Anda di halaman 1dari 312

Advanced Engineering Models for Wind Turbines with Application to the Design of a Coning Rotor Concept

Curran A. Crawford

Trinity College Department of Engineering

This dissertation is submitted for the degree of Doctor of Philosophy October 2006

Engineering is the science of economy, of conserving the energy, kinetic and potential, provided and stored up by nature for the use of man. It is the business of engineering to utilize this energy to the best advantage, so that there may be the least possible waste. Willard A. Smith (1908)

I, Curran Crawford, certify that this dissertation is the result of my own work and includes nothing which is the outcome of work done in collaboration except where specically indicated in the text. The thesis has a total word count of less than 65,000 and includes less than 150 gures.

A This manuscript was prepared using the L TEX2e system implemented in MiKTeX and worked on in TeXnicCenter. The diagrams were prepared in OpenOce.org Draw and the gures in Matlab and Excel.

The text is set in 11 pt font. Copyright c 2006 by Curran Crawford

Title: Advanced Engineering Models for Wind Turbines with Application to the Design of a Coning Rotor Concept Author: Curran A. Crawford Keywords: Wind Turbine, Coning Rotor, BEM Theory, Rotor Optimization

Abstract
The core focus of this thesis is the development of engineering-level aerodynamic and structural models, suitable for the design of highly exible wind turbines with large coning and yaw angles. The enhanced models are integrated to enable optimization of a coning rotor wind turbine concept. GH BLADEDTM is used as a reference industry-standard code, to evaluate the validity of current design tools applied to highly exible concepts. The coning rotor concept combines the load shedding properties of ap-hinged blades with gross change in rotor area, via large coning angles and lengthened blades, to achieve increased energy capture at nominally constant system cost. Based on up-scaling the original detailed design work from the mid-1990s, a comparison to a modern conventional machine indicates that the coning rotor remains a valid alternate technology track. Design theory is also used to justify the coning conceptual approach. The primary theoretical contribution of this work is an enhanced Blade Element Momentum (BEM) method. Utilizing vortex theory to model induction, computationally ecient corrections are derived that are key in more accurately predicting performance for coned rotors. The theory is extended to include wake expansion, dynamic inow, and yawed conditions, as well as considering centrifugal and radialow induced stall-delay. The theory is favourably validated against Computational Fluid Dynamics (CFD) and experimental results for both real and idealized rotors. BLADEDTM was to be modied with the enhanced BEM method for dynamic analyses. To support these analyses, a beam sectional model and Finite Element Method (FEM) approach to the generalized centrifugally stiened beam problem were implemented. Ultimately, the linear structural theory in current codes precluded accurate predictions at large ap angles. In lieu of a fully non-linear exiblebody simulation, a rigid-body dynamic model of the system was developed. The coupled aerodynamic and structural models were then used to analyse steady-state and dynamic operation, including optimal control schedules. Parametric optimization studies were used to examine the interplay between design variables for the coning rotor, relative to a reference conventional machine. Increased blade length, shape and airfoil choice were found to be tightly coupled, yielding energy gains of 1030% over conventional rotors. Airfoil choice and control mechanism were found critical to limiting torque and thrust. The fundamental nonlinear open-loop dynamics were also examined, including ap and edgewise damping behaviour. Low-Frequency Noise (LFN) was computed with a properly implemented physics-based model, to quantify sensitivity to design and operational parameters. The current work is a preliminary, but critical step, in proving the worth of the coning rotor. Controller design and an accurate exible-body code will be required for full load-set simulations, to aect detailed component design and costing. Ultimately, prototype testing will be needed to validate the complicated stalling behaviour of the coning rotor.

Acknowledgements
I would like to begin by thanking Jim Platts and Dr. Bill ONiel for their supervision throughout the process. Without Jims open what if? at our weekly meetings, none of the important questions would have been asked. The support given by members of Garrad Hassan, in terms of time, conversation and code (BLADEDTM ), is greatly acknowledged. In particular, Peter Jamiesons initial overview of the coning rotor concept was invaluable. David Quartons generous oer of BLADED enabled comparison to an advanced, industry-level code. Ervin Bossanyi, Graeme McCann, Tony Mercer, and Robert Rawlinson-Smith always responded quickly to e-mail queries, and their eorts to integrate my changes into BLADED are greatly appreciated. Thomas Wulf and Andrew Streateld at Rothe Erde provided valuable insight into very large bearing design and applicabilty to wind turbines. Robert Mikkelsen at the Technical University of Denmark provided valuable feedback and data for the development of the BEM method. Henk Polinder and Maxime Dubois from the Delft University of Technology provided data and insight into direct-drive generator costing. Dr. Coton at Glasgow University generously provided the source code for HAWTDAWG and helped in attempting to modify it to suit the coning rotor. The team at NREL in the US, especially Patrick Moriarty, Jason Jonkman and Scott Schreck, made available the UAE dataset, as well as insight into their codes and noise prediction. Paul Veers and Jose Zayas at Sandia National Labs also provided a North American perspective to wind energy. Funding by the UK Commonwealth and Canadian NSERC scholarships is gratefully acknowledged. On a more personal note, numerous people at Cambridge provided good company during the course of the PhD, especially Thomas, Andy, and Tams. Carlos proved a good friend at our meetings on both sides of a the Atlantic. I have also enjoyed spending time away from work with all those bound together by Portugal Street and Canadian friends from The Other Place. My family has, as ever, given me support and encouragement throughout my time at Cambridge. Finally, Rina has shared the whole process with me, making the endeavour truly worthwhile.

