty-Fifth Symposium on Naval Hydrodynamics

7
$
7

9 %

*
9

9) 1 %
"
%
)7
5

) 7
9)
8

J)

K)# 9
)

1%

"# >

1%
%

"/

/ % 2)

"

)7

"'
$
#&& &&&
5
"&
@
)

%
1%
/

"I > 1
2

5
%

)$

(5 )

%
5
9 %

9
8

" /

%
)

"&

""

# /3

1%

"

"#

"'

"

(!

"' >

4$
9

5
$

; $4(
)

%
; $4(
*

@D
1

/01

) <

; $4(
#&& &&&

"&

'

&

@D

"&

""

"

"#

"'

"

4$

(
%

!'

!I
!

F
8

)<

%
"&
%

!"&

# /3

1%

" > 7
$

; $4(

)7
8
%

4$
%
@D

(
) 7

9)

);

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

% ) <
%
&&

&
)7
%

@D
)A @

")

"I

)
%

"
'

& &
%
1%
"K2

)
=

)7

% /1 %

) <

%
*
%

%) 7
/1 %
"
&
1%
"

%
:
)<

%
%

) ;
8
%

/ &
<

5)
4

9 %

&)&'
&)&#
&)&#I

>

&)&#'
&)&#

 " #

 " #

&)&#

&)&

)&
) &
) &

&)& I
&)& '
# &

,

&

>

 " #

<

%
%

 "  , 

,

%
2)

% *

%
% "&&
%

'&&

' &

&&

&

I&&

;7

!
1%

!
%

" >@
8

" 9
%

7
%
)
%
@

&)&

&)&'

&)&#
)&
) &
) &

&)&

# &

'&&

' &

&&

&

I&&

I &

;7

1%
1%
&

"J > G
)

"K > @
8

Copyright National Academy of Sciences. All rights reserved.

9
%

"2

Twenty-Fifth Symposium on Naval Hydrodynamics

1%

&

1%

"

)1
%
)<

%
"&&
%

%
&
&
% ) 1

*
%
%

&)"

&)"'

1%

>

) ;
)

&)"'

)&
) &
) &

&)"#

&)"#
# &

'&&

' &

''&

'I&

' &

&&

&

'&

I&

;7

1%

&>G
8

" 9
%

&)"

1%

&)"

#>

)A
%

&

&

&)"'

&)"'

&)"#

)&
) &
) &

&)"#

# &

'&&

' &

&&

&

I&&

;7

1%

">G
8

9
%

;
% %

%
/1 %

&
/1 %

#2
% /1 %

%
&
'2) 7

1%

'>
&

) 7
%
% )

Copyright National Academy of Sciences. All rights reserved.

)A
%)
%

Twenty-Fifth Symposium on Naval Hydrodynamics

8
8

9) $

9 %
%

8
%

J 9
)7

"

%
5

1%

>

1%
J>7
" 9
)

C 3

1%

C 3

8
>C

H
8

%
% )7
%
-$7$A >
)7

3
%
%

>
N

>

R&)&
#)J

5
" J

&)&&
%

R&)&"

&)&&

&)&
")
#)
)
&)&"
!#)J
R")#

J9

>7
)

H
;
@
A

&

/S2 J 9

C
3
I>@
3

>C
H

!'

!I

% )

1%
%

!"&

!"

!"&

)7
%
1%

%
J

1%
5

1%
,
@

K > @
>
)
)

)
*

&

/S2 " 9

8
) ,

>
>

Copyright National Academy of Sciences. All rights reserved.

)
>

Twenty-Fifth Symposium on Naval Hydrodynamics

@
"'&

/1 %

%
:

/S2 J 9

%
%

5
%
1%

1%

8
%

5
/BA=$2

/ ;B@ (C2

D
=

&

)
%

:
K2) $

!"&

)
!"&

&

! % -

;B@ (C

%
) ;
5 /%
2) $

5
%
5

1%

%
J9

)/
7

#" > A

;B@ (C
2

/*

%
%

2)

"KI 2

;B@ (C

@
%

/"&

2)7
)

)7

%
:

"
/1 %

"&

/S2 " 9

;B@ (C
%

/1 %

>

#&2

J 9

#"2)

/S2 " 9

"&

&

!"&

&

/S2 J 9

&

&

"&&

" &

;,
!"&

!"&

1%

#& > A

&

8
/

/S2 " 9

%
" 9
2)

1%
# >
" 9
,A( / *

"&

/*

;B@ (C
2

%
:
2

%
%
;B@ (C /

%
%

Copyright National Academy of Sciences. All rights reserved.

J9

Twenty-Fifth Symposium on Naval Hydrodynamics

1%
#'
BA=$ %
%
"

8
:
)<

J9

/S2 J 9

@
1%

&

#
%

/*

& %

) 7

&

"&&

" &

;,

%
:
2

%
%
;B@ (C /

J9

&

2
7

1%
## >
J9
,A( / *

&

/S2 J 9

!"&

!"&

A :

%
1
/BA=$2
:
) 7
%
5
%
"&
)7

) B

$%

@D
)$
%

!"&

B
9

1%
%
/*

"&

/S2 " 9

&
) ;

BA=$
%
%

/S2 J 9

/S2 J 9

&

> (

1%
)

#I
8
%
%

%
) ;
%
)

&

!
!

&

!"&

!"&

!"&

&

1%
A :

#' > 4
=
%

$%
/

"&

&

/S2 " 9

8
/BA=$2 / *
2)

&

"&&

" &

&&

&

#&&

;,

%
2

1%
J9

#I >
%
) ,A( / *

Copyright National Academy of Sciences. All rights reserved.

BA=$

%
/

Twenty-Fifth Symposium on Naval Hydrodynamics

@
%

5
.
1%

"&S

"
-I&&
G
4 ) 7
#J
1%
# )

J9
)7
9 -

(5 )

(5 )

)
8

&&

&"

&"

&

&

'

1%

&#

&#

&'

1%
/

#K > <
2 " 9

1%
/

'& > <


2
J9

&'

#J > 7
8

" S

/01

&S

"&S

(5 )
S

'

<
8
.

&S
&

&"

&

&#

'

&'

&

&I

) 7

:
1%

> 7
/

1%

#K

) $

/*

1%

/S2
2

'&

>

) (
; $4(
9

%
)
%

8
8

) 7

"

%
%

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

8
:

5
)

#&S
)

%
/01

%
) 7

)1%

'"

(5 )

&S
#

" S
"&S
S
&S
&

"&

"

&
# /3

1%

'# > 7

/S2

9/ *

1%

#&

'" > (5
)

8
%

" 9

IS

)
9/
%

%
8
" 9

2
) S

1%

5
9

&S

'

%
; $4( /
2) 7

9/*

5
9 %
/1 %

'#2)

1%

$ " 9

5
%

&S
9

S) $

'' > <

%
%

) S
S

1,
8
)7

5
;
;

)7

(5 )

9 %
%

)
(5 )

8
%
4$
)B
8

%
) <

(
4$

9
% )
7
"&

"

&

# /3

1%

' >7
/*
2

#&

9
/

%
:

2)

%
*

Copyright National Academy of Sciences. All rights reserved.

&& 2)

Twenty-Fifth Symposium on Naval Hydrodynamics

*
E

%
F

,) ) <
' ''

5 EB
G

$) -

F
"K )

) " ""!"# &

*
%

EB
% E ! (
) 'K R I"

%
) ""

V'

!A :
$

8
*445/

)
,) @
.

)%

( !

@ %

%
%
%
8

)7

,) @
$

"&

) #

VI

E
F '
) ##J R #

E, %
* F
) "'& R "'K

()
1,
)'

F
)"!"

B)G

8
$

%
5

'

=) ,

V"

5 1)

B) 4

) 'J

,)

# *44

4) ,

-) $

# *445

8
# *44

8
*

EB

)@

2
8
@($K ) &)&"
% ,

()
8
$

8
" #
"
)" JR" J

1, F
$

8 )< %
8
)#
V#

E, %

$ *446

) ''

V ""

778/

!
7)

F
! " #
"KK'

-) $

"K

=) ,
!

# "!#'

! "
K

"

-) $

=) ,

%
&

-) $
G
- $

)$)

# *446

)'

F @
(

FG

**
&" $ " ! (
" #$ <
%
T "KK
)=
1
=U

8 E$ 1
1

7
F +%,
&&&

()
*
,) -) $
E3
A
8
=
% 4$ (
./
"
, $
+
# $ 9
,
,) ("&' &
1

F
%

"KK )

=) ,

E1

F1

&" $ "
% # $ '#
) $$ G
4
-) $

4)
8

!"' "KKK

&

B) @
%

F$

EB
. " 1 $$
&&#)
B
8

2 $$ $

4) E $
F "KI

)
$%
# 0 #
&& )

-$7$A F
$# 3 $

Copyright National Academy of Sciences. All rights reserved.

V"

F
) I" ! JI

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Emilio F. Campana
Istituto Nazionale per Studi ed Esperienze di
Architettura Navale, Italy

Samareh JA,"A Survey of Parametrization


techniques", NASA/CP-1999-209136, June 1999, pp
333-343, NASA Langley Research Center, Hampton,
VA 23681
AUTHORS RESPONSE

The applications presented in this paper


offer a good example of the use of a high-fidelity
RANSE solver as a solid basis for an accurate
analysis and design. Promising results should
encourage the Authors to certify the obtained
improvements performing a convergence study of the
numerical solver used in the analysis and the
uncertainty analysis of the experimental data used for
verification. A few points regarding the approach
followed in the paper may be clarified by the authors:
1. Geometrical and functional constraints play a
major role in real-life design avoiding
meaningless solutions. May the Authors give
some detail about the enforced constraints? How
have been implemented the necessary
penalization techniques?
2. Pareto front reported in fig. 29 seems to be
conflicting with the definition adopted in
literature (e.g. Miettinen, 1999). Most of the
solutions indicated as Pareto optimal in this
figure are dominated by others and apparently
only 3 solutions belong to the Pareto optimal set
(those below (10%, 10%)). Can the Authors
clarify the definition of Pareto front they are
using?
3. Is the DOE phase always used as a starting
procedure to sample the design space before
simplex or moga algorithms are applied? In
figures 32, 33 and 36 the optimal solutions seem
to be found during the DOE phase and the value
added by the numerical optimization algorithms
is not evident.
4. In the optimization community it is known that
the Nelder & Mead simplex method proved to be
slow in convergence, and sometime it is not
convergent at all (Dennis and Torczon, 1991).
Do the Authors have experienced low
convergence rate in the use of this approach?
REFERENCES
Dennis JE Jr. and Torczon V, "Direct search methods
on parallel machines", SIAM Journal on
Optimization, Vol. 1, No. 4, pages 448-474,
November 1991.
Miettinen
K,
"Nonlinear
Multiobjective
Optimization", Kluwer Academic Publishers, Boston,
1999

The authors would like first to thanks Pr.


Campana for the interesting questions and remarks.
During the development of the optimization loop in
Bassin d'essais des carenes, we have tested several
solvers, with convergence study, and systematic
comparisons with experimental data available from
tank tests. The results are not presented in this paper,
but can be found in Alessandrinis publications. For
industrial studies the choice of the mesh size is very
important because it is a balance between the quality
of the results and the time you need to evaluate about
200 hulls.
1- The examples presented in this paper have been
carried out with naval shipyard DCN (French
Military Shipyard). The constraints are given by
the naval architects in charge of the project, and
are directly integrated in the modeling software.
We dont use penalization functions because the
hull that are generated with the modeling
software are respecting the constraints, by
construction. The constraints that are generally
used are of course Lwl and displacement, but
also 3D sonar dimensions for the bulb, engine
block dimension, ship stability criterion
2- The definition of Pareto frontier is the classical
one used in the literature. Fig 29 presents a
mistake in the selection of some points.
3- The paper presents a comparison of several
methods: DOE, SIMPLEX and MOGA. They are
not performed sequentially for industrial project.
The use of optimization algorithms such
SIMPLEX can be dangerous because you dont
know if you reach the global optimum. The use
of MOGA is very time consuming, because of a
relatively low convergence rate. We mainly use
DOE in order to give a representation of
hydrodynamic response surface (3d plots
presented in the paper). This can be efficient
only if the number of parameters is limited to 3
or 4. This is the choice we made with the use of
a specific modeling software, compared to the
modification of poles that lead to an important
number of parameters.
4- According to our experience, SIMPLEX
algorithm converges very quickly, and is more
robust than Gradient based methods. We dont
have any bad experience with this algorithm,

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

perhaps due to the fact that the number of


parameters is always limited to 3 or 4.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

DISCUSSION
Hoyte C. Raven
MARIN, Netherlands
This paper shows some examples of hull
design optimisation and ranking using a free-surface
RANS solver. Clearly, CFD-based optimisation is a
desirable new possibility, to be introduced in
practical ship design after further consolidation. Our
experiences (e.g. in the FANTASTIC project)
suggest that the main conditions for successful CFDbased optimisation currently are:
1. Effective parametrisation of the hull form;
2. Accurate prediction of the trends by the
CFD-code.
The paper mainly addresses the second issue. The
examples show a correct ranking of hull form
variants, although the magnitude of the differences
often is underestimated. In our own studies we found
that the predicted dependence of the viscous
resistance on hull form parameters differed
significantly for different grid densities; and quite
dense grids were needed to get accurate trends. Did
you check such grid dependence for your code?
The examples shown in the paper all refer to
fairly slender ships at higher Froude numbers, and are
probably dominated by wave resistance differences.
Do you have experience with similar optimisations or
rankings for fuller ships at lower Froude numbers,
when viscous drag dominates?
AUTHORS RESPONSE
The Authors thank Hoyte Raven for his
interesting remarks. We agree with the two main
conditions for successful optimization. Its true that
the paper mainly presents the ability of the solver to
rank hulls. These remarks encourage the authors to
publish description of the modeling software, perhaps
in a more CAD/ Naval design oriented conference.
We have some experience with hull form
optimization of Aircraft-carrier or Ferry boat, with
Froude numbers around 0.2. At this speed, even if
viscous drag dominate, the variation of the pressure
drag is always more important than the viscous drag
modification. For example, we found that pressure
drag could be modified from 30 % to +50%
(depending of the initial design), whereas
modifications of the viscous drag is always lower
than 5%. This is based on the results of few
optimizations, and must be confirmed on new
projects.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics


St. Johns, Newfoundland and Labrador, CANADA, 8-13 August 2004

A PhysicsBased Simulation Methodology for


Predicting Hydrofoil Singing
Eric G. Paterson, John E. Poremba, Leonard J. Peltier,
and Stephen A. Hambric
(The Pennsylvania State University, University Park, PA, USA)
ABSTRACT

NOMENCLATURE

A physicsbased simulation methodology for predicting hydrofoil singing has been developed. Two
model problems, each of which focuses on specific
aspects of the methodology, namely fully-coupled
fluid dynamics and structural acoustics, and hybrid
RANS/LES simulation for a practical propulsor geometry, have been studied.
For the former, a singing symmetricallybeveled
cantilevered strut was studied to establish the feasibility of coupling unsteady CFD and structural
acoustics simulations. The uncoupled shedding frequency was shown to agree well with experiment,
and linear response of the strut was shown to be
large, but only at a very specific speed which corresponds exactly to coincidence of shedding and modal
natural frequencies. Fullycoupled simulations, on
the other hand, showed a much broader and larger
response indicating lockin of nearresonant conditions. Impact of resonance on the flow field was
shown to be significant with large increase in oscillatory pressure over the strut, reorganization of the
vortex street, and generation of travelingwaves in
the boundary layer due to unsteady changes in angle
of attack.
DES for a propulsor stator blade was undertaken to study the unsteady flow field generated by
trailingedge vortex shedding. Results were shown
to be similar to previous DES work, however, this
case illustrates the effect of angle of attack and camber which were not present in earlier work. Spectra of both the unsteady forces and trailingedge
pressure show 3 distinct peaks corresponding to the
vortexshedding frequency and two super harmonics, the latter of which is directly related to unsteady
drag. All 3 peaks are potentially close to the resonant response of the the circumferential harmonic
of the stator row 1stbending mode, and thus motivates future work in coupling the DES and structural
acoustics for this problem.

b
CD
CL
Cp
D
Di
f
k
L
m
p
p0
Rb
Re
S
St
t
e
t
Ui
U0
xi

Damping coefficient
(kg/s)
D
Drag coefficient = 1 U
2
2
0S

L
Lift coefficient = 1 U 2 S
2 0

0
Pressure coefficient = pp
1
2
2 U0
Drag force (N)
Surface displacement
frequency (Hz)
(1) Turbulent kinetic energy
(2) Stiffness coefficient (kg/s2 )
(1) Chord length (m)
(2) Lift force (N)
Mass (kg)
Pressure (N/m2 )
Reference pressure (N/m2 )
Body radius (m)

Reynolds number = U0 L
2
Wetted surface area
(m )

Strouhal number = fUL0


Time (s)

Nondimensional time = UL0 t


Velocity components (= U, V, W )
Reference velocity
Cartesian coordinates (= x, y, z)
Angle of attack, deg
Time step
Dynamic viscosity
(1) Turbulent dissipation rate
(2) circular frequency
2f
)
(=
p
Natural frequency = k/m
Density