Table of Contents
1 Introduction 1.1 Wind Power . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.2 Scoping the Coning Rotor . . . . . . . . . . . . . . . . . . . . . . . 1.3 Thesis Outline . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 1 2 5

Wind Energy in Context


9 9 9 9 10 11 11 12 12 12 13 13 15 15 16 19 20 21 21 21 23 23 26 26 29 30 30 31

2 System Considerations 2.1 The Resource . . . . . . . . 2.1.1 Scales . . . . . . . . 2.1.2 Characterisation . . 2.1.3 Metrics . . . . . . . 2.2 The Grid . . . . . . . . . . 2.2.1 System Constraints . 2.2.2 Supporting the Grid 2.2.3 Decentralization . . 2.3 The End User . . . . . . . . 2.3.1 Locale . . . . . . . . 2.3.2 Further Impacts . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

3 Variations on a Theme 3.1 Chasing Function . . . . . . . . 3.1.1 Design Theory . . . . . 3.2 COE and CF . . . . . . . . . . 3.3 To Lift or Drag . . . . . . . . . 3.3.1 Governing Equations . . 3.3.2 Lift-Drag Comparison . 3.3.3 Operational Conditions 3.4 VAWT or HAWT . . . . . . . . 3.5 Scale . . . . . . . . . . . . . . . 3.6 Control Strategy . . . . . . . . 3.6.1 Goals and Methods . . . 3.6.2 Dynamic Considerations 3.7 Adapting Structures . . . . . . 3.7.1 Dynamic Motion . . . . 3.7.2 Load Matching . . . . .

. . . . . . . . . . . . . . . i

. . . . . . . . . . . . . . .

. . . . . . . . . . . . . . .

. . . . . . . . . . . . . . .

. . . . . . . . . . . . . . .

. . . . . . . . . . . . . . .

. . . . . . . . . . . . . . .

. . . . . . . . . . . . . . .

. . . . . . . . . . . . . . .

. . . . . . . . . . . . . . .

. . . . . . . . . . . . . . .

. . . . . . . . . . . . . . .

. . . . . . . . . . . . . . .

. . . . . . . . . . . . . . .

. . . . . . . . . . . . . . .

. . . . . . . . . . . . . . .

. . . . . . . . . . . . . . .

. . . . . . . . . . . . . . .

. . . . . . . . . . . . . . .

. . . . . . . . . . . . . . .

3.8

3.9

3.7.3 Soft Approach . . . . . . . . . . . . 3.7.4 Coning Rotor Adaptation . . . . . . Function to Form . . . . . . . . . . . . . . . 3.8.1 Historical Context . . . . . . . . . . 3.8.2 Topology . . . . . . . . . . . . . . . 3.8.3 Pitch Control . . . . . . . . . . . . . Coning Rotor Challenges and Opportunities

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

31 32 32 32 34 36 36

II

Analysis
41 42 42 45 45 47 47 49 53 54 57 58 60 64 72 73 75 76 77 96 97 100 103 109 114 114 119 124 130 136 137 141

4 Analytic Development 4.1 Industry Standard Comparison (BLADEDTM ) 4.2 Coordinate Systems . . . . . . . . . . . . . . . 4.3 Coned Rotor Performance Metrics . . . . . . . 4.4 Aerodynamic Modelling . . . . . . . . . . . . . 4.4.1 Back to Basics . . . . . . . . . . . . . . 4.4.2 Flow Field Kinematics . . . . . . . . . . 4.4.3 Momentum Balances . . . . . . . . . . . 4.4.4 Blade Elements . . . . . . . . . . . . . . 4.4.5 Wake Analysis . . . . . . . . . . . . . . 4.4.6 Relating Disc to Far Field . . . . . . . . 4.4.7 Azimuthal Induced Velocity . . . . . . . 4.4.8 Stall Delay and Centrifugal Pumping . . 4.4.9 Spanwise Flow . . . . . . . . . . . . . . 4.4.10 Closed Equations . . . . . . . . . . . . . 4.4.11 Expanding Wake . . . . . . . . . . . . . 4.4.12 Additional Wake Geometries . . . . . . 4.4.13 Aerodynamic Analytic Optimum . . . . 4.4.14 Dynamic BEM . . . . . . . . . . . . . . 4.5 Acoustic Modelling . . . . . . . . . . . . . . . . 4.5.1 Aerodynamic Noise . . . . . . . . . . . . 4.5.2 Modelling Approaches . . . . . . . . . . 4.5.3 Low-Frequency Noise . . . . . . . . . . . 4.5.4 Post-Processing . . . . . . . . . . . . . . 4.6 Structural Modelling . . . . . . . . . . . . . . . 4.6.1 Sectional Modelling . . . . . . . . . . . 4.6.2 Kinetostatics . . . . . . . . . . . . . . . 4.6.3 Dynamic Modelling . . . . . . . . . . . 4.6.4 Centrifugally Stiened Beam . . . . . . 4.7 Control . . . . . . . . . . . . . . . . . . . . . . 4.7.1 Steady State Operation . . . . . . . . . 4.7.2 Dynamic Control . . . . . . . . . . . . . ii

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

4.8

4.9

Generator Modelling . . . . . . . . . . . . . 4.8.1 Shear Stress Generator Model . . . . 4.8.2 Analytic Magnetic Circuit Generator Cost Modelling . . . . . . . . . . . . . . . . 4.9.1 Proxy Cost Metric . . . . . . . . . . 4.9.2 Magnetic Materials Cost . . . . . . .

. . . . . . . . Model . . . . . . . . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

141 142 143 143 144 144 147 147 148 149 153 155 156 160 164 165 167 169 171 172 177 178 178 179 181 182

5 Validation 5.1 Aerodynamic Validation . . . . . . . . . . . 5.1.1 Uniformly Loaded Rotor . . . . . . . 5.1.2 Uniform Loading Results . . . . . . 5.1.3 Fundamental Model Deciencies . . 5.1.4 Radial and Far-Field Flow . . . . . . 5.1.5 Tjaereborg Rotor . . . . . . . . . . . 5.1.6 NREL UAE . . . . . . . . . . . . . . 5.1.7 Dynamic Inow . . . . . . . . . . . . 5.1.8 Yawed Flow . . . . . . . . . . . . . . 5.2 Acoustics Validation . . . . . . . . . . . . . 5.3 Structural Validation . . . . . . . . . . . . . 5.3.1 Sectional Properties . . . . . . . . . 5.3.2 Beam Model . . . . . . . . . . . . . 5.4 BLADEDTM Validation and Suitability . . 5.4.1 Aerodynamics . . . . . . . . . . . . . 5.4.2 Steady State Operation . . . . . . . 5.4.3 Flapping Hinge . . . . . . . . . . . . 5.4.4 Mode Shape Modication by Coning 5.5 Generator Model Validation . . . . . . . . .

. . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . .

III

Design
185 185 185 186 187 188 190 191 191 193 194 195

6 Rotor Optimization 6.1 Initial Comparison and Updating . . . 6.1.1 Shifting Benchmarks . . . . . . 6.1.2 Up-Scaling CONE-450 Results 6.1.3 Parked Conditions . . . . . . . 6.2 COE and CF Performance . . . . . . . 6.3 Aerodynamic Behaviour . . . . . . . . 6.4 Rotor Optimization Challenge . . . . . 6.4.1 Conventional Rotors . . . . . . 6.4.2 Coning Rotor Optimization . . 6.4.3 CONE-450 Approach . . . . . . 6.5 Parametric Study . . . . . . . . . . . . iii

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

6.5.1 6.5.2 6.5.3 6.5.4 6.5.5 6.5.6 6.5.7

Design Variables and Parameters Optimizer Tuning . . . . . . . . Stepwise Optimization Approach Aerodynamic Optimum . . . . . Blade Length . . . . . . . . . . . Control Strategy Impacts . . . . Aerodynamic Uncertainty . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

197 199 201 203 203 208 211 217 217 217 222 229 230 231 234 237 238 239 241 241 242 243 245 A-1 B-1 C-1 C-1 C-4 C-7 C-10 C-12 C-12 D-1 E-1

7 Design Considerations 7.1 Dynamic Simulation . . . . . . . . 7.1.1 Fundamental Response . . . 7.1.2 Aerodynamic Damping . . 7.2 Tower Thump . . . . . . . . . . . . 7.2.1 Variation with Wind Speed 7.2.2 Wake Prole Inuence . . . 7.2.3 Mitigation by Oset . . . . 7.2.4 Control Strategy Eects . . 7.2.5 Dynamic Motion . . . . . . 7.2.6 Remaining Issues . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

8 Conclusions 8.1 Analytical Contributions . . . . . . . . . . . . . . . . . . . . . . . . 8.2 Design Renements . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.3 The Next Steps . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Bibliography A Geometric Transformations B Cross Section Analysis C Reference Machine Specications C.1 Tjaereborg Machine . . . . . . . . . . . C.2 NREL UAE Phase IV . . . . . . . . . . C.3 1.5 MW Reference Machine (REF-1500) C.4 GH Demo Machine . . . . . . . . . . . . C.5 CONE-450 Concept . . . . . . . . . . . C.6 Coning Rotor Study . . . . . . . . . . . D Historical Machines E Material Specications

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

iv

List of Figures
1.1 1.2 1.3 3.1 3.2 3.3 3.4 3.5 3.6 3.7 3.8 4.1 4.2 4.3 4.4 4.5 4.6 4.7 4.8 4.9 4.10 4.11 4.12 4.13 4.14 4.15 4.16 4.17 4.18 4.19 4.20 4.21 4.22 4.23 4.24 Example Danish concept HAWT machines . . . . . . . . . . . . . . . Example VAWT machines . . . . . . . . . . . . . . . . . . . . . . . . . Coning rotor illustration . . . . . . . . . . . . . . . . . . . . . . . . . . CP and CF for translating airfoils utilizing lift and drag . . . . . . Variation of CP and CF with translation direction for lift device Blade mass trend with rotor diameter . . . . . . . . . . . . . . . . Specic energy capture as a function of machine rating . . . . . . . Power curve regions . . . . . . . . . . . . . . . . . . . . . . . . . . Wind velocity and force vectors for various control strategies . . . Power control in CP and planes . . . . . . . . . . . . . . Coning rotor schematic layout . . . . . . . . . . . . . . . . . . . . . Coordinate systems . . . . . . . . . . . . . . . . . . . . . . . . . . Kinematics setup . . . . . . . . . . . . . . . . . . . . . . . . . . . Kinematics at rotor for single stream-tube . . . . . . . . . . . . . Flow relative to airfoil section . . . . . . . . . . . . . . . . . . . . Thrust CT,model model behaviour . . . . . . . . . . . . . . . . . . Wake decomposition into rings and laments . . . . . . . . . . . Vortex ring geometry . . . . . . . . . . . . . . . . . . . . . . . . . Velocity eld around vortex cylinder . . . . . . . . . . . . . . . . Axial, root and bound vortex laments . . . . . . . . . . . . . . . Azimuthal induction along blade . . . . . . . . . . . . . . . . . . Skewed ow velocity vectors and boundary layer proles . . . . . Eective sweep angle with coning for typical operating conditions Experimental 2D lift curves for varying yaw angle . . . . . . . . Wing section lift curves for 45 swept wing . . . . . . . . . . . . Stall delay behaviour for 2D sections in yawed ow . . . . . . . . Sectional data corrected for spanwise ow . . . . . . . . . . . . . Expanded wake prole . . . . . . . . . . . . . . . . . . . . . . . . Skewed wake integral . . . . . . . . . . . . . . . . . . . . . . . . . Grids for table . . . . . . . . . . . . . . . . . . . . . . . . . . . Skewed wake table transformation . . . . . . . . . . . . . . . . . Tower shadow model . . . . . . . . . . . . . . . . . . . . . . . . . Example total velocity tower shadow ow eld . . . . . . . . . . Dynamic wind vector decomposition . . . . . . . . . . . . . . . . Dynamic ow relative to airfoil section . . . . . . . . . . . . . . . v . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3 3 4 22 22 25 25 26 27 28 34 43 48 48 50 52 55 56 57 59 61 64 65 67 68 68 71 74 79 81 82 84 86 86 87

4.25 4.26 4.27 4.28 4.29 4.30 4.31 4.32 4.33 4.34 4.35 4.36 4.37 5.1 5.2 5.3 5.4 5.5 5.6 5.7 5.8 5.9 5.10 5.11 5.12 5.13 5.14 5.15 5.16 5.17 5.18 5.19 5.20 5.21 5.22 5.23 5.24 5.25 5.26 5.27

Dynamic stall parameters . . . . . . . . . . . . . . Dynamic stall hysteresis loop for cyclic . . . . . Section denition . . . . . . . . . . . . . . . . . . . Blade section triangulation . . . . . . . . . . . . . Denition of section coordinate systems . . . . . . Section layout for kinetostatics . . . . . . . . . . . Kinetostatic oset vectors . . . . . . . . . . . . . . Integrated loading . . . . . . . . . . . . . . . . . . Beam element . . . . . . . . . . . . . . . . . . . . . Strained bre . . . . . . . . . . . . . . . . . . . . . Element centrifugal force . . . . . . . . . . . . . . Operating point solution ow chart . . . . . . . . . Speed windows for optimal and limiting operation

. . . . . . . . . . . . .