INTRODUCTION
Design of small propulsors for highspeed vehicles presents unique challenges, particularly when
new materials or fabrication techniques are em1

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

ployed. While the nondimensional trailingedge


vortex shedding frequency, or Strouhal number, for
a given blade geometry may be fairly insensitive to
Reynolds number (Re), the dimensional frequency
can potentially be on the order of several thousand Hertz due to high velocity and small length
scales. Coupled with the modes of vibration of
lightly damped blades, generation of acoustic tones
and structural fatigue failure could potentially occur
if the shedding frequency were to coincide with any
of the modal resonance frequencies.
It has long been known that spatially coherent
vortices shed from bluff bodies sometimes cause excessive vibrations in underlying structures and high
sound pressure levels in surrounding acoustic media. When the frequency of the vortex shedding is
nearly coincident with the frequency of a receptive
structural or acoustic resonance, the fluctuating velocities of either the acoustic fluid or structure can
reinforce the strength of the shed vortices, resulting
in a feedback loop between vortices and structuralacoustic vibration. In extreme cases, the vortices
and structural-acoustic modes couple and lock-in to
each other, often with disastrous results. The flowvibration behavior leading to the 1940 collapse of
the Tacoma-Narrows bridge is an example of this
phenomenon.
When lock-in occurs in hydrofoils and marine
propellers, it is commonly termed singing, since
the frequency of the shedding and corresponding
structuralacoustic resonance is often in the auditory frequency range of humans1 . Ross [1987] provides a history of early investigations of singing from
ship propellers. Once vortex shedding was established as the forcing function responsible for exciting propeller blades, and the discovery that singing
always occurred at propeller blade resonance frequencies was made, several strategies for eliminating singing were devised. To shift the shedding frequency away from propeller resonance frequencies
and to reduce the spatial coherence of the shed vortices, various modifications to blade trailing edges
were attempted, usually making the edges asymmetric about the blade camber lines. Also, as documented in the literature, e.g., Blake et al. [1977],
lightly damped propellers were found to be more
likely to sing than highly damped ones, thus leading to material selection aimed at reducing vibratory
response at resonance.
More recently, work has focused on the detailed
physics of the singing phenomenon, motivated by

the desire to use computational techniques during


the hydrofoil design process to ensure that singing
would not occur during operation [Blake, 1984, Ilin
and Levkovskii, 1993, Matveev, 2002]. Fortunately,
the coupling between shed vortices and structuralacoustic resonance had already been studied, but for
other applications. Examples include the lock-in of
vortices shed from the leading edge of a cavity with
acoustic resonances within the cavity [Rockwell and
Naudascher, 1978], and the lock-in of vortices shed
from flow over cylinders with structural resonances
of the cylinders [Skop and Griffin, 1973].
Many investigators of lock-in phenomena make
use of the simple liftoscillator model proposed by
Hartlen and Currie [1970] to simulate the nonlinear
coupling of the unsteady lift with the vibration. The
model employs a modified Van der Pol equation for
the fluctuating lift and the usual dynamic equations
of motion for the body. While it has been used to
simulate the likelihood of lock-in, or singing in hydrofoils [Blake et al., 1977], it is not a predictive
tool since there are several coefficients that depend
on the specifics of the flow field, and which must be
tuned using experimental measurements.
Direct simulation of vortexinduced vibration
(VIV), and fluidstructure interaction in general,
has been limited to simple geometries, such as cylinders and cables, and low Re (e.g., Evangelinos and
Karniadakis [1999], Lucor et al. [2002]). While these
simulations are impressive in scope due to the largeamplitude motions and direct numerical simulation
(DNS) resolution of turbulence, they lack both the
geometric complexity of modern marine propulsors
and the highfrequency vortex shedding in a fully
turbulent wake at ReL > 105 .
To undertake such simulations,
a flow
solver/model which can resolve the largest and
most energetic turbulent eddies and which can
resolve complex geometries, including surface
displacements, is required. While large-eddy simulation (LES) has the ability to resolve small-scale
inertial-range eddies in the turbulent boundary
layer, resolution of vortex shedding requires that
the entire geometry be modeled so that the unsteady
circulation and lift can be captured. Moreover,
Blake [1984] has shown that vortex shedding,
and particularly VIV, has a strong influence on
the spanwise integral length scales of pressure.
Minimally, when using periodic spanwise boundary
conditions, the size of the domain in this direction
should be approximately 510 blade thicknesses. To
our knowledge, both of these requirements have yet
to be achieved in stateoftheart blade LES, e.g.,
[Wang and Moin, 2000].

1 A related phenomenon is flutter of airfoils and airplane


wings, the frequencies of which are usually very low, outside
the auditory frequency range of humans

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Therefore, the objective of our work is to develop a physicsbased simulation tool which can predict hydrofoil singing and which eventually can be
used within a typical design cycle. Our approach
is to extend recent work by Paterson and Peltier
[2004], on hybrid Reynoldsaveraged Navier-Stokes
(RANS)/LES detachededdy simulation, in particular for trailingedge flows, by coupling the
method with a structural dynamics algorithm and
with the existing dynamic gridding capability of
CFDSHIP-IOWA [Paterson et al., 2003]. In the following sections, we will present the model problems
and flow conditions, and briefly summarize the computational approach. Grid generation and computational parameters will be discussed, followed by an
analysis of the simulation results. Finally, a summary and list of future work will be provided.

w = 11.7 cm
2.54cm

h = 0.635cm
t = 0.18cm

Figure 1: SymmetricallyBeveled Strut Geometry

The loss factors of the strut were measured at varying flow speeds. Below the shedding frequency, it
is about 0.002, which is typical of lightly damped
metals like brass. To use structural damping in a
forced transient vibration analysis (see later section
on structural-acoustic modeling), it must be converted to an equivalent viscous damping constant at
the analysis frequency, so that bn = kn / = 6.71
kg/s at 163 Hz.

MODEL PROBLEMS AND FLOW


CONDITIONS
Singing SymmetricallyBeveled Strut
The first model problem in our study was part of an
experimental and analytical investigation by Blake
et al. [1977] into the effects of trailingedge thickness
and structural damping on flowexcited vibration of
simple cantilever hydrofoils. Blakes type IIIB strut,
which is shown in Figure 1, is made of brass and
has a chord length L=11.7 cm, span length of 50.8
cm, and a thickness h=0.635 cm. The strut trailing
edge is a symmetric bevel of 12.5 degrees originating
from a blunt end of thickness t=0.18 cm. Blake et al.
[1977] report several standard bending and torsional
modes of the strut, which may be approximated as
those of a cantilever plate. Leissa [1993] summarizes
several semi-analytical solutions for the resonance
frequencies and mode shapes of a cantilever plate,
which agree reasonably well with those reported by
Blake et al. [1977]. In particular, Blake et al. [1977]
focused on the third cantilever mode of the plate,
where acceleration levels indicated a strong singing,
or lock-in phenomenon, at a free stream velocity of
2.28 m/s, corresponding to ReL = 2.67x105 . Based
on the information above and the formula provided
by Leissa [1993], a modal mass and stiffness of 1.604
kg and 3.44x106 kg/s are estimated. Since measurements were made with the strut immersed in water, an entrained fluid mass of 52% of the structural
mass is estimated as 1.67 kg [Ross, 1987], giving an
inwater modal resonance frequency of
s
kn
1
= 163Hz.
(1)
fn =
2 mn + mf luid

Propulsor StatorBlade Analysis


The second model problem of our study is based
upon an ARL Penn State propulsor which was machined out of a solid block of aluminum. A notional
representation of this machine is shown in Figure
2. It consists of upstream duct supports, rotor, and
postswirl stators. Operating conditions correspond
to Re = U0Rb = 2.1x106 where Rb is the body radius
and U0 is the vehicle speed.
In addition to geometry, Figure 2 shows pressure
contours from a poweringiteration analysis, a procedure which will be briefly discussed in the next
section, and typical first bendingmode resonant response of the aft stator blades. The nondimensional
b
frequency of this mode is about
e = R
U0 = 115,
which is potentially close to estimates of the shedding frequency; thus motivating much of the work
herein.
To simplify the analysis, we have restricted our focus to the midspan section of the stator blade. The
geometry at this radius was extracted, projected to
a plane, and then extruded in the spanwise direction
so that a periodic detachededdy simulation (DES)
could be undertaken. Inflow angle of attack was iteratively adjusted so that the modeled section loading
matched the poweringiteration analysis. This was
achieved by setting the blade section to an angle
ofattack = 13.7 .
3

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

COMPUTATIONAL APPROACH
CHAMP
ARL/Penn State uses a suite of hydrodynamic
and structural acoustic tools, integrated under the
acronym CHAMP (Combined Hydro-acoustic Modeling Programs) [Hambric et al., 2004], to assess
the vibrations and radiated sound characteristics
of operating marine propulsors and other fluid
dynamic machinery. CHAMP separately models
flow-induced sources due to turbulence ingestion
(TI) [Lysak, 2001, Lysak and Brungart, 2003], turbulent boundary-layer (TBL) forcing [Peltier and
Hambric, 2004], and trailing-edge vortex shedding.
Fluctuating pressure fields acting on the surfaces of
the propulsor are estimated using CFD powering
iteration solutions as inputs to the TI and TBL
tools. The structural-acoustic response characteristics of the components are computed using Finite Element (FE) and Boundary Element (BE)
software[Koopman and Fahnline, 1996, Hambric
et al., 2003]. Typically, the pressure fluctuation
and structural-acoustic computations are uncoupled,
such that the pressure fields are computed assuming a rigid underlying structure (blocked pressures).
While this is acceptable for broadband turbulent
boundarylayer forcing, it is clearly unacceptable for
stronglycoupled problems such as the singing phenomenon described earlier. As such, it represents
one of the weaknesses of the current design/analysis
procedure and provides motivation for costly water
tunnel testing of all designs.
Central to the entire procedure is the CFD
poweringiteration analysis, the purpose of which
is to predict the steady designpoint powering of
a propulsor using highfidelity RANS simulations.
The approach couples axisymmetric RANS simulations of the entire vehicle with body forces derived
from detailed three-dimensional RANS simulations
of each blade row. This approach eliminates many
of the ambiguities, such as definition of the effective
wake, which are often encountered when using an
inviscid propeller performance code to obtain body
forces. In addition, detailed resolution of boundary
layers, wakes, and tip vortices provides critical information for both estimation of structural acoustic
forcing functions and cavitation physics. The only
external inputs to a powering iteration are the vehicle speed and shaft rpm, so results from it may
be used in the same way as results from a water
tunnel test. Currently, highly customized versions
(UNCLE-TF [through-flow] and UNCLE-REL [relative frame]) of the UNCLE flow solver are used to

(a) Surface pressure from poweringiteration analysis

(b) Typical first bendingmode response of stator


blades

Figure 2: Three-dimensional view of notional


propulsor geometry and components of CHAMP
analysis

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

perform the powering iteration, e.g., Boger et al.


[2003].

0.1

conv
0.90
0.80
0.70
0.60
0.50
0.40
0.30
0.20
0.10
0.00

RANS Region

Flow Solver
Y

0.05

While converged RANS solutions may be used to estimate TI and TBL fluctuating pressure fields with
reasonable accuracy, timeaccurate CFD solutions
are required to simulate the vortexshedding process.
CFDSHIPIOWA [Paterson et al., 2003] serves as
the core flow solver for the unsteady RANS and
DES computations reported here. CFDSHIPIOWA
is a general-purpose parallel unsteady incompressible RANS CFD code. The computational approach
is based upon structured overset-grid, higher-order
finite-difference, and pressure-implicit split-operator
(PISO) numerical methods. Production turbulence
model uses a linear closure and the blended k/k
SST 2-equation model[Menter, 1994]. Efficient
parallel computing is achieved using coarsegrain
parallelism via messagepassing interface (MPI) distributed computing. For timeaccurate unsteady
simulations, global solution of the pressurePoisson
equation is achieved using preconditioned GMRES
and the PETSc libraries. [Balay et al., 2001]
DES [Strelets, 2001] is a threedimensional unsteady numerical method using a single turbulence
model, which functions as a subgrid-scale model in
regions where the grid density is fine enough for a
LES, and as a RANS model in all other regions.
Implementation of DES in CFDSHIP-IOWA was
accomplished by modifying the turbulence model
and convective-term discretization. The turbulence
model is modified by introducing a DES length scale
`e = min (`k , CDES )

"LES" Region
-0.05
3.2

k 1/2
, = max (x , y , z ) .

k 3/2
`k

3.5

Figure 3: Contours of convective scheme blending


function
shifts energy from sub-grid, or modeled, scales to resolvable scales as defined by the filter width CDES .
The second modification aims to reduce numerical
dissipation inherent in the upwind convectiveterm
discretization scheme. The implemented approach is
based upon a hybrid central/upwind approximation
of the convective terms.
Finv = (1 ) Fctr + Fupw

(5)

where is defined as

m
CDES
|U |t
= max tanh
, tanh
lk

(6)
The result is that smoothly transitions between
1.0 in the RANS regions, resulting in a local upwind
scheme, and near 0.0 in the LES regions, where a
centered scheme is desired. In addition, a Courantnumber constraint of 1.0 has been imposed which
requires that time step be sufficiently small to support turbulent eddies. The coefficients n, m permit
the interface between RANS and LES regions to be
arbitrarily sharpened, however, currently we use
n = m = 1 due to the fact that higher-order coefficients have resulted in unstable simulations. It
is also noted that Equation 6 was developed since
implementation of Strelets [2001] function for was
unsuccessful.
Figure 3 shows color contours of for the stator
blade simulation. The red region is interpreted
to be URANS. DES is active in the green/blue region. Note the abrupt transition from RANS to DES
upstream of the bevel on the pressure and suction
sides in response to an abrupt increase in spanwise
grid resolution. A thin-band of red is observed
near the foil surface, indicating that the near-wall
boundary-layer is modeled by URANS, as desired.
In CFDSHIP-IOWA the convective terms are discretized with the following higher-order upwind for-

(2)

(3)

CDES is a model constant with a value between 0.78


and 0.61 weighted by the Menter k /k blending function [Menter, 1994]. The new length scale
`e replaces `k in the the dissipative term of the ktransport equation

k 3/2
.
`e
(4)
The effect of this modification is to increase dissipation in LES regions such that the turbulence budget
k

DRAN
S = k =

3.4

which compares the turbulence length scale from


RANS, `k , to the local grid size, , where is
based on the largest dimension of the local grid cell:
`k =

3.3

k
= DDES
=

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

mula
Ui

Uj
i

1
(Ui + |Ui |) i Uj
2
1
+ (Ui |Ui |) +i Uj .
2

elements (BE). In the case of hydrofoil singing, this


assumption is violated, and the pressure becomes
p(, , Ui [x x0 , t ], t ),and the system of equations becomes coupled:
Z Z Z
Ui (x, t) =
p(, , Ui [x x0 , t ],

(7)

The discrete upwinding operators i and +i are calculated using a 5point stencil
i Uj

= Wmm Uj,i2 + Wm Uj,i1 + Wn Uj,i


+ Wp Uj,i+1 + Wpp Uj,i+2
(8)

+i Uj

= Wpp Uj,i2 Wp Uj,i1 Wn Uj,i


Wm Uj,i+1 Wmm Uj,i+2
(9)

MD(t)
+ BD(t)
+ KD(t) = F(t)

(11)

(12)

where D(t) is a vector of displacements at all locations x, M, B, and K are (typically large) matrices
of mass, damping, and stiffness terms, and F(t) is a
time-varying force vector. A common and computationally efficient approach to solving for structural
vibrations is to express the displacement in terms
of the systems structural-acoustic mode shapes, or
eigenfunctions:

Structural-Acoustic Modeling
The velocity response Ui (x, t) at a point on a structure excited by a fluctuating pressure field p is found
by integrating over the loaded region:
Z Z Z
Ui (x, t) =
p(, , t )hi (x, x0 , ) d d d

Rather than attempting to couple directly the


Navier-Stokes and structuralacousticresponse
equations into a single simulation, we instead solve
Equation 11 iteratively over time, where the CFD
computational domain will be adjusted according
to the instantaneous displacement vector Di (x, t),
and the instantaneous velocities and pressure field
are updated via the flow solver.
When using the finite element method to analyze structural response, the structural geometry is
subdivided into elements and an overall system of
equations is generated based on discretized stiffness,
mass, and damping, where the displacement D is
computed by solving the system of equations:

with the coefficients Wi defined as a weighted sum


of the coefficients wi for the 4th-order central and
the 2nd-order upwind schemes, Wi = (1 ) wi4c +
wi2u .
Finally, as discussed in later sections, CFDSHIPIOWA is an overset-grid capable CFD code with an
interface to PEGASUS 5.1[Suhs et al., 2000]. This
capability will be exploited to perform local grid refinement and flow adaptation in the wake and LES
regions.

t )hi (x, x0 , ) d d d

D(t) = (t)