. . . . . . . . . . . . .

. . . . . . . . . . . . .

. . . . . . . . . . . . .

. . . . . . . . . . . . .

. . . . . . . . . . . . .

. . . . . . . . . . . . .

. . . . . . . . . . . . .

. . . . . . . . . . . . .

. . . . . . . . . . . . .

. . . . . . . . . . . . .

90 91 115 117 118 120 120 123 130 133 134 139 140 149 150 151 151 152 155 157 158 159 160 161 163 164 166 168 169 170 170 171 171 173 174 174 175 175 177 179

Uniformly loaded axial induction ( = 0 , CT = 0.89) . . . . . . . . . Uniformly loaded axial induction as a function of loading CT ( = 0 ) Uniformly loaded axial induction ( = 20 , CT = 0.89) . . . . . . . . . Uniformly loaded axial induction ( = 20 , CT = 0.89) . . . . . . . . Uniformly loaded axial induction as a function of for Bladed thrust model (CT = 0.89, no wake expansion) . . . . . . . . . . . . . . . . . . Uniformly loaded radially induced velocity as a function of (CT = 0.89) Tjaereborg axial induction ( = 0 ) . . . . . . . . . . . . . . . . . . . Tjaereborg axial induction ( = 20 ) . . . . . . . . . . . . . . . . . . . Tjaereborg axial induction ( = 20 ) . . . . . . . . . . . . . . . . . . Tjaereborg aerodynamic loading as a function of . . . . . . . . . . . Power coecient map for rotor operating points as a function of . . NREL UAE Sequence S measurements and predictions . . . . . . . . . NREL UAE Sequence F measurements and predictions . . . . . . . . Tjaereborg root bending moments in yawed ow . . . . . . . . . . . . Simple isolated rotor acoustic prediction . . . . . . . . . . . . . . . . . SPL for upwind rotor with tower inuence for varying azimuthal step size . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Eect of tilt on section out-of-plane velocity . . . . . . . . . . . . . . . Eect of tilt on pressure signature for upwind rotor . . . . . . . . . . . Eect of tilt on downwind rotor SPL . . . . . . . . . . . . . . . . . . . Acoustic footprint of downwind rotor with tilt and shear . . . . . . . . Uniform blade Out-of-Plane (OP) rotating modes . . . . . . . . . . . . Uniform blade In-Plane (IP) rotating modes . . . . . . . . . . . . . . . Uniform blade OP rotating modes for teetered hub . . . . . . . . . . . Free-hub rotating uniform blade IP modes . . . . . . . . . . . . . . . . lhub = 0.5 free-hub rotating uniform blade with IP modes . . . . . . . Collective IP mode shapes for Demo blade . . . . . . . . . . . . . . . . BLADEDTM computed steady power curve for PTS . . . . . . . . . . vi

6.1 6.2 6.3 6.4 6.5 6.6 6.7 6.8 6.9

6.10 6.11 6.12 6.13 6.14 6.15 6.16 7.1 7.2 7.3 7.4 7.5 7.6 7.7 7.8 7.9 7.10 7.11 7.12 7.13 7.14 7.15 7.16 7.17 7.18

Operational rotor thrust curves for upscaled machines relative to REF1500 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Betz limit variable area rotor performance . . . . . . . . . . . . . . . . Industry machine performance . . . . . . . . . . . . . . . . . . . . . . Performance maps for actual tip radius . . . . . . . . . . . . . . . . . Performance maps for reference tip radius and control variable . . . . Chord and twist prole control . . . . . . . . . . . . . . . . . . . . . . Typical section layup . . . . . . . . . . . . . . . . . . . . . . . . . . . . Finite dierencing prediction for CP as a function of solution tolerance tol and step size h for chord and twist control points . . . . . . . . . . Power coecient CP , annual energy yield Eann and total blade mass variation with aerodynamic design tip speed design,aero and cone angle design,aero . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Blade chord proles for design,aero = 68 . . . . . . . . . . . . . . . . Relative energy yield and maximum thrust variation with blade length Stip and balance angle balance . . . . . . . . . . . . . . . . . . . . . . Energy yield relative to REF-1500 varying with mean wind speed and hub height . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Operational curves for optimal rotors at balance = 30 . . . . . . . . . Generator cost variation with diameter and torque/speed levels . . . . CP curves for = 0 and varying airfoil set . . . . . . . . . . . . . Variation in operational curves with stall delay models for set II rotors Flap angle solution expansion terms . . . . . . . . . . . . . . . . . . . Blade apping without aerodynamic contribution . . . . . . . . . . . . Hub loads for blade apping without aerodynamic contribution . . . . Dynamic situation with apping blades and rotating imbalance . . . . Logarithmic decrement d for ap hinge damping . . . . . . . . . . . . Damping constant variation with control strategy for 15% thick set I airfoil . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Damping constant variation with airfoil for 15% thick airfoils and PTS Variation of Angle of Attack (AOA) and total twist angles over wind speed range with airfoil choice (15% thick) for PTS . . . . . . . . . . . Damping constant variation with airfoil for 21% thick airfoils and PTS SPL variation with wind speed . . . . . . . . . . . . . . . . . . . . . . Acoustic spectrum variation with wind speed, including stall delay . . SPL variation with wake width . . . . . . . . . . . . . . . . . . . . . . Acoustic spectrum variation with wake width . . . . . . . . . . . . . . SPL variation with wake decit . . . . . . . . . . . . . . . . . . . . . . Acoustic spectrum variation with wake decit . . . . . . . . . . . . . . SPL variation with oset from tower . . . . . . . . . . . . . . . . . . . Acoustic spectrum variation with oset from tower . . . . . . . . . . . SPL variation with cone angle and observer azimuth angle . . . . . . . vii