(10)
where Ui (x, t) is a velocity response in direction i at
time t, x is a vector in space defining the response
location, and is the time delay between t and t0 .
The loaded region is represented by a surface parameterized by curvilinear coordinates and . The
impulse response function hi (x, x0 , t, t0 ) = Ui (x, t)
when p(, , t) = (x x0 )(t t0 ), where denotes the Dirac delta function and x0 is a vectorin
space defining the location at the parametric coordinates (, ). In most cases, it is acceptable
to assume that Ui (x, t) has a negligible effect on
p(, , t ), such that p(, , t ) is estimated assuming blocked boundary conditions, e.g., the fluid
is flowing over a rigid surface. When this assumption is made, p(, , t ) may be computed using
CFD analysis of flow over a stationary grid of points,
and hi (x, x0 , t, t0 ) may be computed using numerical
techniques like finite elements (FE) and boundary

(13)

where is a matrix of mode shape column vectors


defined over all x (the number of columns corresponds to the number of mode shapes included in the
solution), and (t) is a vector of coefficients defining the participation of each of the mode shapes at
time t. The eigenvalues (resonance frequencies) n
and eigenfunctions are computed for each mode n
by solving:
2

n M + K n = 0
(14)
assuming time-harmonic dependence eit .
The structural stiffness, damping, and mass are
converted into modal, or generalized coordinates by
pre-multiplying the system of equations by the mode
shapes:
h
i
+ B(t)

T M(t)
+ K(t) = T F(t) (15)

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

so that:
+ b(t)

m(t)
+ k(t) = f (t)

interpolation stencils is expensive, gridpoint redistribution is confined to the nearblade region so that
the donorreceiver relationships are not disturbed
during coupled simulation. This is accomplished using a tanh blending function to smoothly reduce the
redistribution to zero at some predefined distance
from the wall. It is further noted that the grid velocities x,
y,
z,
resulting from the unsteady redistribution, are also introduced into the no-slip boundary
conditions, i.e., U = x,
V = y,
W = z,
and the convective terms of the momentum equations, i.e.,

(16)

where m, b, and k are the generalized mass, damping and stiffness matrices, and are typically much
smaller than M, B, and K, leading to significant
computational time savings.
For reasonably sturdy structures vibrating in air,
m, b, and k are diagonal, such that the modes are
uncoupled, and each combination of diagonal terms
completely defines the response of each corresponding mode. Structures vibrating in water, however,
encounter significant fluid loading over their surface,
which leads to cross-coupling, and full m, b, and k
matrices.
When the modes are uncoupled and the applied
forces are deterministic, each modes displacement
contribution may be solved for analytically, where:
fn (t)

n (t) = p

(kn mn 2 )2 + (bn )2

Ui
Ui
DUi
=
+ (Uj xj )
.
Dt
t
xj

While simple in principle, there are several practical aspects of the approach to consider prior to its
implementation. Paterson and Peltier [2004] have
shown that to obtain accurate DES solutions, the
time step must be small enough to support turbulent eddies. Because these eddies are much smaller
than the organized vortex shedding, the latter is over
resolved with 100-200 time steps per period. The
finite element transient solution, however, requires
only about 10-20 time steps per shortest frequency
cycle of interest. In either case, the CFD is several ordersofmagnitude more expensive than the
structural acoustics portion of the simulation.
The direct integration approach used in the FE
transient analysis is unconditionally stable. However, the solution at a given time step depends on
the solution at the previous time step. At the beginning of an analysis, the solutions at earlier time
steps are not generally known, and if set to some arbitrary value (zero, perhaps), lead to a discontinuity
in the response. If structural damping is low, which
is often the case for conditions conducive to singing,
and the time periods of some of the modes in the
structural-acoustic system are long, as they are for
low-frequency resonances, then a significant number
of time steps are required for the solution to reach
a steady state condition. This behavior leads us to
make two observations:

(17)

and the phase delay between the displacement and


applied force is computed from:
tan n =

bn
kn m n 2

(19)

(18)

When the modes are cross-coupled, or when the


phasing of the applied force is distributed randomly
over the surface (which is the case for a turbulent
flow field), a numerical time integration algorithm
is required to solve Equation 16, which is a simple
second order Ordinary Differential Equation (ODE).
We employ a 4th order Runge-Kutta algorithm, implemented in the open source library RKSUITE,
finding that it provides solutions in agreement with
Equations 1718 when applying a simple point drive
with known phase angle.
Coupling Flow and Dynamic Equations via
Dynamic Gridding
Given the expected smallamplitude displacements
D(t), the existing dynamicgrid algorithms in
CFDSHIPIOWA, which were originally developed
as part of the freesurface wavetracking capability,
can be directly applied to the singing problem.
The algorithm is based upon a rigid-wire assumption where D(t) is used to move the surface
grid points in the coordinate direction normal to the
noslip surface. The baseline gridpoint distribution
is stored, and used to redistribute the field points at
each time step. Since computation of oversetgrid

Since the CFD computational requirements are


high, and the CFD solution requires many more
time steps per shedding cycle than the FE/BE
solutions do, some provision must be made to
eliminate or mitigate the transient behavior of
the initial FE/BE solution.
A means of accomplishing this is to find an improved initial estimate of the FE/BE solutions
at the time step prior to t=0 (the beginning of
the coupled solution).
7

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

With the above considerations in mind, our computational approach is summarized below.
1. Solve time-accurate CFD problem (using DES
or URANS) with *blocked* boundaries, ignoring dynamic coupling
2. Compute force spectrum from fast Fourier
transform (FFT) of converged, blocked boundary, CFD solution
3. Compute structural-acoustic modal response
due to fluiddynamic forces using Equations 17
18

Figure 4: Overset grid system for Symetrically


Beveled Strut: Overall view

4. Compute coupled response (via the moving


CFD grid approach), using as initial conditions
the blocked CFD solution and the uncoupled
structural-acoustic solution

is noted that all boundary conditions, including


the oversetgrid trilinearinterpolation formula, are
built directly into the global coefficient matrix for
the pressure-Poisson equation which results in significant acceleration of the iterative convergence.
Algorithm-related parameters to control the simulation include: 2nd-order upwind convective scheme
for the RANS regions; 2nd-order in time; PETSc
GMRES solver; 50 sub-iterations maximum for momentum and turbulence equations; 3 PISO steps;
250 sub-iterations maximum for pressure-Poisson
equation; and convergence criteria of 104 for momentum (residual based upon change between iterations) and 104 for pressure (normalized residual based upon error of discrete equation). So as
to excite the strut below, at, and above its 3rd
cantilever resonance, simulations were conducted
at U0 = 1.8, 2.28, 3.5 m/s, which correspond to
Re = 2.38x105 , 2.67x105 , 2.90x105 . Dimensionless
time steps e
t = U0Lt = 0.005 and 0.002 were used.

GRID GENERATION, BOUNDARY


CONDITIONS, AND COMPUTATIONAL
PARAMETERS
Singing SymmetricallyBeveled Strut
The computational domain extends from 2.0
x/L 3.0, 2.0 y/L 2.0, and 0.0 z/L 0.2.
Since the focus of this problem is vibration response
to vortex shedding, it is run in psuedo-2D mode
and is restricted to 5 points in the spanwise (z) direction. Near-wall spacing was set to Ly = 1.0x105 ,
which is an estimate based upon a turbulent flat
plate boundary layer, Re = 2.67x105 , and a target
y + = 1. Wake refinement grid was designed to have
nearly isotropic cells with = 0.0025.
Grids were generated using hyperbolic and translational extrusion capabilities of the Gridgen software from Pointwise, Inc. This process produces
high-quality nearly-orthogonal cells. The overset
grid system is shown in Figure 4. It is comprised
of 13 blocks with a total of 296,710 points and a
maximum block size of 30,805 points.
All grid systems contain a mixture of overset
and pointwise-continuous multiblock boundary interfaces. For the overset boundaries, PEGASUS
5.1 [Suhs et al., 2000] is used to compute double
fringe interpolation coefficients for all outer and
holeboundary fringe points. Level2 interpolation
is used to minimize overlap and improve interpolation quality.
Non-overset boundaries are prescribed to be either
noslip, inlet, farfield, or exit conditions. Details
of these conditions are provided in the CFDSHIPIOWA users manual [Paterson et al., 2003]. It

Propulsor StatorBlade Analysis

The computational domain extends from 1.0


x/L 6.0, the range of y/L is defined by the blade
number, and 0.0 z/L 0.2, the latter of which
is 4 times the strut thickness and 8 times the transverse width of the separated region at the trailing
edge. Near-wall spacing was set to Ly = 6.0x106 ,
which is an estimate based upon a turbulent flat
plate boundary layer, Re = 2.18x106 , and a target
y + = 1. Wake refinement grid was designed to have
nearly isotropic cells with = 0.001.
Grids were generated in the same fashion as the
strut, resulting in the overset grid system shown in
Figure 5. It is comprised of 17 blocks with a total of
1.47x106 points and a maximum block size of 96,600
points. In comparison to grids used by Paterson and
Peltier [2004], this grid system is fairly compact in
8

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

reduced, two-dimensional forms of the CFD and


structural-acoustic simulations. As such, URANS
flow solution is sufficient since DES is inherently a
3D flow solution technique.
Since lock-in of shed vortices and structural resonances only occurs when the respective frequencies
are very nearly coincident, we investigate only the
response of a single mode of the strut, the third
cantilever plate bending mode defined previously.
When considering only a single mode, Equation 16
is greatly simplified, and has only a single degree of
freedom to solve for. The generalized forcing function applied to the blade (F(t)) is also simplified,
as we assume the twodimensional mode shape of
the strut to be of constant amplitude, so that the
generalized force becomes simply the integrated unsteady lift, and:

Figure 5: Overset grid system for stator blade section

+ bn (t)

mn (t)
+ kn (t) = L(t).

(20)

Since the mode shape amplitude is constant over


the cross-section, the displacement is just equal to
(t). To simplify selection of parameters and develop a consistent numerical scheme, Equation 20
is nondimensionalized in the same fashion as the
RANS equation

e
e
e
m
fn (t)
+ ben (t)
+ kf
n (t) = CL (t)

Figure 6: Periodic linear cascade

(21)

2bn
2kn

n
e
f
e
where m
fn = 2m
L3 , bn = U0 L2 , kn = U02 L , = L
and CL is the lift coefficient. This effectively reduces the number of free parameters to one the
Reynolds number.

that highfidelity wake resolution extends to only


1/3 chordlength downstream.
A unique aspect of our work is the spanwise variation of grid resolution. On the fore- and mid-body
and in the far-field, only 5 points are used to span
the z-direction of the domain. In contrast, in the
LES region, 101 uniformly spaced points are used.
In comparison to the strut, several other boundary
conditions are used for this problem. In the spanwise direction, translational periodicity is assumed.
In addition, as shown in Figure 6, translational periodicity is assumed in the ydirection so as to model
the statorblade row as a periodic linear cascade.
Algorithm-related parameters to control the simulation are the same as described for the strut.
Flowrelated parameters include Re = 2.18x106 ;
0 t
=
and dimensionless time steps e
t = UR
b
0.001, 0.0005, 0.00025.

Uncoupled Simulation
Uncoupled URANS simulations are conducted to determine the blocked flow field, frequency of vortex
shedding, amplitude of unsteady lift, and the linear
response of the strut.
Figure 7 shows instantaneous contours of the vorticity and eddy viscosity. A symmetric, vortex street
is clearly shown.
The corresponding unsteady lift and drag are
shown in Figure 8. The lift coefficient shows that the
wake instability develops from 0 e
t 3, a transient
phase occurs from 3 e
t 16, and that stationary
limitcycle behavior is achieved for e
t 16, the latter of which corresponds to 130 shedding periods.
The simulated nondimensional shedding frequency
is fes = fUs0L = 8.22 , which is very close to the experimental fes = 8.21 when computed using values
reported by Blake et al. [1977], i.e., fs = 235 Hz at
U0 = 3.35 m/s. If the chord length is replaced by

RESULTS AND DISCUSSION


Singing SymmetricallyBeveled Strut
To establish the feasibility of the methodology, we
investigate Blakes singing Type IIIB strut using
9

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

vorticity: -5.00E+ 01 -4.00E+ 01 -3.00E+ 01 -2.00E+ 01 -1.00E+ 01

5.00E+ 00

1.50E+ 01

2.50E+ 01

3.50E+ 01

4.50E+ 01

0.02

ZUT: 1.00E-05 3.11E-05 5.22E-05 7.33E-05 9.44E-05 1.16E-04 1.37E-04 1.58E-04 1.79E-04 2.00E-04

fs = 160 Hz

0.01

CL

Figure 7: Contours of Vorticity and Eddy Viscosity


(h + 2), where is an estimated momentum thickness, the computed Strouhal number is St = 0.25.
To study the linear (uncoupled) response of the
strut, Equation 21 is driven with the constant amplitude lift coefficient over a wide range of Re. Figure
9 summarizes the results of this exercise by plotting nondimensional displacement (normalized by
maximum displacement at resonance) vs. Re, where
Re = 2.67x105 corresponds to the speed at which fs
and fn are exactly coincident.

-0.01

-0.02

10

15

20

25

time (U 0t/L)
(a) Entire simulation

Coupled Simulation
When the flow and dynamic simulations are coupled,
the response at resonance should increase significantly. Fullycoupled simulations were performed at
3 speeds cooresponding to a priori estimated vortex
shedding frequencies of 0.9n , 1.0n , and 1.1n .
Each case was integrated for 20,000 time steps
with e
t = 0.002. Unsteady lift and drag coefficients are shown in Figure 10. The resonant and
nearresonant conditions show CL magnitude ranging from 2 to 10, which are 150 and 700 times larger,
respectively, than the uncoupled value shown in Figure 8. Note that these results are somewhat preliminary at the time of manuscript submission in that
the resonant and nearresonant cases had not yet
achieved fully stationary results after 20,000 time
steps.
In addition to the uncoupled response, Figure
9 also shows the displacement magnitude for the
coupled simulations. Once again, displacement is
normalized by the maximum displacement of the
blocked solution, so that amplification of response
due to the lock-in process is easily seen in the plot.
When the shedding and resonance frequencies exactly coincide, the structural response increases by
about 50 dB over that of the uncoupled case. At
speeds 10% lower and higher than the coincident
speed, the amplification is still very large (30-40 dB),

0.02

fs = 160 Hz

0.0075

0.01

CD

CL

0.007
0

0.0065
-0.01
CD
CL

-0.02

20.6

20.8

21

0.006

time (U 0t/L)
(b) Detailed view over several shedding periods

Figure 8: Uncoupled time history of lift and drag


coefficients

10

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

indicating that the vortex shedding process is nearly


locking in to the resonance frequency. The vibration
amplification is consistent with the unsteady lift amplification shown in Figure 10.
The amplifications observed by Blake et al. [1977],
also plotted in Figure 9, are not as strong as those in
the coupled URANS/structural-dynamic simulation.
Since the quasi-2D URANS simulation is perfectly
coherent in the spanwise direction (the vortex shedding and structural response are assumed constant
along the strut span), this discrepancy is not surprising. Future simulations will model the strut in threedimensional space, employing full DES CFD analysis and finite element structural-dynamics analysis.
The amplification of lift and structural response at
lock-in will be compared again to Blakes data to
assess the model accuracy.
For the resonant case, the impact of coupling on
the flowfield is shown in Figure 11. The pressure
field displays large departure from the uncoupled
case. As the strut vibrates, the periodic displacement of fluid creates a large high and low pressure
region on alternating sides of the strut, which is obviously responsible for the large CL shown in Figure
10. Vorticity and eddyviscosity contours show that
the wake structure has been significantly altered in
comparison to that shown in Figure 7. Instead of
the typical vortex street with alternating vortices,
the resonant case shows two lines of vortices which
follow a nearly constant path downstream. Also the
magnitude of both vorticity and eddy viscosity are
a factor of 2 larger than the uncoupled simulation.
Finally, the leadingedge view of the vorticity field
shows that with strut vibration, the unsteady change
in angleofattack sets up a travelingwave excitation of the boundary layer. Wave length near the
leading edge is nearly equal to the strut thickness,
however details of the boundary layer response have
not yet been fully examined.

60

20log10(D/D max (blocked))

40

Coupled
From Blake (1977)
Blocked (uncoupled)

20

-20

-40

-60
1.0E+05

1.0E+06

Re=U oL/

Figure 9: Strut response: maximum diplacement vs.


Reynolds number

CL

0.006

CD

0.008

0.004

-2
30

30.2

30.4

30.6

30.8

0.002

time (U 0t/L)
10

0.08

0.06

CD

CL

-5
0.04

-10
31.4

31.6

31.8

32

Propulsor StatorBlade Analysis

time (U 0t/L)

0.008

0.006

In this section, DES results are presented for the


rigid foil only. Fully coupled simulations will be undertaken soon and reported elsewhere. Here, DES
was integrated over 93,400 time steps, which is almost 2 orders-of-magnitude longer than the simulations reported by Paterson and Peltier [2004].
The simulation was accomplished using 54,500 CPU
hours (3202 wallclock hours or 133 days) in the
background queue on the IBM SP3 at ARLMSRC.
Analysis of the statistics computed over various sample sizes indicates that this integration time was excessive. Stationary results with sufficiently accurate

CD

CL

-2
0.004
-4
13.6

13.7

13.8

13.9

14

14.1

14.2

14.3

14.4

time (U 0t/L)

Figure 10: Time history of lift and drag coefficients: coupled simulations. Re from top to bottom:
2.38x105 , 2.67x105 , 2.90x105

11

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

U
1.20
1.10
1.00
0.90
0.80
0.70
0.60
0.50
0.40
0.30
0.20
0.10
-0.00
-0.10
-0.20

(a) Pressure field

Figure 12: Instantaneous axial velocity

(b) Vorticity and eddyviscosity

Figure 13: Iso-surfaces of intrinsic swirl parameter


ized helicity, show that the first two instability
modes have been resolved, i.e., spanwise rollers
and longitudinal ribs. Also apparent are horseshoe vortices and small scales very near the trailingedge surface. These observations are similar to those
made for DES of flat struts [Paterson and Peltier,
2004].

(c) Leadingedge vorticity

Figure 11: Flowfield visualization for resonant conditions


statistics can be achieved with 2500 time steps. Results of this simulation are presented through discussion of the instantaneous flow field, and statistical
and Fourier analyses.

Statistical Analysis
Mean and root-mean-square (RMS) statistics
are computed for each dependent flow variable
(U, V, W, k, , p) at every point in the grid. For
the results presented here, statistics were calculated over 73000 time steps. This corresponds to
a non-dimensional integration period of 18.25 for
t = 0.00025. Since the statistics were nearly constant in the spanwise direction, it is assumed that
the uncertainty due to sample size is small.
Although not shown here, the mean axial-velocity
and pressure fields display characteristics typical of
steadyflow simulations of a hydrofoil; thin bound-

Instantaneous Flow Field


The instantaneous flow field at an arbitrary time
step is shown in Figures 12 and 13. Axialvelocity
contours show a sinuous wake, characteristic of vortex shedding, velocity increase on the suction side
of the blade, and the effects of periodicity in the
transverse direction.
Isosurfaces of the intrinsic swirl parameter
[Berdahl and Thompson, 1993], colored by normal12

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

ary layer and wake, increased velocity and decreased


pressure on the suction side, and a small region of
flow separation near the trailing edge.
The RMS axial-velocity and pressure fields are
shown in Figure 14. RMS velocity shows a double
lobed pattern in the near wake with a peak value of
0.15, typical of blunt trailingedge hydrofoils, e.g.,
[Blake, 1975], and where each lobe corresponds to
the trajectory of shed vortices. Upstream of the
trailing edge, RMS velocity is close to zero.
RMS pressure is similar to the velocity field in that
the peak value occurs in the near wake. However,
the RMS pressure field shows an increase over the
aft portion of the blade on the suction side, and over
the leading edge/forebody.

0.025

0.8
0.6

0.02
Cp mean
Cp rms

0.4

0.015

Cp rms

Cp mean

0.2
0

-0.2

0.01

-0.4
-0.6

0.005

-0.8
-1
3.2

3.25

3.3

3.35

3.4

Urms
0.20
0.18
0.16
0.14
0.12
0.10
0.08
0.06
0.04
0.02
0.00

Figure 15: Mean and RMS surface pressure


the aft portion of the blade, and increases, on both
sides of the foil, towards the leading edge. The reason for this unknown, however; we suspect that the
effect of and camber generate an interaction which
is absent in the non-lifting struts.
Fourier Analysis
(a) RMS axialvelocity contours

The time histories of CL and CD , and the corresponding fast Fourier transforms (FFT), are shown
in Figure 16. Three peaks are observed in both
FFTs,
e =87, 175, and 332, however, CL is larger
for the lower two frequencies, and CD is larger for
the third. Note that in the timehistory curves, the
various colors represent individual simulations, each
of which were computed in a 48hour queue.
Figure 17 shows analysis of the trailingedge pressure at two points across the trailing edge, locations
of which are indicated by red symbols. The time histories of pressure show both the relative difference
across the trailing edge, and the relatively rich frequency content. As expected, the FFTs show the
same 3 peaks as were displayed in Figure 16.
In addition to the FFTs of the DES results,
the spectra produced by CHAMPs turbulent
boundarylayer forcingfunction (TBLFF) model
[Peltier and Hambric, 2004] are included. TBLFF
is a stochastic model for the spacetime TBL wall
pressure spectrum which uses statistical data from
RANS, or the the mean field from a DES, as input. The TBLFF formulation permits the surface
pressure power spectra to respond appropriately to
favorable, zero, and adverse pressure gradients. This

PRrms
0.07
0.07
0.06
0.06
0.05
0.04
0.04
0.04
0.03
0.03
0.02
0.01
0.01
0.01
0.00

(b) RMS pressure contours

Figure 14: RMS Flow Field


Details of the surface pressure are shown in Figure 15. Again, mean pressure shows a typical hydrofoil solution; near uniform loading over most of
the blade, adverse pressure gradient on suction side
trailing edge, and constant pressure across separated
flow region. RMS pressure, on the other hand, departs from simple flat struts [Blake, 1975, Paterson
and Peltier, 2004] in that it shows a large hump over
13

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

0.2675
0.0400

0.040

0.2650

0.035
0.030

0.0390

0.2625

0.025
0.020

0.2600

0.015

0.2575

Suction
Pressure

0.010

Pressure

CL

CD

0.0380

0.0370

0.2550

0.005
0.000
-0.005
-0.010

0.0360

-0.015
-0.020

0.2525

-0.025
5

10

15

20

10

time (U 0t / R b)

15

-0.030

20

time (U 0t / R b)

-0.035
-0.040

10

15

20

time ( U 0t / R b)

(a) Lift time history

(b) Drag time history


(a) Probe location

(b) Pressure time history

60

70
175

332

10 log10 (pp() U o / R b U o)

90

10 log10(pp() Uo / Rb Uo)

-60

87

80

100

110

120

130 1
10

CD
CL

-80

-100

-120

10
Rb/Uo

-140 1
10

DES, Suct. Side


DES, Pres. Side
TBLFF, Suct. Side
TBLFF, Pres. Side

10

(c) FFT of lift and drag time histories

Figure 16: Analysis of Lift and Drag Forces

10

10

R b/U o

(c) Pressure autospectral densities

1
Coherence

0.8

effect, as shown by the difference between pressure


and suction sides, is resolved. Moreover, the DES
and TBLFF spectra show good agreement in magnitude for
e < 400. Above this point, the DES
spectra falls off rapidly due to the filter width of the
local grid size.

0.6
0.4
0.2
0 1
10

10
R /U

Phase angle (degrees)

Finally, the coherance and phase angle between


the two points show that at
e = 87, they exhibit a
coherence of 0.5 and are -180 out of phase. This frequency corresponds to the primary vortex shedding
frequency from each side of the foil and to the first
peak in the CL spectrum. At
e > 300, the correlation goes to 1.0, and the phase angle between the
two points becomes zero. This corresponds to the
peak in CD and indicates a mode which is 4 times
the principal shedding frequency.

200
100
0
100
200
1
10

10
Rb/Uo

(d) Coherence and phase angle between pressure and


suction sides

Figure 17: Analysis of trailingedge pressure

14

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

cumferential harmonics clustered around a single frequency, each of which will couple to shed vortices in
a different manner. Since the relative phasing of the
modal vibrations of adjacent blades depends on circumferential modal harmonic, they will induce different relative phasing of the vortex shedding from
adjacent blades, leading to a very complex, threedimensional coupled flow and structural-vibration
field.

SUMMARY AND FUTURE WORK


A physicsbased simulation methodology for predicting hydrofoil singing has been developed. Two
model problems, each of which focuses on specific
aspects of the methodology, namely fully-coupled
fluid dynamics and structural acoustics, and hybrid RANS/LES simulation of trailingedge turbulent flow for a practical propulsor geometry, have
been studied.
A singing symmetricallybeveled cantilevered
strut was studied to establish the feasibility of coupling unsteady CFD and structuralacoustics simulations. The uncoupled shedding frequency was
shown to agree well with experiment, and linear response of the strut was shown to be large, but only
at a very specific speed which corresponds exactly
to coincidence of shedding and modal natural frequencies. Fullycoupled simulations, on the other
hand, showed a much broader and larger response
indicating lockin of vortex shedding and structural
vibration at nearresonant conditions. The impact
of lockin on the flow field was shown to be significant with large increase in oscillatory pressure over
the strut, reorganization of the vortex street, and
generation of travelingwaves in the boundary layer
due to unsteady changes in angle of attack.
DES for a propulsor stator blade was undertaken to study the unsteady flow field generated by
trailingedge vortex shedding. Results were shown
to be similar to previous DES work, however, this
case illustrates the effect of angle of attack and camber which were not present in earlier work. Spectra of both the unsteady forces and trailingedge
pressure show 3 distinct peaks corresponding to the
vortexshedding frequency and two super harmonics, the latter of which is directly related to unsteady
drag. All 3 peaks are potentially close to the resonant response of the the circumferential harmonic
of the stator row 1stbending mode, and thus motivates future work in coupling the DES and structural
acoustics for this specific problem.
In the near future, the entire Type IIIB strut
will be re-analyzed using DES coupled with structural dynamic motions. The coupled response will
be compared to both the simulations presented and
to the measurements of Blake et al. [1977] to evalute
the anticipated improved accuracy, due to resolution
of turbulence interaction with the stuctural dynamics.
Also, a coupled, multi-modal analysis will be conducted using the procedures described here. This is
important, since multi-bladed propulsor components
typically have many blade resonances of different cir-

ACKNOWLEDGEMENTS
This work was supported by ARL Penn State IR&D.
The DoD High Performance Computing Modernization Office and Army Research Laboratory-Major
Shared Resource Center are acknowledged for providing computing resources.
REFERENCES
Satish Balay, Kris Buschelman, William D.
Gropp,
Dinesh Kaushik,
Matt Knepley,
Lois Curfman McInnes, Barry F. Smith,
and Hong Zhang.
PETSc home page.
http://www.mcs.anl.gov/petsc, 2001.
C.H. Berdahl and D.S. Thompson. Eduction of
swirling structure using the velocity gradient tensor. AIAA Journal, 31(1), 1993.
W.K. Blake. A statistical description of pressure and
velocity fields at the trailing-edges of a flat strut.
Report 4241, David W. Taylor Naval Ship Research and Development Center, Bethesda, MD,
December 1975.
W.K. Blake. Excitation of plates and hydrofoils by
trailing edge flows. Journal of Vibration, Acoustics, Stress, and Reliability in Design, 106:351
363, July 1984.
W.K. Blake, L.J. Maga, and G. Finkelstein. Hydroelastic variables influencing propeller and hydrofoil
singing. In Robert Hinkling, editor, ASME Winter
Annual Meeting, Atlanta, GA, Nov 1977. ASME.
D.A. Boger, J.W. Lindau, A. Arabshahi, R.B. Medvitz, and J.E. Roremba. Computational modeling
and validation of the steadyahead performance
and cavitation breakdown behavior of the first
generation torpedo defense vehicle. ARL-PSU
Technical Memorandum File No. 02-077, ARL
Penn State, State College, PA, July 2003. Limited Distribution.
15

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

C. Evangelinos and G.E. Karniadakis. Dynamics


and flow structures in the turbulent wake of rigid
and flexible cylinders subject to vortexinduced
vibrations. Journal of Fluid Mechanics, 400:91
124, 1999.

F.R. Menter. Two-equation eddy viscosity turbulence models for engineering applications. AIAA
Journal, 32(8), 1994.
E.G. Paterson and L.J. Peltier. Detachededdy simulation of high reynolds number beveledtrailing
edge flows and wakes. In Symposium on LES Advancements and Applications, ASME FED Summer Meeting, Charlotte, NC, July 2004. to appear.

S.A. Hambric, M.L. Jonson, J.B. Fahnline, S.C.


Conlon, R.L. Campbell, and P.D. Lysak. Simulating the vibroacoustic response of fluidloaded
structures excited by distributed fluactuationg
forces using a modal approach: a framework for
the champ cpability. ARL Technical Report In
Progress, ARL Penn State, State College, PA,
2004.

E.G. Paterson, R.V. Wilson, and F. Stern. Generalpurpose parallel unsteady RANS ship hydrodynamics code: CFDSHIP-IOWA. Technical Report 432, IIHR Hydroscience and Engineering,
The University of Iowa, Iowa City, IA, Nov 2003.

S.A. Hambric, L.J. Peltier, J.B. Fahnline, D.A.


Boger, and J.E. Poremba. Structural and acoustic noise sources due to turbulent flow through
and elbow formulation of analysis methods,. In
Proceedings of 2003 ASME IMECE, number SME
IMECE2003/NCA-41525, Washington, DC, Nov
2003.

L.J. Peltier and S.A. Hambric. Estimating turbulent


boundary layer wall pressure spectra from cfd rans
solutions. Journal of Fluids and Structures, 2004.
Submitted and in review.
D. Rockwell and E. Naudascher. Review - self sustaining oscillations of flow past cavities. ASME
Journal of Fluids Engineering, 100:152165, 1978.

R.T. Hartlen and I.G. Currie. Liftoscillator model


of vortexinduced vibration. ASCE J. Engineering Mechanics Division, 96:577591, 1970.

D. Ross. Mechanics of Underwater Noise. Peninsula


Publishing, 1987.

V.P. Ilin and Y.L. Levkovskii. Influence of edge


vibrations on hydrofoil and propeller trailing edge
noise. Acoust. Phys., 39(3):254258, 1993.

R.A. Skop and O.M. Griffin. A model for the vortex


excited resonant vobrations of bluff bodies. Journal of Sound and Vibration, 27(2):225233, 1973.
M. Strelets. Detached-eddy simulation of massively
separated flows. AIAA Paper 2001-0879, Reno,
NV, January 2001. In 39th AIAA Aerospace Sciences Meeting and Exhibit Proceedings.

G.H. Koopman and J.B. Fahnline. Designing Quiet


Structures. Academic Press, 1996.
A. Leissa. Vibration of plates. Acoustical Society of
America, pages 7687, 1993.

N.E. Suhs, W.E. Dietz, S.E. Rogers, S.M. Nash, and


T. Onufer, J. Pegasus users guide, version 5.1e.
Technical report, NASA, 2000.

D. Lucor, C. Evangelinos, L. Imas, and G.E. Karniadakis. Flowinduced vibrations of nonlinear


cables. part 2: Simulations. International Journal
of Numerical Methods in Engineering, 55:557571,
2002.

M. Wang and P. Moin. Computation of trailing-edge


flow and noise using large-eddy simulation. AIAA
Journal, 38(12):22012209, December 2000.

P.D. Lysak. A model for the broadband unsteady


forces in a marine propulsor due to inflow turbulence. Technical Report TR 01-007, ARL Penn
State, State College, PA, Dec. 2001.
P.D. Lysak and T.A. Brungart. Velocity spectrum
model for turbulence ingestion noise from computational fluid dynamics calculations. AIAA Journal, 41(9):18271829., 2003.
K.I. Matveev. Tone generation on a hydrofoil of
a highspeed ship. Ocean Engineering, 29:1283
1293, 2002.
16

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics


St. Johns, Newfoundland and Labrador, CANADA, 8-13 August 2004

CHARACTERIZING AND ATTENUATING THE LARGE-SCALE


OSCILLATIONS DOWNSTREAM OF SHALLOW CAVITIES
COVERED BY A PERFORATED-LID
STEPHEN A. JORDAN (Naval Undersea Warfare Center, Newport, RI)
ABSTRACT
Flow past cavities covered by perforated lids pose a
challenging problem for design engineers. Like their
open cavity counterpart, Kelvin-Helmholtz waves
appear early in the separated shear layers that quickly
mature into large-scale coherent structures.
This
evolution is sustained by a hydrodynamic feedback
mechanism within the cavity even when its aft wall is
far removed from the lid. Herein, the results from largeeddy simulations together with previous experimental
measurements
show
analogous
fundamental
characteristics between open and perforated-cover
cavities. Both cavities adequately scale the fundamental
frequency of the large-scale disturbance into a
dimensionless form using the characteristic freestream
velocity and the cavity width (or lid length). These
frequencies jump to higher modes at equivalent
increasing length scales. Unlike an open recess, the
perforated-lid cavity offers options for mitigating (or
even eliminating) the instability. LES results (for both
laminar and turbulent upstream separation) show the
perforation spacing as the key factor.
While
maintaining the same fundamental frequency, one can
easily dampen its spectral peak to complete
disappearance (for laminar separation) by extending the
perforation spacing. Extension sleeves placed within
the perforations gave an additional modest reduction of
the spectral pressure peaks as well.
INTRODUCTION
Incompressible flow past open cavities can generate a
large-scale instability that remains coherent for many
cavity lengths downstream. This instability originates
early within the separated upstream shear layer that has
destabilized due to a feedback mechanism from a
hydrodynamic event occurring within the cavity itself.
Given certain geometric and kinematic conditions, this
process can be self-sustaining whether upstream

separation is laminar or turbulent and the cavity is


shallow or deep. This phenomenon is most likely
responsible for the large-scale pressure disturbances
detected downstream of certain shallow cavities
recessed within US submarines as well as the internal
vibrations within the cavity itself. Other distinct
examples include automobile sunroofs and bomb bays
of low-speed aircraft. The difference between these
geometries and that of the US submarine is the latter
opening is covered by a perforated lid. While the
destabilized flow traverses over the covered cavity, the
pressure feedback perturbations travel upstream
internally; specifically, under the holes. Even with flow
communication through the holes, the cavity lid
provides sufficient separation of these two
hydrodynamic characteristics that permits them to exist
simultaneously. While external spoilers adequately
quiet the destabilized hull flow, they do not treat this
prevailing root-cause.
Knowing that the open cavity can sustain periodic
oscillations regardless of the upstream conditions,
attention is given to both cases involving upstream
laminar and turbulent separation.
The cavity is
considered relatively shallow which rules out
oscillations attributed to resonant transverse acoustic
waves as the destabilizing forcing function. One of the
earliest investigations characterizing this cavity flow
problem stems from the laboratory experiments by
Sarohia1,2 who tested three axisymmetric models. In
each experiment, upstream separation was laminar. His
experiments lead to certain quantitative criteria
necessary to suppress an oscillatory state. For
dimensionless cavity depths d/o > 2.5, a maximum
dimensionless cavity width b Re o / o ~ 300 will
inhibit amplified disturbances in the shear-layer while
traversing the cavity opening. Beyond this width, the
cavity will instigate growth into large-scale instabilities
of the first mode (f1b/U). If the cavity is widened, the
initial value f1b/U modestly rises until a sudden jump is