186 188 189 190 191 197 200 201

204 205 206 207 209 210 211 215 218 219 220 221 223 226 227 228 228 230 231 232 232 233 233 234 235 236

7.19 Acoustic spectrum variation with cone angle . . . . . . . . . . . . . . 7.20 Footprint SPL changes with variation in cone angle . . . . . . . . . . . 7.21 Rotor thrust variation over azimuth for control strategy choice . . . . B.1 Section integral coordinate transform . . . . . . . . . . . . . . . . . . . C.1 C.2 C.3 C.4 C.5 C.6 Tjaereborg blade details . NREL UAE blade details REF-1500 details . . . . . Demo machine details . . CONE-450 details . . . . Airfoil coecients (NACA . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63-4XX) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

236 237 238 B-1 C-2 C-6 C-8 C-11 C-13 C-14

viii

List of Tables
2.1 3.1 4.1 4.2 5.1 5.2 5.3 5.4 5.5 6.1 6.2 7.1 7.2 C.1 C.2 C.3 C.4 C.5 IEC Wind Turbine Classes . . . . . . . . . . . . . . . . . . . . . . . . 11 19 112 145 152 169 176 177 181 200 200 238 239 C-1 C-5 C-7 C-10 C-12 D-2 E-2 Axiomatic/TRIZ design method similarities . . . . . . . . . . . . . . . Windowing functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . Active magnetic specic material costs . . . . . . . . . . . . . . . . . . Uniformly loaded power coecients (CT = 0.89) . . . . . . . . . . Upwind SPL predictions . . . . . . . . . . . . . . . . . . . . . . . . Uniform blade with long hub eigenfrequencies . . . . . . . . . . . . Collective IP eigenfrequencies for Demo blade . . . . . . . . . . . . Steady ap angle variation with actuator moment for =50 RPM . . . . . . . . . .

Outer skin layup schedule . . . . . . . . . . . . . . . . . . . . . . . . . Web layup schedule . . . . . . . . . . . . . . . . . . . . . . . . . . . . SPL variation with control strategy above rated . . . . . . . . . . . . . SPL variation with dynamic motion and cone angle . . . . . . . . . . . Tjaereborg parameters . . NREL UAE parameters . REF-1500 parameters . . Demo machine parameters CONE-450 parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

D.1 Soft machine history . . . . . . . . . . . . . . . . . . . . . . . . . . . . E.1 Section layup material specications . . . . . . . . . . . . . . . . . . .

ix

List of Acronyms
AMT AOA BEM BOP DV CF CFD CHP COE CS DES DFT DNS DOE DOF DP EOM FD FEM FFT FR FSS FW-H GDT GH HAWT HVDC IP LE LES LFN LVRT NA NPV NREL NS ODE

Axial Momentum Theory Angle of Attack Blade Element Momentum Balance of Plant Design Variable Capacity Factor Computational Fluid Dynamics Combined Heat and Power Cost of Energy Coordinate System Detached Eddy Simulation Discrete Fourier Transform Direct Numerical Simulation Design of Experiments Degrees of Freedom Design Parameter Equations of Motion Finite Dierence Finite Element Method Fast Fourier Transform Functional Requirement Fixed Speed Stall Ffowcs Williams-Hawkings Japanese General Design Theory Garrad Hassan and Partners Horizontal Axis Wind Turbine High Voltage Direct Current In-Plane Leading Edge Large Eddy Simulation Low-Frequency Noise Low Voltage Ride Through Neutral Axis Net Present Value National Renewable Energy Laboratory Navier-Stokes Ordinary Dierential Equation xi

OP PDE PDF PMG PTF PTS PV RANS SGSF SQP SPL TE TRIZ UAE URANS VAWT VSS WTC

Out-of-Plane Partial Dierential Equation Probability Density Function Permanent Magnet Generator Pitch to Fine Pitch to Stall Process Variable Reynolds-Averaged Navier-Stokes Savitzky-Golay Smoothing Filters Sequential Quadratic Programming Sound Pressure Level Trailing Edge Theory of Inventive Problem Solving Unsteady Aerodynamics Experiment Unsteady Reynolds-Averaged Navier-Stokes Vertical Axis Wind Turbine Variable Speed Stall Wind Turbine Company

xii

List of Symbols
Given the range of equations covered, only the most generic and consistent are included here. The remaining symbols are dened in context as they are introduced and used. English Symbols a a Axial induction factor Tangential induction factor Aspect ratio Number of blades Chord length Sectional (2D) drag coecient Sectional (2D) lift coecient Sectional (2D) moment coecient Power coecient P/ 1/2V 3 D2 Thrust coecient T / 1/2V 2 D2 Mass Power; Rotation frequency and its multiples e.g. 1P , 3P Pressure Radius Reynolds number V l/ Spanwise coordinate, blade length Thrust Velocity (wind speed if no subscript specied) Induced velocities in momentum and vortex theories respectively

A
B c cd cl cm CP CT m P p r Re s, S T V w, u

Greek Symbols Airfoil AOA Cone angle (positive downwind) Aerodynamic skew angle
BEM correction factors; Strain

xiii

tot twist set pitch t yaw

Vortex lament strength Vortex strength per unit area Total blade section twist angle (all pitch angles positive towards feather) Blade section geometric twist angle Blade xed pitch oset (set angle) Active blade pitch angle Sweep angle Tip-speed ratio r/V Dynamic viscosity Kinematic viscosity / Rotor rotation speed Nacelle tilt angle Nacelle yaw angle Density Blade solidity Bc/2r; Stress

Notation xy xi xn A x I x x A set of 2D (or 3D) axes; Plane containing 2D axes Element index i; where i is an axis label, indicates vector component in that direction Time index n Matrix; also a b c d a b