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

detectable into a second mode. This higher mode is


synonymous with an abrupt shortening of the instability
wavelength. A transition phase is possible over a
narrow band where both modes exist simultaneously.
But by slightly widening the cavity, the first mode
instability will disappear altogether. With further
widening, higher modes are possible until the cavity
reaches wake mode3 where the separated flow is
suddenly visualized as a bluff-body wake. Under a
fixed Reynolds number, Sarohia2 developed an
empirical relationship between the disturbance
wavelength and the cavity width; b/ = N+1/2 where N
is the mode integer.
Gharib and Rosko3 provided a detail understanding
of the modal behavior in dimensionless units. For
example, given the Reynolds number Re = 95 and a
shape factor H = / = 2.5 at upstream separation
(laminar flow), Gharib and Rosko measured the second
mode beginning at cavity width b/ = 81 and the third
mode disappearance marked at b/ = 155. They
detected a transition phase characterized by 122 < b/o <
132 where the predominate frequency of the third mode
was nearly equivalent to the previous mode. In the third
mode, the momentum thickness grew linearly (d/dx =
0.031) through the separated shear layer that housed the
disturbances. Gharib and Rosko noted that this growth
is close to the measured value (0.035) of the turbulent
mixing layer as reported by Liepmann and Laufer4.
Our understanding of the physical mechanism
responsible for sustaining an oscillating cavity flow is
largely attributed to the extensive efforts of Rockwell et
al.5-9. The essential feedback mechanism is actually
present in a much wider class of geometries including
planar (or axisymmetric) jets and mixing layers
impinging on a downstream structure. In the case of the
open cavity configuration, the flow impinges onto the
downstream wall. This act instigates an instantaneous
upstream propagation that subsequently destabilizes the
separated shear layer. Kelvin-Helmholtz instabilities
form within the shear layer that amplified into largescale coherent structures prior to their impingement
onto the downstream wall. Rockwell and Naudascher6
categorized this vortex evolution as fluid-dynamic
whose process is devoid of an acoustic delay. Recently,
Rowley et al.10 successfully applied this physical
perception to exposed cavities where the flow was
compressible and separation was laminar.
Detection of self-sustained oscillations at low-speed
turbulent separation can be traced back to the early
measurements by East11. These pressure measurements
were taken in an aerodynamic deep cavity at Mach
number Ma = 0.18. East detected strong acoustic tones
emanating from the opening whose frequencies

correlated well with the fundamental acoustic resonance


of the cavity. Similar to the case of laminar separation,
the dimensionless fundamental frequencies jump to
higher modes by widening the cavity opening.
Conversely, multiple modes can co-exist over a wide
range of cavity depths. Overall, lower dimensionless
frequencies occur for turbulent separation at a particular
mode as compared to the laminar case. East noted that
these predominate tones are evident only when the
disturbances within the separated turbulent layer are
amplified due to coupling between this layer and the
cavity modes. More specifically, East concluded that
the shear layer itself houses the feedback mechanism
necessary to sustain the salient acoustic tones.
Contrary to the open geometry, the perforated-cover
cavity has received little attention outside of isolating
the fluid-resonant feedback mechanisms7. King el
al.12 observed amplified instabilities within longitudinal
slots that were inline with the streamwise flow direction
of a cylindrical duct. Recently, Celik and Rockwell13
reported self-sustaining purely hydrodynamic
instabilities over a deep cavity when covered by a
specific perforated lid and approached by a laminar
boundary layer. Besides demonstrating a frequency
reduction of the salient spectral peaks, subharmonics
and modal jumps with increased impingement length,
they also found that the instability wavelength scales
with the streamwise extent of the perforations rather
than the diameter of the circular perforations
themselves. The respective dimensionless frequencies
(Strouhal number, St) agreed closely with the values
expected for the open cavity under similar geometric
and upstream flow conditions (0.5 < St < 0.6). Their
observations suggest that an analogous feedback
mechanism exits whether the cavity is covered or open.
Ozalp et al.14 subsequently extended the
experiments of Celik and Rockwell13 to investigate the
same phenomena, but approached by a turbulent
boundary layer. For hole diameters D/ < 4, they
visualized large-scale coherent vortices well above the
fine-scale structures of turbulence. They reaffirmed the
prior conclusion that the dimensionless frequency of the
large-scale disturbance remains consistent for a specific
mode of oscillation, and does not scale upon the hole
diameter. Conversely, the characteristic wavelength of
the Reynolds stress components was strongly influenced
by the perforations. Another important conclusion of
their work deals with the spanwise coherency of the
instabilities. Unlike the free shear layer spanning an
open cavity, they noted little correlation of the spanwise
structure directly adjacent to the perforated lid.
Ekmekci and Rockwell15 found self-sustained
oscillations for a slotted-lid cavity (d/b = 1.43, s/D =

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

1.67 and t/D = 1.52) that were analogous to the periodic


character of the perforated cover and open cavities.
However, unlike the previous works where the feedback
was purely hydrodynamic they investigated the
coupling effects of the first resonant mode of the cavity
with the instability periodicity given a turbulent inflow.
They easily detuned frequency coupling above and
below the plate by reducing the inflow freestream. A
self-sustaining oscillation was detected at impingement,
but its magnitude was reduced approximately four-fold.
Although these previous works uncovered very
interesting characteristics of the perforated-lid cavity,
many questions still remain in particular view of
quantifying the flow physics similar to open cavity.
Additionally, modifying the lid itself offers potential for
mitigating the instability while still ensuring its
functionality. Herein, further knowledge about the
large-scale instability downstream of this particular
geometry including its restoring hydrodynamic
perturbations is gained by examining the flow physics
resolved through large-eddy simulations (LES). Jordan
and Ragab16 among others17,18 have clearly
demonstrated the capability of the LES strategy for
resolving the fine-scale physics within separated shear
layers including final roll-up of a large-scale Karmantype structure. In view of the previous experimental
observations15, the spanwise hole spacing is sufficiently
tight to model them as slots. Thus, the streamwise
length scale is defined by the leading and trailing edge
of the first and last slot, respectively.
RESOLVED FIELD EQUATIONS
Resolving growth of the shear-layer instabilities along
the perforated cavity lid demands stretch grids to ensure
the proper spatial resolution. Stretched grid topologies
typically require reformulating the governing equations
into a curvilinear coordinate system (, , ). The
corresponding LES equations can be derived according
to the procedure suggested by Jordan19 where the first
spatial operation is formal transformation of NavierStokes equations to the curvilinear system. The
resultant system is a set of direct numerical simulation
equations that are solved in the computational domain.
Instituting a formal filter operation derives the
respective LES formulation.
For the present
application, we will assume that the filter width and
local grid spacing are equal. Thus, the resolved and
filtered turbulent fields are mathematically the same.
The differentiation and filtering commute similar to the
filter operation applied to the Navier-Stokes equations
in Cartesian coordinates. The resultant grid-filtered
equations in curvilinear coordinates become

Uk
=0
k

Continuity:

(1a)

Momentum:
~ ~k
~
g u i U k u i g x j p ik 1 ~ ~ kl u i
+
=
+ k +
gg

t
Re k
k
k

(1b)
where the transformation operation redefines the real
SGS stress ( ) in terms of the resolved Cartesian
( u , v, w ) and contravariant ( U, V, W ) velocity
k
i

components; specifically, =U u U u ) . During the


computation, each contravariant velocity component is
evaluated in terms of their resolved counterparts using
~
the definition U = ~g u .
The overbar in this
k
i

k
xj

definition denotes the filter variable whereas the tilde


symbolizes implicit filtering of the metric coefficients
~
( xkj ) and Jacobian ( ~g ) through their numerical
approximation. Finally, this LES equation system was
numerically approximated and time-advanced according
to the procedure outlined by Jordan19.
DYNAMIC SUBGRID SCALE MODEL
Turbulence scales beneath the grids spatial resolution
are termed the subgrid-scales (SGS) of the LES
computation. To model these scales, we used an eddy
viscosity relationship that has been modified for
dynamic computation of the model coefficient20-23. The
dynamic improvement22 promotes the correct
asymptotic behavior of the modeled turbulent stresses
when approaching no-slip walls. Besides this feature,
numerous applications show that the dynamic version
will correctly contribute little in the low to zero
turbulence regions of the flow domain. The model in
the transformed space is expressed as
ik + 1 3~ik ll = 2C2 | S | Sik

(2)

where C is evaluated dynamically based on the resolved


scales and ~ik is a filtered metric term defined as
~
~
~ k = ~
g xkj ij = xki . The turbulent eddy viscosity (T) is
i

defined by T = C2 | S | where S =

2Sij Sij and is

the grid-filter width. The resolvable strain-rate field


~
( Sik ) is expressed as Sik = ~g xkj Sij .
Evaluating the model coefficient dynamically in the
computation involves a unique derivation that starts
with explicitly filtering the grid-scale LES equations
(Eq. 1) a second time by a test filter20. This derivation
leads to an algebraic expression that determines the

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

coefficient over a local narrow band of finest resolved


scales as given by the test filter kernel with width ( ).
The test filter employed herein is actually a form of
volume averaging24 where the filter width is twice the
local grid spacing; = 2 . The user has the choice of
test filtering along the curvilinear lines in either the
physical domain or computational space that depends on
whether the SGS term is implemented in a nonconservative or conservative form23. However, the user
must avoid the associated second-order error produced
in the physical domain by numerically evaluating each
contributing first-order derivative prior to the explicit
filter operation. The modified Leonard term ( Lki ) of the
test-filter equations has the cognate form
t

Lki = Tik ik

(3)

in the computational space where the bar over the


Reynolds stress denotes the test filter operator. In this
identity, the modified Reynolds stress ( T ) is defined by
k
i

Tik 1 3 ~xkj Tll = 2C2 S Si k

(4)

Substituting this definition and the test-filter SGS stress


into the identity (Eq. 3) gives a second form of the
transformed Leonard term from which we determine C;
Lk 1 ~ k L = 2C2 M k
(5)
k
i

3 x j ll

The model stress ( M ) is defined in terms of resolvable


k
i

tensors as M ik = 2 S Si k S Sik where the filter width


ratio = 2.
Contracting Lki in Eq. 5 gives a unique expression
for the model coefficient C. The derivation must yield
rotational invariance of the coefficient in both the
physical domain and transformed space. The proper
procedure requires writing the error function in the
physical domain before minimizing its square. Along
the curvilinear lines in the physical domain, the
contravariant error tensor ( E ) becomes
k
i

~
E ik = ~
g xkj Lij

~k
3 x j L ll

~
2C2 ~
g xk j M ij

(6)

where matrices Lij and Mij are not summed


independently.
Minimizing this error gives the
expression
C=

Lki M ik

(7)

2 M km M km

for the model coefficient that operates on the inner


product of the Cartesian tensor components of L and
M in the computational space.
k
i

k
i

Inasmuch as the model coefficient is dependent on


the local instantaneous strain-rates of the resolved field,
either positive or negative values are possible in its
evaluation. While the positive coefficients denote
forward scatter in the turbulent energy spectrum, the
negative values symbolize backscatter, or the reverse of
energy from the SGS scales to the finest scales of the
resolved field. Experience has demonstrated that the
backscatter physics correlate well over long execution
times, which quickly demand "ad hoc" measures to
guarantee stability. However, in most simulations the
statistical contribution of the SGS model is usually
small compared to the resolved fields. Thus, the best
choice for maintaining a stable computation at the
instantaneous level is to truncate all negative
contributions from the dynamic model to yield a net
zero effect.
RESULTS AND DISCUSSION
Before considering its mitigation, we first intend to fully
characterize the large-scale instability that is detectable
downstream of the cavity covered by a slotted-lid.
Although the simulations can target a generic study of
slotted-lid cavities, the basic geometric scale herein
mimics a shallow cavity recessed within a US
submarine. The cavity itself is classified as shallow
because the streamwise length (b) of the lid slots
exceeds the cavity depth (d) as sketched in figure 1.
Additionally, the cavity is elongated meaning that its aft
wall boundary is far removed from the last set of slots
such that impingement is not necessarily the governing
mechanism for self-sustaining the periodic oscillation.
As noted earlier, the spanwise spacing of the
perforations are sufficiently tight to model the lid
geometry as a streamwise series of slots.
This
modification is justified in view of the analogous results
observed and reported by Rockwell24 for both staggered
holes and slots perforating the lid of his deep cavity
configuration (b/d < 0.5).
The final slot size and their streamwise distribution
in the LES computations copy the staggered hole pattern
of the particular scaled inlet. This configuration is
sketched in figure 1 where the slot gap width (D) to
spacing (s) is s/D = 0.25. This ratio remained constant
in the simulations because varying D (fixed s) by as
much as 30% revealed no discernible differences in the
dimensionless oscillation frequency fb/U. Besides s/D,
the cavity depth (d) also remained fixed in the
simulations with the lid length (b/d) ranging as 1.4 < b/d
< 5.4 which depicts lids perforated by three to eleven
slots. Inside the cavity the fore and aft walls were
modeled 2D and 12D beyond the slots.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

8D

d
x

Figure 1. Example Slot Geometry (five slots)


of the LES Computations.
Grid Generation and Flow Conditions. Externally,
the inflow and outflow boundaries were placed 8D
upstream and 12D downstream from the lids leading
and trailing edges, respectively. The respective upper
boundary limit was situated at least ten boundary layer
thicknesses (10o) above the cavity lids leading edge.
Various structured grids were generated that obviously
depended on the number of lid slots. The largest grid
(11 slots) housed 565x131x17 points in the streamwise
(), transverse () and spanwise (z) directions,
respectively. A spanwise spacing of z = 0.03 in the
simulations was suitable for resolving transition in the
separated shear layers as well as the energy scales of the
subsequent turbulent field.
In reference to the
wavelengths measured by Mansey et al.25, this spacing
resolves each spanwise eddy by a minimum of four
points. Using an exponential distribution function, the
grid lines were cluster towards the lid boundary such
that the first field point above the wall near exit
remained as <y+ > < 5.
At the outlet boundary, the transformed form of the
Euler equations was found satisfactory to exit the
coherent structures with little distortion. The Blasius
boundary layer equations were applied to create each
LES computation of a laminar inflow condition. At
separation the shape factor (H) of these simulations
ranged as 2.56 < H < 2.61. A fixed streamwise velocity
was upheld along the upper boundary with the twodimensional continuity equation used to determine the
respective vertical component given the spanwise
component set to zero. Finally, each simulation had a
fixed time step tU/d = 0.001.
Test cases requiring turbulent separation used a
modified form of the clever procedure called rescale
and recycle26 to generated turbulence in the inflow.
For the cavity computations, the instantaneous flow
velocities at a -z plane that were chosen downstream
from the inflow boundary and prior to separation are
rescaled and recycled to create the upstream turbulence.
In the original technique, the law-of-the-wall scales u+
and y+ define mean similarity at both planes. These

2
1.75

mode 3

1.5

slots3
open1

1.25

open1

fb/U o

u(y)

staggered2
(honeycomb)

1
0.75

Re o d o

mode 1

0.5

920
1107
2860
4000

0.25

5
82
10
2.5

0
5

10

15

20

25

30

35

40

b/

(a) Conventional Length Scaling


2

Staggered2
d/0 = 82

1.5

fb/U

scales are rescaled and recycled to the inflow boundary


according to the differences in momentum thickness.
However, exercising this original technique for the
covered cavity simulations produced undesirable
periodic vorticity within the inflow that influenced the
fundamental frequency of the oscillation. This vorticity
arose due to the proximity of the separation point to the
recycle boundary coupled with the incompressibility
constrain that is satisfied at each time step.
As an alternate choice, Prantls one-seventh-power
law for the mean flow provided a suitable means of
rescaling and recycling turbulence; u/U = (y/)1/7.
Knowing the boundary thickness at both the inflow and
recycle planes, the fluctuating component at the
recycled plane can be easily extracted and superimposed
onto the mean inflow velocity. This modified procedure
produces a near consistent shape factor H = 1.3 between
the inflow and recycled boundaries.

Slots
4.5 < d/0 < 12

3rd mode

1st mode
0.5
t/D = 0.86
t/D = 0.25

0
0

500

b/
b Re o / o

1000

1500

(b) Expanded Length Scaling


Figure 2. Scaling the Large-scale Instability Frequency
and Cavity Length for Three Geometric Configurations
(open, staggered and slotted).

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Laminar Separation Characteristics: Quantifying the


large-scale disturbance downstream of a perforated lid
begins by attempting analogies to the well-understood
physics of the open cavity. Comparisons of the previous
experimental
measurements
and
the
present
computational results of the instability dimensionless
frequency fb/U as generated for laminar flow separation
over both open and covered cavities are shown in figure
2a. Two sets of hot-wire anemometry measurements1
are included for the open cavity geometry at Reynolds
numbers Re = 920 and 2860, which are based on the
boundary layer thickness at separation (o). Conversely,
the third experimental dataset13 depicts pressure
measurements taken just downstream of a deep cavity
(d/ o = 82) that is covered by a staggered hole pattern.
Each hole was offset streamwise by one hole diameter
(D = 6.4mm) and drilled through a 5.5mm thick plate
(t/D = 0.86). Most notably, these data adequately depict
the large-scale instability in terms of the downstream
predominant frequency, which is properly scaled by the
length (b) and the freestream velocity (U).
6x102
6x10

b(Re )1/2 /

Slots (slots)
55Holes
Slots (slots)
33Holes
UU
oo

dd

,,,,UU
oo

dd

Oscillations

Figure 3 gives the dimensionless length b Re o / o

No Oscillations

10

15

analogy in the disturbance character even though


significant differences exist among the cavities
geometric scales. For example, although the cavity
depth d/ o of the staggered-hole cover13 and present
slotted lid differ by a large factor both configurations
show modal jumps at the same length scale. Sarahio2
measured a first to second modal jump in his open axisymmetric cavity at length scale b Re o / o = 433,
which surprisingly agrees with the same jump computed
herein for the slotted lid. Moreover, each modal stage
indicates a similar modest rise in the dimensionless
frequency fb/U when subjected to an increased length
b Re o / o .

Open Cavity1

we note that the previous mode oscillation had not fully


disappeared from the frequency spectrum. Rockwell3
also detected jumps between modes in his staggeredhole lids, but as depicted in figure 2a they occurred at
lengths (b/ o) much higher than the open or slottedcover cavities.
As noted earlier, Sarohia1 first
recognized this important characteristic by expressing a
direct integral connection between the disturbance
wavelength () and length b; b/ ~ N+1/2. For the
present geometry, /b = 0.64 just before the frequency
jump which indicates its first state of oscillation (N ~ 1).
Better agreement between the present LES
computations and the experimental evidence can be
reached by expanding the cavity length scale (b/ o) in
terms of the boundary layer properties at separation. By
replotting the dimensionless frequency versus the length
scale b Re o / o , figure 2b illustrates remarkable

20

25

d /

Figure 3. Open1 and Slotted-Cover Cavity Results


Indicating the Minimum Streamwise Length of an
Oscillatory Flow for a Specific Cavity Depth.
A second self-similar characteristic is clearly
discernible among the four datasets in figure 2a. Each
geometry supports jumps to higher modes of the largescale disturbance when subjected to a lengthened cavity
width or lid cover (b/ o). Furthermore, while under a
particular mode of oscillation, all three configurations
indicate a moderate increase in fb/U by expanding b/ o.
Like the open cavity, computations of the instability for
the slotted-lid cavity will switch between modes simply
by adding or reducing the respective number of slots.
For laminar flow at separation, b/ o > 12 caused a jump
to the higher mode. However, at b/ o ~ 14 (11 slots)

that is required to instigate the large-scale instability for


either an open1 or slotted-lid cavity. Note that these
lengths depict generation of the disturbance very early
inside the separated shear layer after laminar separation.
The circles and squares denote the present LES results
with the open symbols specifying no oscillations
sustained in the long-term. The notations , , Uo and d
adjacent to each solid symbol note the specific
parameter that was increased to destabilize the initial
shear layer using the non-oscillatory results as the initial
flow condition. In one instance, the three parameters
, , Uo were slightly amplified to instigate the
disturbance. Of special interest is the oscillation/nooscillation boundary for d/ o < 5. At first glance, a
slightly longer length b Re o / o is necessary to sustain
the large-scale instability in the slotted-cover case
compared to the open cavity for the same depth scale
d/ o. This result is indicated for both the three-slotted
and five-slotted configurations. Not shown are the
measured dimensionless frequencies downstream of the

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

staggered-hole cover13 because b Re o / o

> 400 in

each experiment, which is well above the lower


boundary limit denoting an oscillatory state.

experimental measurements. Taken collectively,


similarity of the large-scale disturbance is evident when
placed in units of dimensionless frequency fb/U and
streamwise length b Re / o , whether the cavity is open
o

or covered by either a staggered-hole or slotted pattern.


All three configurations produce periodic instabilities
based on the long wavelength that scales according to
the cavity opening or perforation length.
Most
surprisingly, the model jumps occur at the same
dimensionless length scale b Re / o . In particular,

2.5
eOpen
cOpen
eStaggered
eStaggered
eSlots
3 Slots
5 Slots(s)
5 Slots(d)
9 Slots

fb/U

1.5

the first jump to the second mode occurs at


approximately b Re o / o = 2150 whereas the second

0.5

jump to the third mode takes place at approximately


b Re / o = 3500.
o

Mode

0
0

1000

2000

3000
4000
1/2
b(Re ) /

5000

6000

Figure 4. Comparisons of the Dimensionless LargeScale Frequency (fb/U) to the Dimensionless Cavity
Length Scale ( b Re / o ); the open symbols indicate
o

LES results, eOpen denotes experimental open cavity


data11, eStaggered signifies staggered-lid experimental
and
eSlots
symbolizes
slotted-lid
data14,
measurements15.
Turbulent Separation Characteristics: Turbulence in
the initial shear layer can inhibit growth of small-scale
disturbances into large-scale structures.
A good
example is the loss of strong coherent structures in the
wake of a circular cylinder when upstream separation is
turbulent.
Ozalp el al14 reported self-sustained
oscillations downstream of perforated-lid cavities, but
their results demonstrated similarity under a single
mode and excluded slotted covers. The following
turbulent separation results establish similarity of the
large-scale instability for multiple oscillating modes
under a purely hydrodynamic event including
comparisons of measured and LES results for both open
and covered cavities. Once the oscillation is sustained,
the flow characteristics of the turbulent shear layer are
similar in nature to the laminar separation case. Thus,
the present study dealing with turbulent separation is
only focused on the proper dimensionless scaling of the
cavity length as well as the relative values for each
oscillation mode.
Quantification of the large-scale oscillation where
the inflow condition is turbulent is plotted in figure 4 for
several cavity configurations. Given the same cavity
configuration, the fundamental frequency as predicted
by the LES computation agrees closely with the

Under each mode, all three cavities produce a


modest rise in fb/U with increased width b Re / o .
o

This behavior agrees in principle with that observed


earlier for laminar separation even though each is
expressed using a different length scale to establish
similarity. The most distinctive difference between the
two states at separation is the magnitude of fb/U. While
fb/U for the first mode is above 0.5 when separation is
laminar, the turbulent condition drops the dimensionless
frequency to values under 0.5. Once jumped to higher
modes, both conditions at separation suggest analogous
values.
When the inflow is turbulent, the most significant
difference in the large-scale oscillation among the
cavity configurations rests essentially on their
frequency. Given equivalent dimensionless widths
b Re / o , the absolute frequency of the large-scale
o

instability in figure 4 depends largely on whether the


cavity is open or covered by a perforated lid. In
particular, the cover holding the least open area to the
cavity below (eStaggered) caused the uppermost rise in
the frequency of the hydrodynamic oscillation.
Interestingly, this result contrasts the measurements of
Ekmekci and Rockwell15 for the case where the cavity
resonance is fully coupled to the shear layer oscillation.
They reported a higher value for the dimensionless
frequency of the coupled open cavity (fb/U = 0.320)
compared to the slotted plate configuration where fb/U
= 0.244. To a lesser degree, other incremental factors
such as deepening the cavity or thickening the lid will
lower the absolute frequency. For example, deepening
the slotted-lid cavity from ratio d/b = 0.42 to d/b = 2.5
in the LES computation b Re / o ~ 1600 (5 slots)
o

lowered the dimensionless frequency from fb/U = 0.52


to fb/U = 0.46. Similarly, Rockwell24 at the same
dimensionless width showed a 30 percent dropped in

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

fb/U after thickening the plate ratio from t/d = 0.86 to


t/d = 3.44.
Mitigating the Large-Scale Instability. The LES
results of both laminar and turbulent separated flow
over rectangular cavities that are covered by a
perforated plate reveal the generation of coherent
periodic structures whose fundamental frequency scales
according to the streamwise length of the lid. These
large-scale structures evolve from much smaller scale
disturbances (K-H waves) that grow almost
exponentially as they traverse the lid. The latter waves
appear early in the separated shear layer due to a small
transverse pressure perturbation that arrives from a
hydrodynamic feedback event within the cavity itself.
60
Spoiler (7 holes)
Re = 10,000

50

fb/U = 1.49
fb/U = 0.609

Sp

40
30

Downstream
Inside Cavity

20

Above 4th Slot

10
0
0.01

0.1

fb/U

10

(a) Spectral Amplitudes for Single Spoiler


1.5

However, mounting only a single spoiler upstream of


the perforations (or slots) provides little relief for
dampening growth of the initial disturbance. This
observation becomes clearly evident in view of the
pressure spectrum in figure 5a where higher
fundamental amplitudes are predicted during early
growth as oppose to further downstream when the
instability has matured to the leading coherent structure.
This figure also indicates a decrease in the fundamental
frequency of the instability as it traverses the cavity lid.
Notably, the cavity aft end reverberates at the same
frequency as the periodic oscillation. This correlation is
analogous to the test cases without spoilers and further
suggests that the origin of the pressure feedback lies
within the aft region of the cavity. Within this region
the spectral amplitudes at the fundamental frequency are
significantly lower than those outside the cavity.
The relative pressure spectra of the scaled model
with one and two spoilers are shown in figure 5b. Both
spectra depict sample points 8D for the lid trailing edge.
All dimensionless frequencies using two spoilers
indicate spectral amplitudes of the downstream pressure
that are substantially attenuated when referenced to the
single spoiler results.
A fundamental large-scale
organized structure is mildly discernible in the pressure
spectra. Its spectral amplitude is approximately 1/4th of
the single spoiler instability with a relative reduction in
the corresponding dimensionless frequency. These
reductions provide clear evidence that supports
continued use of external spoilers for successfully
interrupting the growth of large-scale instabilities over
cavities covered by perforated lids.
2.0

1.25

2 Spoilers
1

Downstream
Inside Cavity

SApp

1.5

0.75

1.0

p/ U

1 Spoiler
2 Spoilers

0.5
0.25

Pulsed
Speed

Constant
Speed

0.5

0.0

0
0.1

1
fb/U

10

(b) Spectral Amplitudes of the Both Spoilers


Figure 5. Pressure Spectrum of a Single and Double
Spoilers Mounted on Lid Housing Nine Slots.
External spoilers mounted on the cavity lid
interrupts communication between the separated shear
layer and the adjacent upper cavity, which inhibits
generation and growth of the small-scale instabilities.

-0.5
50

100

150

200

tU/d

250

300

350

400

Figure 6. Time Traces of Relative Pressure for Steady


and Pulsed Flow over the Slotted-Lid Cavity
Before accepting external spoilers as a suitable
corrective measure for interrupting the evolution of
large-scale instabilities over perforated lids, they should

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

be tested to explore other flow conditions. Under this


premise, simple steady inflow computations whether
laminar or turbulent does not cover the entire spectrum
of probabilities. More specifically, the correctness of
the external spoilers under unsteady inflow deals with
body maneuvering. One such example is shown in
figure 6 where the inflow was pulsed by twenty-five
percent of the respective steady case. The LES results
at a point downstream of the lid are not encouraging
because the time traces suggest return of a strong
oscillation. Even after sixty dimensionless time units,
the figure indicates little dampening of the pressure
response over time. On the average, peak values of
downstream pressure for the pulsed flow simulation
increased four-fold compared to that of the steady
inflow results.
1.2
s/D = 0.25
s/D1 = 1.5
s/D2 = 1.5

Sp

0.8
0.6
0.4
0.2
0
0.1

1
fb/U

10

Figure 7. Pressure Spectra Attenuation of the Instability


by Increasing the Relative Spacing of Subsequent Slots.
1.4

s/D = 0.25
s/D2 = 1.5
s/D5 = 1.5

1.3

/ 0

1.2

1.1

Understanding the analogies between the open and


slotted-lid cavities lends options for mitigating or
possibly eliminating the streamwise growth of the largescale disturbance. More specifically, options can be
assessed that passively control the disturbance
amplitude and frequency downstream of the cavity by
manipulating its growth inside the separated shear layer.
Herein, we focus on the lid perforations themselves and
their streamwise distribution. Knowing that the level of
shear layer growth across an open cavity is similar to
that of a turbulent mixing layer, the ratio s/D becomes
an important length scale for attenuating the Reynolds
stresses within that layer. Small values tend toward the
open cavity configuration while longer lengths approach
flat plate flow.
Groundwork results from several test cases for
laminar flow separation are shown in figure 7 and figure
8 where the slot spacing was incrementally increased to
s/D = 1.5 (fixed D) along the cavity lid. In particular,
the notation s/Dn in the figure denotes the n number of
slot spaces increased beyond the first. Notation s/D5
indicates conversion of the nine-slotted lid to five-slots,
but with the same cover length (b). Each test used the
LES resolved field for a nine-slotted lid (s/D = 0.25) as
the initial oscillating flow state. Clearly, lengthening
the slot spacing attenuated the large-scale instability.
The s/D1 modification reduced the peak amplitude by
40% and the s/D2 configuration by 66%. A pressure
spectrum is not plotted in figure 7 for the s/D5 test case
because the initial oscillations eventually damped to an
insignificant level.
Interestingly, the fundamental
frequency of the disturbance is unaffected by the ratio
s/D which suggests that the length scale of the feedback
mechanism remained intact within the cavity volume.
Examining the average spatial growth of the
momentum thickness () within the lid shear layers is a
leading indicator of the instabilitys streamwise
evolution. In the original nine-slotted configuration
(s/D = 0.25), figure 8 suggests that the disturbance
completed transition to a turbulent state after traversing
approximately 60% over the lids length. Growth of the
momentum thickness within the turbulent shear layer is
d/dx = 0.0123 ( b Re o / o = 1850) which is
substantially lower than d/dx = 0.022 for the open
cavity ( b Re o / o = 1635) and d/dx = 0.035 for the

0.9
0

20

40

60

80

x/ 0

Figure 8. Reduction in the Streamwise Growth of the


Momentum Thickness by Increasing the Relative
Spacing of Subsequent Slots.

turbulent mixing layer. Extending the second and third


slot spacing (case s/D2) not only delayed transition, but
also lowered the instability growth rate as reflected by
its reduced peak in the pressure spectra. Simulating a
five-slotted lid ( b Re o / o = 2296) with s/D = 1.5 for
all the slots inhibited long-term growth of the
disturbance altogether. Observations showed that even

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

1.2
9 Slots
Variable Slots

f1b/U = 0.68

0.8
0.6

Sp

the separated shear layers in the latter slot gaps


reattached to the lid. Moreover, the free shear layers
and bounded layers remained laminar across the lid.
However, the slot gaps did cause jumps in the growth
rate such that the averaged value d/dx = 0.0003
exceeds that of a fully attached laminar boundary layer
(d/dx = 0.0002) under equivalent upstream conditions.

0.4
2.8

0.2
2.4

0
0.01

0.1

/ 0

2.0

fb/U

10

(a) Pressure Spectra

1.6

100

1.2

0.8
0

50

x/ 0

100

150

10

200

Figure. 9. Reduction in the Streamwise Growth of the


Momentum Thickness by Increasing the Relative
Spacing of Subsequent Slots and Adding Sleeves.
Restraining growth of the initial disturbance by
altering the streamwise slot spacing for the test case of
turbulent inflow is illustrated in figure 9 ( b Re o / o =
2580). The original lid in this case holds nine equally
spaced slots. Unlike the spatial averages shown in the
previous figure (figure 8), the present figure shows
time-averages (five periodic cycles) at each grid point
along the lid surface. Early growth of the momentum
thickness across the nine-slotted lid is approximately
d/dx = 0.0192, which approaches the turbulent open
cavity value of d/dx = 0.022 and far exceeds the
turbulent flat plate value of d/dx = 0.002 at this
Reynolds number (Reb = 66000). Selectively extending
the spacing between slots (especially to control the high
level of early growth) played a dramatic role in the
overall evolution of the large-scale instability. In no
case was the oscillation eliminated as in laminar
separation. The objective here was to create the
disturbance wave, then widen the subsequent slot
spacing until the final oscillation dampened to an
acceptable level. Figure 9 illustrates the slot spacing
needed to meet these criteria. Growth of the momentum
thickness was lowered to approximately d/dx ~ 0.0021,
which is only five percent higher that that of a fully
turbulent boundary layer across the lid. Further minor
reductions are possible by introducing elongated sleeves
through the slots as illustrated in the figure.