Column vector; also Identity matrix

Geometric vector; also x, y Unit geometric vector

xiv

Chapter

Introduction
Energy lies at the heart of modern society, enabling everything from heating, lighting, computers, and food production through to manufacturing, medicine, and transport. To properly appreciate the context of this thesis, consider that worldwide in 2003, wind power provided 0.5% of the 13.5% slice electricty represented of the 125 PWh consumed at end-use from all primary sources [1, 2]. A growing appreciation of the deleterious climatic eects of anthropogenic greenhouse gas emissions is increasing regulatory pressure to mitigate or avoid CO2 release into the atmosphere [3]. Simultaneously, many populous nations are rapidly aspiring to Western standards of living, intensifying the use of and competition for energy. The tractability of tackling stationary energy use and supply, as opposed to transport fuels, and in particular electricity production, has made this area of key public importance [4]. However, it is important to bear in mind the overall scale of our societys energy use, in all forms and with myriad correctable ineciencies, when considering enhancement of electricity production. Ultimately, the highly competitive electricity market will select the most costeective technology. Unfortunately, inclusion of externalities, e.g. a carbon tax, unknown future decommissioning costs, as well as subsidies to all forms of energy, creates an articially skewed playing eld. To compete in this economic environment, wind power must deliver power of grid-standard quality and at a competitive price relative to conventional generation. It is in this overall context that this thesis seeks to develop engineering tools capable of properly analysing and optimizing a potentially much more cost-eective wind turbine concept, the coning rotor.

1.1 Wind Power


Wind power has for the past 5 years increased in installed capacity by 20-30% each year [2], indicating commercial viability with proper siting and regulatory framework. The turbines are among the best energy sources in terms of energy payback 1

Chapter 1 Introduction

time1 (3-7 months [5]) and ratio2 (1739 [6]). Once the turbines are built, their impact is virtually nil, given proper design and consideration of local communities and wildlife. These strong economic and sustainability credentials make wind power attractive from multiple perspectives. Wind turbine design has achieved a high level of sophistication, enabling production of 120 m diameter, 5 MW machines, sited both on-shore and o-shore. Progress from the 6 m diameter, 15 kW machines of the early 1980s to today has been achieved by technical and scale evolution of the Danish concept. Modern examples are shown in Fig. 1.1, typied by rigid three-bladed rotors placed upwind of the towers. In a modern machine, a gearbox transfers torque from the rotor to an actively controlled generator, although some directly drive a low-speed generator. Pitch actuators at the roots of the blades directly control the aerodynamic power input to the rotor. Almost universally, a Horizontal Axis Wind Turbine (HAWT) is chosen (i.e. the rotation axis of the rotor is near horizontal), although the Vertical Axis Wind Turbine (VAWT) (examples shown in Fig. 1.2) with a vertical rotation axis received much attention in the past, and may still nd application at extremely small and large scales. Key modern development work is aimed at increased reliability, particularly for o-shore sites, and load mitigation to reduce structural cost. The latter is being tackled by more active control (in turn made possible by power electronics and fast actuators) and rened structural designs and materials. It should be appreciated that wind turbines, although sharing many engineering principles with aircraft design, are quite distinct in their objective (to deliver reliable energy) and therefore demand more cost-sensitive solutions.

1.2 Scoping the Coning Rotor


The coning rotor was studied in the mid-1990s as an alternative technology track [7], potentially capable of delivering signicant cost of energy reductions relative to then state-of-the-art machines. The coning rotor builds on a lineage of exible machine designs, as opposed to the rigid Danish concept, in an attempt to both capture more energy and achieve a lighter (cheaper) structure. The essential physical features of the coning rotor, illustrated in Fig. 1.3, are hinges at the blade roots,
1 Energy payback time represents the time taken to produce as much energy as used in production of the generating device. 2 Energy payback ratio is the lifetime energy produced divided by energy used to produce the technology.

1.2 Scoping the Coning Rotor

(a) 1.3 MW Bonus machines at Lambrigg wind farm

(b) REpower 5 MW HAWT being installed oshore near Beatrice oil platform in the North Sea (Image from http://www.repower.de))

Figure 1.1 Example Danish concept HAWT machines

(a) 6 kW XC02 Quiet Revolution QR5 VAWT (Image from www.quietrevolution.co.uk)

(b) 300 kW Flowind Darrieus-type VAWT (Image from www.ecopowerusa.com)

Figure 1.2 Example VAWT machines

Chapter 1 Introduction

enabling the blades to ap downwind, thus sweeping out a cone. The rotor must obviously be downwind of the tower. The hinges aord a steady load-matching advantage by turning the blades from bending beams into tension members, as the centrifugal and aerodynamic forces balance each other.

Figure 1.3 Coning rotor illustration

Dynamic movement of the blade may also further mitigate transitory loading. Flap-hinged blades have long been known to aord this benet [810]. The dierentiating features of the coning rotor are larger operational and parked coning angles. In shut-down, extreme wind conditions, the blades are fully-coned to approximately 85 . By avoiding extreme loading, the blades can be relatively longer, capturing more energy. Even with these positive aspects, exible concepts have not achieved commercial success, particularly at scales greater than a few hundred kilowatts. The root diculty has been in achieving accurate models, both aerodynamically and struc-

1.3 Thesis Outline

turally.1 These are required to adequately size components and synthesize conservative operational proles. The inherent exibility of conventional rigid machines is a diculty in itself, without the further non-linearities of gross-coning. Failures have been typically caused by dynamic behaviour that was not accurately predicted, making a clear understanding of underlying assumptions important for all types of wind turbines. Numerous areas of study present themselves for further validating the coning rotor concept. This thesis is limited to enhancing the engineering models that will underpin any future development work. Without properly formulated tools, any more detailed analysis and design is premature. In particular, the novel aspect of the concept, the rotor itself, is of primary interest, with due consideration given to the capabilities of other components. Furthermore, technical analysis is pursued, rather than economic estimates. Having developed the requisite analysis suite, a parametric study is made of the concept, focusing on blade design and operational strategy. This step represents the most optimistic view of performance benet relative to traditional machines, and must be completed in order to justify and motivate future work. This rst stage is required in any case as a base for future dynamic simulations and further renement of the design loads. Although the focus is on the coning rotor, the developed theory and optimization procedure have relevance to current conventional machines, operating at large yaw angles with increasingly exible blades.