E11 /U2d

-50

b Re / = 2580
1

b Re / = 3240

0.1

0.01
0.1

fb/U

10

100

(b) Streamwise Energy Spectra


Figure 10. Attenuation of the Pressure and Streamwise
Energy Spectra of the Large-Scale Instability by
Increasing the Relative Spacing of Subsequent Slots;
Separation is Turbulent.
Attenuation of the instability pressure signal due to
varying the slot spacing in the turbulent inflow test case
is plotted in figure 10a. Both profiles depict data
collected 8D downstream of the lid where the peak
spectral amplitude of the nine-slot computation was
used to scale each profile. By maintaining constant
spacing between the slots, a strong pressure oscillation
of the large-scale instability is clearly produced that
would be easily detected in the far field. Two strong
harmonics (2f1b/U = 1.27 and 4f1b/U = 2.54) are also
unmistakably visible in the pressure spectrum. Beyond
these harmonics the spectral amplitudes drop to
insignificant levels. Extending the streamwise slot
spacing not only lowered the pressure amplitudes of the
instability, but also obscured detection of its
fundamental frequency. Instead, the pressure spectrum
suggests several small-scale structures of varying

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

frequencies that house only low amplitudes because


they lack an apparent hydrodynamic feedback event
within the adjacent cavity to self-sustain strong growth.
The spectral energy profiles shown in figure 10b
support this observation. These profiles represent the
streamwise velocity fluctuations (u2) that were taken
inside the last slot of each configuration. A single
fundamental frequency is clearly apparent only in the
energy spectra of the nine-slot simulation. More
importantly, the energy levels at all frequencies are
substantially reduced (at least and order-of-magnitude)
by varying the slot spacing along the lid. The dashed
lines in the figure denote Kolmorgovovs 5/3s law for
isotropic turbulence. Comparing these lines with the
LES results indicates use of a local grid resolution in the
computation sufficient to resolve the inertial sub-range
of the turbulent energy spectrum.
FINAL REMARKS
Diminishing the streamwise growth of a large-scale
instability across a recessed cavity that is covered by a
series of perforations is an important attribute in design
engineering. At the very least, we desire to avert its
growth into a resilient periodic structure that remains
coherent far downstream as well as cleanly detectable in
the remote field. The present investigation showed that
both open and perforated-cover cavities all instigate the
same disturbance. Proper scaling of the fundamental
frequency into a dimensionless form requires the same
characteristic length and velocity scales as the open
cavity. Moreover, the instability jumps to higher modes
given the same dimensionless length scale in terms of
the cavity width and boundary layer parameters at
upstream separation. Sustaining the self-oscillation is
achieved by destabilizing the initial separated shear
layer through a hydrodynamic feedback mechanism that
exits whether the cavity is open or covered and deep or
shallow, even if the aft inside wall is far remove from
the perforations as illustrated in the present study.
Options for passively mitigating the oscillation
become obvious when dealing with perforated-cover
cavities. Although external spoilers are an effective
means of control, lengthening the streamwise
distribution of the perforations provides an enhanced
flexibility for dampening this disturbance. This option
is effectual for both laminar and turbulent upstream
boundary layers. We can easily mitigate the oscillation
to any level simply by invoking non-uniform hole
spacing, but this control should take place obviously
before the wave reaches maturity. Conversely, the
dimensionless frequency of the damped instability
appeared unaffected by the variable distribution.

ACKNOWLEDGEMENTS
This author wishes to thank R. D. Joslin (program
officer, Office of Naval Research, Code 333) and R. B.
Philips (NUWC, Code 10) for their support of LES
research related to flow over covered cavities. The
computational work was supported in part by a grant of
HPCMP resources from the Arctic Region
Supercomputing Center.
NOMENCLATURE

C
=
Cr
=
D
=
E
=
=
Eij
H
=
Lij
=
Mij
=
N
=
Re
=
=
Sij
Sp
=
Tij
=
U,V,W =
=
Uo
Uc
=
b
=
d
=
f
=
~
g
=
ij
g
=
m, n =
p
=
s
=
t
=
u,v,w =
u
=
x,y,z =

=
t
=
g
=
=

=
=
T

=
, , =
~
=

model coefficient
cross-correlation coefficient
slot gap width
model error tensor
spectral energy
shape factor
modified Leonard term
model stress tensor
oscillation mode
Reynolds number
strain rate tensor
pressure spectra
modified Reynolds stress tensor
contravariant velocity components
freestream velocity
instability velocity
cavity streamwise length
cavity depth
frequency
filtered transformation Jacobian
metric coefficients
mixed mode integers
pressure
slot spacing
lid thickness
Cartesian velocity components
streamwise turbulent intensity
Cartesian coordinates
filter width ratio
test filter width
grid filter width
boundary layer thickness at separation
momentum thickness
turbulent eddy viscosity
wavelength
fluid density
Reynolds stress
curvilinear coordinates
filtered metric coefficient

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

REFERENCES
1.

2.

3.

4.
5.

6.

7.

8.

9.

10.

11.

12.

13.

Sarohia, V., (1975), Experimental and Analytical


Investigation of Oscillations in Flows Over Cavities,
PhD Thesis, California Institute of Technology.
Sarohia, V., (1977), Experimental Investigation of
Oscillations in Flows Over Shallow Cavities, AIAA
Journal, Vol. 15, No. 7, pp. 984-991.
Gharib, M. and Roshko, A., (1987), The Effect of Flow
Oscillations on Cavity Drag, Journal of Fluid
Mechanics, Vol. 177, pp. 504-530.
Liepmann, H.W. and Lufer, J.,(1947), Investigation of
Free Turbulent Mixing, NACA Technical Note No. 1257.
Rockwell, D., (1977a), Vortex Stretching due to Shear
Layer Instability, Journal of Fluids Engineering, Vol.
99, pp. 240-244.
Rockwell, D., (1977b), Prediction of Oscillation
Frequencies for Unstable Flow Past Cavities, Journal of
Fluids Engineering, Vol. 99, pp. 294-300.
Rockwell, D. and Naudascher, E., (1979) Self-sustained
Oscillations of Impinging Free Shear Layers, Annual
Review Fluid Mechanics, Vol. 11, pp. 67-94.
Rockwell, D. and Knisely, C., (1980) Observations of
the Three-dimensional Nature of Unstable Flow Past a
Cavity, Physics of Fluids, Vol. 23, No. 3, pp. 425-431.
Knisely, C. and Rockwell, D., (1982), Self-sustained
Low-frequency Components in an Impinging Shear
Layer, Journal of Fluid Mechanics, Vol. 116, pp. 157186.
Rowley, C.W., Colonius, T. and Basu, A.J., (2002), On
Self-sustained
Oscillations
in
Two-Dimensional
Compressible Flow over Rectangular Cavities, Journal
of Fluid Mechanics, Vol. 455, pp. 315-346.
East, L.F.,(1966), Aerodynamic Induced Resonance in
Rectangular Cavities, Journal of Sound and Vibration,
Vol. 3, No. 3, pp. 277-287.
King, J.L., Boyle, P. and Ogle, J.B., (1958), Instability
in Slotted Wall Tunnels, Journal of Fluid Mechanics,
Vol. 4, pp. 283-305.
Celik, E. and Rockwell, D., (2002), Shear Layer
Oscillation along a Perforated Surface: A Self-excited
Large-scale Instability, Physics of Fluids, Vol. 14, No.
12, pp. 4444-4447.

14. Ozalp, C., Pinarbasi, A., and Rockwell, D., (2003) Selfexcited Oscillations of Turbulent Inflow along a
Perforated Plate, Journal of Fluids and Structures, Vol.
17, pp. 995-970.
15. Ekmekci, A., and Rockwell, D., (2003) Self-sustained
Oscillations of Shear Flow Past a Slotted Plate Coupled
with Cavity Resonance, Journal of Fluids and
Structures, Vol. 17, pp. 1237-1245.
16. Jordan, S.A. and Ragab, S. (1998), A Large-eddy
Simulation of the Near Wake of a Circular Cylinder,
Journal of Fluids Engineering, Vol. 120, pp. 243-252.
17. Kravchenko, A.G.. and Moin, P., (2000), Numerical
Studies of Flow over a Circular Cylinder at ReD = 3900,
Physics of Fluids, Vol. 12, No. 2, pp. 403-417.
18. Breuer, M., (1998), Numerical and Modeling Influences
on Large-Eddy Simulations for the Flow Past a Circular
Cylinder, International Journal of Heat and Fluid Flow,
Vol. 19, pp. 512-521.
19. Jordan, S. A. (1999), A Large-eddy Simulation
Methodology in Generalized Curvilinear Coordinates,
Journal of Computational Physics, Vol. 148, pp. 322340.
20. Jordan, S. A, (2003), Resolving Turbulent Wakes,
Journal of Fluids Engineering, Vol. 125, pp. 823-834.
21. Smagorinsky, J., (1963), "General Circulation
Experiments with the Primitive Equations, I. The Basic
Experiment," Monthly Weather Review, Vol. 91, pp. 99164.
22. Germano, M., Piomelli, U., Moin, P., and Cabot W.H.,
(1991) A Dynamic Subgrid-Scale Eddy Viscosity
Model, Physics of Fluids, Vol. 3, pp. 1760-1765.
23. Jordan, S. A. (2001), Dynamic Subgrid-Scale Modeling
for Large-Eddy Simulations in Complex Topologies,
Journal of Fluids Engineering, Vol. 123, pp. 1-10.
24. Rockwell, D., (2003), Personal Communication.
25. Mansy, H., Yang, P., Williams, D.R., (1990),
Quantitative Measurements of Three-Dimensional
Structures in the Wake of a Circular Cylinder, Journal
of Fluid Mechanics, Vol. 270, pp. 277-296.
26. Lund, T.S., Wu, X. and Squires, K.D., (1998),
generation of Turbulent Inflow Data for SpatiallyDeveloping Boundary Layer Simulations, Journal of
Computational Physics, Vol. 140, pp. 233-258.

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

25th Symposium on Naval Hydrodynamics

Hull excitation by fluctuating and rotating


acoustic sources at the propeller
O. Rath Spivack1 , R. Kinns2 and N. Peake1
(1 DAMTP, University of Cambridge, Cambridge, U.K.)
2
( RKAcoustics, Glenavon, Back Road, Clynder, G84 0QQ, Scotland )

Abstract

floating body. The dependence on the number of


blades and on shaft frequency is also explored. We

We present here a preliminary study of the effects

give quantitative results regarding the effects of the

of acoustic sources associated with a rotating pro-

rotating forces and volumes, and of the net fluctu-

peller on a symmetric hull. The propeller is mod-

ating tailshaft force, which should prove valuable in

elled as a combination of acoustic sources, with rotat-

establishing a limit on what can be achieved through

ing monopoles and dipoles representing thickness and refinement in hull and blade design.
loading noise respectively, and stationary sources representing contributions from unsteady volumes and

Introduction

fluctuating forces at the hub. The pressure field and


forces on the hull surface are calculated using an effi-

It has been usual to ignore the effects of fluid com-

cient Boundary Element code to solve the Helmholtz

pressibility in computation of fluctuating hull pres-

equation, taking fully into account the effects of com-

sure distributions due to propeller sources. These

pressibility. The effects of the free surface are also

sources arise from the rotation of a propeller in a

examined for various types of acoustic source using

spatially non-uniform and unsteady flow field. In ef-

image techniques. The selected body is a floating el-

fect, the hydrodynamic aspects of the problem have

lipsoid with the source beneath the surface. This is

been combined with the calculation of the unsteady

a preliminary representation of surface ship hull ex-

pressure field near to the propeller, assuming incom-

citation by general acoustic sources.

pressible flow apart from unsteady voids associated

We show considerable differences between the effects


of stationary and rotating sources. We also find that

with propeller cavitation (Cox et al 1978, Breslin and


Andersen 1994).

the contribution of higher harmonics from a rotat-

In reality, rotating or unsteady sources at the pro-

ing source decays fairly rapidly when the hull is fully

peller cause both near and far-field pressure distri-

submerged, but remains significant in the case of a

butions, the former associated with unsteady pres-

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

sure on the hull, the latter associated with underwa-

can, for example, estimate the magnitudes of the ro-

ter sound. The transition between the hydrodynamic

tating steady thrust and fluctuating thrust that will

and acoustic fields due to fluctuating hub forces oc-

have similar effects on near or far field sound pressure

curs close to the propeller at frequencies of practical


interest for computation of hull vibration response

levels.
We have compared the effects of four fundamentally

(Kinns and Bloor 2004).

different types of acoustic source. These are: stationary monopoles, stationary dipoles, an array of rotat-

Complete solution of the combined hydrodynamic

ing volume sources and an array of rotating steady

and acoustic problems can be achieved in principle

forces centred on the propeller hub. Results are pre-

by solving the Ffowcs Williams-Hawkings equation

sented in non-dimensional form where the ratio of the


sound wavelength to the body diameter is a principal

(Salvatore and Ianniello 2002). The numerical problems are challenging and it is difficult to check the

non-dimensional parameter. This allows the results

accuracy of derived solutions for realistic geometries

to be compared with those from earlier work where

and flow conditions. It therefore remains important

the fluid is incompressible and the sound wavelength

to develop a full understanding of the effects of spe-

is infinite (Chertock 1965).

cific types of acoustic source at the propeller on hull


excitation and underwater radiated noise.

The effects of the free surface are examined for various types of acoustic source using image techniques.

In previous work (Kinns and Bloor 2002, 2004), the

We do not model the propeller explicitly, regarding

nature of hull vibration excitation by stationary fluc-

it instead as a combination of acoustic sources lying

tuating volume and fluctuating force sources at the

in the plane of a notional propeller.

propeller location of a modern cruise liner was explored. The Helmholtz equation was solved using

boundary element modelling techniques to include


the effects of a finite speed of sound, for specified

The acoustic field of a propeller source

acoustic sources. We then showed (Kinns et al 2002,


Rath Spivack et al 2004) how the fluctuating pres-

The acoustic field of a static source obeys the inho-

sure and force distributions on submerged and float-

mogeneous wave equation:

ing bodies are influenced by the combined frequency-

1 2p
2 p = Q(x) .
c2 t2

dependent effects of diffraction around the body and

(1)

of the free surface.


If the source is a monopole with frequency , the

We have now extended earlier work to include the effects of rotating steady forces and steady volumes, so

source field is given by:

that it is possible to compare the effects of stationpi (r, t) = M0 eit

ary fluctuating and rotating steady components. We


2

Copyright National Academy of Sciences. All rights reserved.

eir
,
r

Twenty-Fifth Symposium on Naval Hydrodynamics

where M0 is the monopole strength.

rotates at the propeller frequency . The cylindrical

If the source is a dipole, the source field is given by:

coordinates (r, , z) are defined as follows:

it

pi (r, t) = D0 e

i
1
kr

r = radial distance from propellers axis,

eikr
cos
,
r

= angle of rotation around propellers axis,


z = propellers axis.

where is the angle between the dipole axis and r,

This gives the acoustic pressure as a sum of harmon-

and D0 is the dipole strength.

ics:

An acoustic source in motion produces a field that to

an observer will appear both shifted in frequency and


pi (x, t) =

with a different amplitude from the field radiated by

eimt eim Pm (x).

(5)

m=

a stationary source.
The particular cases when the dipole axis is aligned

The pressure perturbation due to a dipole q (of con-

parallel or perpendicular to the propeller axis repre-

stant strength) located at a moving point x = xs ( )

sent thrust and torque respectively.

obeys the time-dependent equation (see, for example,

The equation obeyed by the acoustic pressure due to

Dowling and Ffowcs Williams 1983):

a rotating monopole of constant strength Q located

1 p
2 p = q(x xs ( )) .
c2 t2

at a moving point x = xs ( ) is :

(2)

p
1 2p
2 p =
(Q(x xs ( ))
2
2
c t
t

By using the free-space Greens function


G(x, y) =

(t | x y | /c)
,
4 | x y |

(6)

(3) (see e.g. Crighton et al 1992).


As in the case of the rotating dipole, the acoustic

and denoting the cylindrical coordinates of the ob-

pressure is again decomposed into individual harmon-

server at x by R, , Z, then, for an axial dipole, the

ics. In the assumption of chordwise compact sources,

following expression is derived:

i.e. blade chord cL << , which applies in all the


cases of interest here, the term for the mth harmonic

pi (x, t) =

qz X imBtim 2 eim(BS/c)
is approximated by:
e
d ,
2
2
8 m=
S
0
(4)

Pm =

where

m2 B 3 2 0
iMx 2
(1 +
) Im ,
8 2
mBMt z

(7)

where
S=

|x

|2

2aR cos .

Here we have assumed that the source is located at


a fixed radius a in the propeller plane at z = 0 and

xtip

Im =
x0

hcL eimB(+tMt )
ddx
S0

Copyright National Academy of Sciences. All rights reserved.

(8)

Twenty-Fifth Symposium on Naval Hydrodynamics

and the distances


anm =

p
1
S 0 = | x |2 2aR cos
lB

M X
N
X

Anmij bij

(11)

j=1 i=1

where anm represent the known incident field at zn ,


0 =

m , b the unknown surface field, and Anmij represents

Mx (z z1 ) + S 1
(1 Mx2 )1/2 lB

the discretisation of the rest of the integral in 10.


This defines a set of (M N ) equations, which can

are normalised by the blade length lB . Mx is the

be written as a matrix equation,

axial Mach number, Mt is the tip Mach number, and


h is maximum blade thickness (see Peake 1991).

c = Bb .