1.3 Thesis Outline


This thesis is divided into three distinct parts, following a natural progression from problem denition to solution. Part I deals with the fundamental aspects of harvesting wind energy. Chapter 2 outlines the relevant characteristics of the wind resource, together with the constraints placed on the wind energy converter by the energy system being supplied. Chapter 3 provides a fundamental overview of possible functional and physical approaches, based on the set of constraints imposed by the overall energy system. It also more clearly denes the operational and physical aspects of the coning rotor, justifying the concept from a design theory perspective, relative to other conceptual proposals.
There have of course been cases in which elements of the machine were designed incorrectly for nancial or expediency reasons.
1

Chapter 1 Introduction Part II presents the analysis tools required to explore highly exible wind turbine

concepts, motivated by the analysis requirements created by the physical aspects of the coning rotor in Chapter 3. Chapter 4 presents the theories that have been developed and implemented in a unied analysis suite, ExcelBEM. In particular, a novel Blade Element Momentum (BEM) method is presented, incorporating corrections to more correctly model coned and yawed rotors in steady and unsteady ow. The implementation of a Low-Frequency Noise (LFN) model is presented next. A structural model for the blade sections is then developed, followed by dynamic Equations of Motion (EOM) and a centrifugally stiened Finite Element Method (FEM) beam model. Finally, an optimal control scheduling algorithm is described, followed by a brief reference to a generator model and cost metrics. The work carried out to validate the various theories, to the extent possible, is presented in Chapter 5. The BEM method developments draw on the authors published material [11]. The remaining limits and shortcomings of the analytic methods are discussed, as is the suitability of currently available industry codes. It should be noted at this point that the reference code, BLADEDTM , is clearly capable of its intended use in simulating conventional machines; it is the suitability for unconventional concepts such as the coning rotor that is being explored. Using the analytic tools developed in Part II, Part III explores the fundamentals of designing a coning rotor, including aspects of the authors published papers [12, 13]. Chapter 6 rst up-scales the results of the original CONE-450 study to a modern MW-scale machine. The aerodynamic characteristics of the coning rotor are then discussed, with relevance to the optimization of rotors. A set of optimization and parametric studies of coning rotors is then presented and discussed. The control strategy, airfoil choice and blade length are primary foci. Chapter 7 then explores some critical elements for the further design of the coning rotor. The fundamental dynamic response and aerodynamic damping are presented rst, followed by a parametric study of LFN trends with design parameters. The thesis culminates with Chapter 8, by highlighting the main outcomes of this work, covering both the theoretical contributions to wind turbines in general, and the specic coning rotor concept study. The thesis concludes with suggestions for future work to be carried out to address the outstanding issues and move the concept forward towards reality.

Part

Wind Energy in Context

Chapter

System Considerations
This chapter gives only the most brief of introductions to the source, transmission and end-use of wind energy. The emphasis is on the salient points which must be considered when proposing and analysing an alternative wind turbine concept. Numerous authors inform the following sections, and provide more in-depth discussions of the facets presented [1419]. The resource itself is introduced in 2.1, followed by transmission via the grid in 2.2 to the end user in 2.3.

2.1 The Resource


Wind energy is actually a form of solar energy, ultimately originating from the nuclear reactions in the sun. Solar radiation dierentially heats the atmosphere and earths surface. This creates unbalanced pressure forces that drive the kinetic motion of the air.

2.1.1 Scales
Wind energy is present at three spatial scales: global (e.g. trade winds); secondary scales impacting the lower atmosphere (e.g. hurricanes, monsoon circulation); and tertiary localized winds (e.g. land/sea breezes, valley/mountain winds, thunderstorms, tornadoes). Temporal variation exists on an inter-annual, annual, diurnal (day/night) and short term basis.

2.1.2 Characterisation
An atmospheric boundary layer exists above the surface of the earth, so that the wind-speed grows with height z above reference height zref and wind-speed Vref : V = Vref roughness. 9 z zref

(2.1.1)

The power-law exponent is typically 1/7, varying with wind conditions and surface

10

Chapter 2 System Considerations

The annual temporal variation is described well by a Weibull distribution with Probability Density Function (PDF) f (V ): f (V ) = dF k = k V k1 exp dV c V c
k

(2.1.2)

The shape factor k is in the range 1.82.2, and the scaling parameter c is used to t the mean wind-speed. The specic power content of the wind is 1/2V 3 . This cubic relationship with velocity V means that higher wind speeds deliver much more energy than low speeds. Total energy yield from a turbine with power output function P = P (V ) is computed from:
Vco

E = 8760
Vci

P (V )f (V )dV

(2.1.3)

where Vci and Vco are the lower and upper bounds of the wind turbines operational wind speeds. The 8760 = 24 365 constant in the formula requires V in units of m/s and P in kW to yield E in kWh, the standard energy unit in the electricity industry. More detailed temporal and spatial description of the wind-speed is provided by the following [17]: Turbulence intensity Probability density function Autocorrelation function Integral time Power spectral density function Oshore wind conditions are generally favourable compared to those on-shore. The relatively smooth and unobstructed surface increases the shear rate (i.e. higher speeds at lower heights) and reduces turbulence levels.

2.1.3 Metrics
Potential installation sites are characterised by their annual mean wind-speed. The standard at 10 m height ranges from Class I (0.04.4 m/s) through to Class VII (7.09.4 m/s). Wind energy cost is typically quoted for Class VI sites. Bear in mind that the V 3 relationship means that that a Class VII site has 50X more power than a Class I site. Accurate knowledge of local conditions is critical for delivering veracious energy yield predictions. For the purposes of loads calculation, the IEC standard [20] species four classes of wind speed shown in Table 2.1, dened at the hub height.