(12)

By making the further approximation Mx 0, valid


for an underwater propeller in the normal operational

Finding the solution vector b requires inverting the

range of submarine and commercial vessels, the har-

(M N ) (M N ) matrix B, whose elements are given

monic term simplifies further to

by Anmnj . The solution was found to be stable with


respect to change in discretisation, which is deter-

Pm

m2 B 3 2 0
=
8 2

xtip

x0

Z
0

mined by wavelength and by geometry. We achieved

hceimB(Mt S )
ddx.
S0

good accuracy with 80 32 points. The code (Peake


and Rath Spivack 2002) was fast, running in about

(9)

3 minutes on a standard 600MHz PC. Computation

Once the incident field is given, then the total field

time rises rapidly with N and M , approximately

can be found by solving the wave equation. Since the

O(N 3 M 3 ).

source is harmonic, calculations are carried out in the


frequency domain. Using the appropriate incident
field for each type of acoustic source, as given above,

the total field p(r) on the hull surface is found by


solving the integral form of the Helmholtz equation:
Z
pinc (r) = p(r)

G(r, r0 )
p(r0 )dr0
n

Typical Fluctuating and Rotating Sources

The rotating volume and force sources are a nec(10) essary feature of thrust generation by a propeller.
They arise even if the flow field at entry to the pro-

where G(r, r ) is the free space Greens function,

peller is perfectly uniform and provide a lower bound

This is carried out by discretising the surface as a

on achievable hull excitation. The unsteady forces

function of the coordinates z and , so that at each are caused by imperfections in the wake field, which
point zn , m the discretised form of 10 is given by

the ship or submarine designer can seek to minimise


4

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

by careful attention to the combination of hull, ap-

the vertical, transverse and axial directions are often

pendage and propeller design.

The same applies

found to be about 0.01T at blade passing frequency

to unsteady volume sources that are introduced by

bpf = 2B, in the case of a surface ship with con-

propeller cavitation. Our interest is in establishing

ventional shafts and propellers. These may be re-

whether there is any potential conflict between design duced by an order of magnitude for a submarine or a
for minimum fluctuation and design for minimum ra-

surface ship with tractor propellers.

diated noise. It is also in establishing how far it is

All results presented here have been obtained for

useful to reduce fluctuating forces.

axisymmetric hull shapes. Calculations have been

In deriving pressure fields due to rotating forces and carried out for acoustic wavelengths in the range
0.5d < < 10d. In the case of a 100 metre long

volumes, we have assumed that these sources are dis-

tributed equally between B blades and that both can body, the frequency is then in the range from 7.5 Hz
to 150 Hz, assuming that the speed of underwater

be regarded as a circumferential array of B point

sources at 0.7rp , where rp is the propeller radius. sound is 1,500 metres/sec. This includes representaTherefore, they have a circumferential velocity of tive values of propeller blade passing frequency and
0.7rp , where is the rotational frequency of the

its multiples.

propeller: = 2N , with N denoting the number of

shaft revolutions per second.

The Hull Model

In order to compare the typical influence of rotating


The principal part of our analysis is for a simple el-

forces and volumes, we have assumed that the pro-

lipsoidal body whose length to diameter ratio is 5.

peller has a thrust coefficient KT = 0.2, where KT is

The body is either submerged in an infinite sea or

defined by:

floating with the free surface of the sea in the plane


of symmetry. Our chosen length to diameter ratio is

KT = 4 2 T /(2 d4p ) ,

less than that of a typical submarine hull or a surface


where T = total thrust, distributed between B ship hull at the waterline. Real hulls tend to have
blades, = density of water, dp = 2rp .
an extended region of nearly constant beam as well
The total volume of the propeller blades Vp is assumed to have the typical value Vp =

0.1rp3 ,

as lines that are fuller near the stern, so disturbing

dis-

forces determined for the ellipsoid of a specified beam

tributed between B blades.

are likely to be increased for a real hull shape.

The proportional volume and thrust coefficient might

The aspect ratio is chosen to give the best approx-

be expected to vary in practice by no more than a


factor of 2 from these nominal values.

imation to the curvature of a real hull in the region where the source-related pressure on the hull

The fluctuating forces are assumed to act at the pro-

is greatest. This means, in practice, a section of a

peller hub. The rms fluctuating forces F in each of few wavelengths downstream of the propeller. Some
5

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

results are shown for different aspect ratios, to illus-

sion of fluctuating forces via the propeller shaft, but

trate how surface pressures change as the hull changes

with the associated pressure fields that drive the hull

from nearly spherical to very elongated.

surface. The effects of unsteady cavitation are rep-

The chosen simplifications are designed to illustrate

resented by a fluctuating volume source at the hub.

the principal characteristics of each type of acoustic

The layout of the body and propeller sources is shown

source, but a similar analysis can be carried out for in Figure 1.


any combination of source type and position relative
to any chosen shape of body. The hull is assumed

1.0

to be rigid in this analysis, so we avoid having to


y /d

describe the dynamic properties of the hull and focus


instead on the incident and scattered propeller fields

0.0

in the absence of induced hull vibration.

-1.0
-1.0

1.0

0.0

1.0

2.0

3.0

4.0

5.0

6.0

y /d

x/d
0.0

-1.0
-1.0

hull

propeller

sea surface

Figure 2
0.0

1.0

2.0

3.0

4.0

5.0

In the case of the floating ellipsoid, we position the

6.0

notional propeller sources beneath the hull, as shown

x/d
hull

propeller

in Figure 2. The propeller diameter in this case has

axis

length dp = d/5. The propeller centre is d/6 below the hull, and d/4 forward of the stern waterline.
Figure 1

The notional clearance is then dp /3. The fluctuat-

In the case of the submerged body, the sources are in ing forces and fluctuating volume source, representa plane d/4 downstream of the ellipsoid extremity, on ing unsteady cavitation, act at the propeller hub.
the ellipsoid principal axis. The propeller is assumed
to have a number of blades B between 5 and 20. The
propeller diameter is dp = d/2.

Fluctuating forces, representing the effects of tempo-

Scattering by submerged ellipsoidal hulls

ral and spatial variations in the wake field at entry


to the propeller, are assumed to act at the propeller We present here calculated pressures and forces on
hub. Our present concern is not with the transmis-

ellipsoidal hulls, induced by acoustic sources with fre6

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

quency 1 , associated with a wavelength of underwa- ing noise, the axial cumulative force shown is the roter sound of 10d for a 5-bladed propeller. On a 100 m

tating component at multiples of bpf, which leads to

hull, this corresponds to 7.5 Hz. For a 20-bladed pro-

pressure variations at a fixed point on the hull. For

peller, 1 then correspond to 30 Hz. Results with the stationary monopole and dipole sources, which
higher frequencies (Peake and Rath Spivack 2003)

represent contributions from steady volumes and fluc-

show similar trends.

tuating forces at the hub, the graphs show total axial

Using the values given in section 3, we first fix the

cumulative force.

geometry of the propeller blades. We also fix the frequency of the underwater source, as above, and we
take the density of water to be constant, with value
103 Kg/m3 . This determines the source strength of
the rotating monopole, which represents rotating volumes. The total thrust T is then determined by taking the thrust coefficient KT = 0.2. The strength of
the third source considered, the static dipole, is determined, in turn, by taking the fluctuating forces to
be three orders of magnitude (exactly) smaller than
the rotating steady thrust. We present the hull forces
generated by these sources in units of the total thrust
Figure 3a

T.
Hull forces produced by unsteady volume sources,
represented by stationary monopoles, which arise
from propeller cavitation, are quite different in nature and not directly related to the total thrust. We
show these in units of M0 d, where M0 is the rate
of change of mass flux, and d is the reference length
scale.
Figures 3a, b, c and d show the scaled axial cumulative force on the hull due to a rotating monopole
source, a rotating (thrust) dipole, a stationary
monopole and a stationary dipole source respectively,
in the case where the propeller is 5-bladed. For the

Figure 3b

first two sources, which represent thickness and load7

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

force on the hull decays quite fast with higher harmonics in the case of rotating sources, while in the
case of a stationary dipole the decay with higher multiples of bpf is much smaller.
This is of course directly related to the decay in amplitude of the harmonic components of the incident
acoustic pressure Pm , given by equation (4) and (9)
for rotating dipoles and monopoles respectively. The
components at integers multiples of bpf for a stationary dipole do not show similar decay.
For rotating sources, when we consider the first harFigure 3c

monic only, the contribution from loading noise is


dominant with respect to that from thickness noise:
the difference is about two orders of magnitude.
For higher harmonics, though, the contribution from
thickness noise becomes comparable to that from
loading noise. The exact ratio of the contribution
from thickness and loading will of course vary with
blade geometry, and indeed we have obtained different values by changing maximum blade thickness and
chord (Peake and Rath Spivack 2003), but within
the parameter range appropriate for the normal operational conditions of commercial vessels and submarines the same general observation is valid: the

Figure 3d

contribution from loading noise is dominant for the

We can see that the major contribution to the axial

first harmonic. This applies generally in the subsonic

cumulative force on the hull comes from the rotating

regime.

thrust. A few immediate observations can also be


made regarding the behaviour of higher harmonics

The differences between a rotating and a stationary

for different type of sources.

the acoustic pressure along a section of the hull. We

source can also be seen by plotting the variation of

There are considerable differences in the scaling of show this in Figures 4a and 4b, where arbitrary lonthe amplitude of harmonics between rotating and sta-

gitudinal sections along the hull have been chosen.

tionary sources. The magnitude of the cumulative

The comparison here is between the result obtained


8

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

from the first harmonic component of a rotating axial

of bpf with the static source. For the higher harmon-

dipole and from a stationary dipole at bpf in Figure

ics, the pressures resulting from stationary sources

4a, and between the third harmonic and 3 bpf in have much greater amplitude than those resulting
Figure 4b. The modelled propeller is 5-bladed, with from rotating sources, because, as shown earlier, the
shaft frequency 1 . The pressure is shown in units of amplitude of the harmonic components decay very
T /d2 on a logarithmic scale.

fast in the case of a rotating source, but hardly at all


for higher multiples of bpf in the case of a stationary
source. The surface pressure resulting from the static
dipole is different in both behaviour and magnitude
from that resulting from a rotating dipole in the section of the hull closest to the propeller, then decays
much more slowly along the hull. These differences
increase slowly with higher number of blades, and
more markedly with higher harmonics.

This can be seen from Figures 5a and 5b, where the


same quantities are shown as in Figures 4a and 4b
respectively, but in the case of a 10-bladed propeller,

Figure 4a

i.e. doubling the number of blades.

Figure 4b
Similar differences occur also for higher harmonics of
the rotating source, compared with higher multiples

Figure 5a
9

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Figure 5b

Figure 6a

In Figures 6a,b and 6c,d we show the changes in the


axial cumulative force when the number of blades is
doubled from 5 to 10, and again to 20, for a rotating axial dipole and a stationary axial dipole respectively. In the case of the rotating dipole the two figures shown are for the first and third harmonic, in
the case of the stationary dipole they correspond to
bpf and 3 bpf . It is clear that the forces on the
hull generated by loading noise decrease significantly
Figure 6b

as the number of blades increase.


10

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

from different sources, though, appears to be strongly


dependent on the aspect ratio. This is an interesting
result, which will be explored further. We will also
aim to aim to establish whether any effects dependent
on the aspect ratio are enhanced by asymmetries in
the hull geometry, and particularly in the case where
appendages are present.

Figure 6c

Figure 7a

Figure 6d
The effect on the axial cumulative force of changing
the aspect ratio of the hull is shown for these two
sources in figures 7a and 7b, going from the extreme
case of a spherical hull (ratio = 1) to a very elongated
hull with ratio = 10. The qualitative behaviour of
the force along the hull does not vary as the ratio is
changed, with the exception of the spherical hull case.
The relative magnitude of the axial forces resulting

Figure 7b
11

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

Scattering by floating ellipsoidal hulls

fect of the sea surface in the case of rotating sources.

We include here some results obtained with a floating

case. It start to show appreciable decay only with

hull, to show the effect of the sea surface. We have

the fourth harmonic, which is still not small enough

We can see that the main effect is that the contribution from higher harmonic is quite significant in this

chosen representative results with different types of to be disregarded.


sources associated with a 5-bladed propeller (Figures
8a,b,c), and in figures 9a and b we show the effect of
increasing blade number in the cases when the source
is a stationary and a rotating axial dipole respectively.

Figure 8b

Figure 8a
In order to make a comparison with the results shown
in the previous section for a submerged hull, we show
again graphs of the first to fourth harmonic components of the axial cumulative force. In fact, in the
case of a floating hull the rotating axial dipole source
will give rise to a net fluctuating force on the hull,
which is not shown here.
The axial cumulative force on the hull resulting from
loading noise due to a 5-bladed propeller, shown in
Figure 8a, is given as a typical example of the ef-

Figure 8c
12

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

to model thickness and loading noise . Stationary


dipole sources have been used to model the effect of
a net fluctuating tailshaft force at chosen multiples
of bpf .
The effect of these sources on a hull has been explored
by calculating pressures and forces for different propeller geometries and varying the aspect ratio of the
hull. Realistic geometrical parameters, frequencies,
and source amplitudes have been employed throughout, giving new important results regarding the ef-

Figure 9a

fect of different forces on a submarine hull and their


ratio. More accurate quantitative results should be
achieved using actual shapes and values.
Here we have shown that a traditional hydrodynamic
analysis tends to underestimate the pressures on the
hull surface and therefore the effects of the external
pressure fields on hull excitation. We have also provided quantitative results regarding the effects of the
rotating forces and volumes, and the effects of the
net fluctuating tailshaft force. The ratio of these effects is particularly important in establishing a limit
on what can be achieved through refinement in hull
and blade design.
Our results confirm that only the first few harmonic

Figure 9b

components of the acoustic pressure are likely to provide the most significant contribution to the vibra-

tional response of the hull due to a rotating pro-

Conclusions

peller source. They also show that the contribution

We have used stationary and rotating source to model

of higher harmonics varies considerably for different

non-cavitating sources associated with a propeller,

types of sources.

for a range of parameters of interest to existing ves-

Finally, our Boundary Element code has proved fast

sels.

enough to run for up to 7000 elements in well under

Rotating monopole and dipole sources have been used 10 minutes on a fast Pentium 4 machine, and flexi13

Copyright National Academy of Sciences. All rights reserved.

Twenty-Fifth Symposium on Naval Hydrodynamics

ble enough to allow input from a variety of different

Dowling

sources and of different hull shapes, thus providing a

Sound and sources of sound Ellis Horwood 1983

very useful tool for comprehensive parametric stud-

Kinns,

ies. The flexibility of the code together with its very

Hull

high limit on the number of elements allows to input

RINA Transactions Part A, vol.

the hull shape from CAD data structures without sig-

41-70.

nificant loss of information, so acoustic pressures and


forces on real hulls can be accurately modelled.

A.P.

R.
Forces

and

and
due

Ffowcs

Bloor,
to

Williams

C.D.,

J.E.

Fluctuating

Propeller

Cavitation,

144, 2002, pp

Kinns, R. and Bloor, C.D., Hull Vibration Excitation due to Monopole and Dipole Propeller Sources,
Journal of Sound and Vibration, vol. 270, 2004, pp.

Acknowledgements

951-980.
Kinns,

The authors are grateful to BAE SYSTEMS, QinetiQ

R.,

Peake,

N. and Rath Spivack O.,

Hull Vibration Excitation by Propeller Sources:

and Dstl for their support and encouragement of the

a link between hydrodynamics and marine acous-

work described in this paper.

tics, Proceedings of the 24th Symposium on Naval


Hydrodynamics, Japan, Fukuoka, 2002.

References

Peake N. The Aeroacoustics of High-Speed ProBreslin


J.P.
and
Andersen
P.
Cambridge
Hydrodynamics of Ship Propellers,

pellers and Fans, PhD thesis, Cambridge, 1991


Peake N. and Rath Spivack O. unpublished data,

University Press, 1994.

2002
Chertock, G. Forces on a Submarine Hull Induced
by the Propeller Journal of Ship Research, Vol.9,
Sept. 1965, pp 122-130.

Peake N. and Rath Spivack O. unpublished data,


2003
Rath Spivack O., Kinns R. and Peake N Acoustic

Cox B.D., Vorus W.S., Breslin J.P. and Rood Excitation of Hull Surfaces by Propeller Sources,
E.P. Recent Theoretical and Experimental JMST, 2004, in print.
Developments

in

the

Prediction

Forces

on

of

Nearby

Propeller

Induced

Vibration

Bound-

aries,

Proceedings of 12th Symposium on Naval

Salvatore F. and Ianniello S. Preliminary Results on Acoustic Modelling of Cavitating Propellers, Proceedings of IABEM 2002, International

Hydrodynamics, 1978, pp 278-299.

Association for Boundary Element Methods,

Crighton D.G., Dowling A.P., Ffowcs Williams J.E.,

versity of Texas at Austin, May 2002.

Heckl M. and Leppington F.G. Modern Methods in


Analytical Acoustics, Lecture Notes, Springer Verlag, 1992
14

Copyright National Academy of Sciences. All rights reserved.

Uni-

Anda mungkin juga menyukai