2.2 The Grid Table 2.1 IEC Wind Turbine Classes (Adapted from [20]) Item Reference wind speed Annual average wind speed 50 yr return gust speed 1 yr return gust speed Uref Uave 1.4 Uref 1.05 Uref I 50 10 70 52.5 Class II III 42.5 8.5 59.5 44.6 37.5 7.5 52.5 39.4 IV 30 6 42 31.5 (m/s) (m/s) (m/s) (m/s)

11

2.2 The Grid


The electricity grid interconnects generators and consumers, often across municipal, state and national boundaries. It has evolved to facilitate large central generating stations exploiting economies of scale. New transmission technologies, such as High Voltage Direct Current (HVDC) [21], are permitting increasing transmission distances. A mix of generation technologies is used to constantly adjust power to match the demand load. The time-constant of the generator, whether it be baseload (e.g. coal, nuclear) or quick load-following (e.g. natural gas), determines the relative economics and hence dispatch choice.

2.2.1 System Constraints


The market structure can play a dominant role in wind powers economic success. The closer the scheduling of supply is to actual dispatch, the more accurate the forecast and hence system value. The Capacity Factor (CF) (see 3.2) of the machine will also aect the ecient utilization of transmission resources. Geography has a large impact as well, as transmitting electricity long distances is quite costly. Moving oshore only increases this burden. Unfortunately, wind, as with many renewables, has a low power density and is often best captured far from load centres. Wind power is often ascribed a low value in displacing conventional plant, and described as intermittent. While it is true that the wind does not blow all the time, nationally distributed wind-power plants will tend to smooth output uctuations. It is more accurate to say that wind power is variable, but then so to is electricity demand. Detailed studies and recent experience has indicated that up to a 20% market share of electricity generation, little backup capacity is required [22].

12

Chapter 2 System Considerations

2.2.2 Supporting the Grid


Wind turbines have been grid-connected successfully for many years [23]. In the past, simple induction generators were used to passively soft-couple the rotor to the grid frequency. At very low penetrations of wind power, and for smaller scale machines, this approach was, and is quite eective. For structural and power-capture reasons, machines are increasingly variablespeed, further decoupling the grid and rotor. With increasing levels of wind plant on the grid, regulators have toughened the requirements for connection. In particular, Low Voltage Ride Through (LVRT) is a requirement in many localities, as is the ability to provide reactive power to actively control the power factor. This has placed increasing demands on power electronics and generators to supply these services to the grid, either at the individual machine or overall wind farm level.

2.2.3 Decentralization
There has been recent interest in returning to a distributed generation model, as it was at the dawn of the electried age in the early 1900s. Savings accruing to Combined Heat and Power (CHP), including the costs to some users associated with network outages, have prompted fossil-fuel based machines of modest scale to be installed. Embedding generation within the grid network of course poses its own set of technical and regulatory challenges. It has been suggested that given the inherently distributed nature of renewables, it makes sense to pursue a distributed model [24]. At the same time, the scale eects of wind turbines must be born in mind when considering small installations, including the quality of the wind resource which grows with height (i.e. size of machine). Until such time as a cost-eective energy storage solution (e.g. compressed air, ywheels, batteries, hydrogen [25, 26]) is available, the grid will be required to redistribute power to continually satisfy demand regardless of wind resource. Nevertheless, the social engagement with the power system generated by embedded generation may in itself be benecial in terms of demand management.

2.3 The End User


The nal consumers of electricity are not interested in energy per se, but rather the services provided by that energy. The typical consumer does not engage with the

2.3 The End User

13

generation technology unless it either adversely aects them, or they are concerned about the wider impacts of the technology (e.g. climate change).

2.3.1 Locale
As mentioned in 2.2.1, wind power is frequently generated close to a minority population to serve larger centres. Apart from the technical impact of remoteness, this requires careful consideration of populations local to the machines, when the generated power is many hundreds or thousands of miles away [27, 28]. The involvement of large companies in nancing and installing wind parks has enabled the rapid deployment seen in recent times. However, there is a danger of alienating consumers if they are not properly consulted, requiring standards of practice in installation and machine design to be upheld [29].

2.3.2 Further Impacts


Apart from the subjective impact of wind turbines and wind parks mentioned in 2.3.1, a number of quantiable concerns are often expressed. With proper machine design and siting, these can be mitigated. Avian Largely addressed by appropriate siting, a better understanding of the visual [30] and audible [31] signals that birds use for navigation, and a shift away from multi-element truss towers attractive for nesting. Aviation Proper siting away from major ight-paths and markers on the machines minimise the possibility of collision. MoD objections to a number of projects on the grounds that national security is threatened by blocking radar sites [32] may largely be alleviated by radar placement and even stealth blades [33]. Land Use While it is often assumed to encompass the entire area of the wind park [34], in reality it only consists of the tower bases, access roads and any sub-stations that may be present. Animals are happy to graze close to turbines and so farming can continue unabated. Noise Noise was a large problem with early turbines, originating from the blade aerodynamics (see 4.5.1) and mechanical sources in the nacelle. Nacelle acoustic insulation, variable-speed machines (matching noise to background wind noise), and blade shaping (tip and Trailing Edge (TE)) have largely alleviated these problems. Two-bladed machines tend to produce

14

Chapter 2 System Considerations more noise than three-bladed machines [35], and tower thump (see 7.2) from downwind machines must be taken into account.

Visual Flicker Proper siting must avoid sunlight shining though the spinning blades and casting a ickering shadow on nearby building occupants.

Chapter

Variations on a Theme
All wind turbines operate in the context set out in Chapter 2, with the goal to transform energy from the wind into useful work. It is possible to propose myriad concepts for achieving this overall function. Indeed, many researchers and practitioners have done and continue to do so. Jamieson [36] gives a good overview, including various wind concentrators, as well as charged particle, airborne, sail-based [37] and multi-rotor concepts. To be ecient in selecting ideas for more detailed study and eort, it is possible to utilize approaches from design theory to tease out the fundamentals of wind energy. Section 3.1 provides a brief summary of this body of theory. The principles are then applied to the justication of the coning rotor and similar concepts in 3.2 through 3.7, based on the functional requirements. The topology and operational principles of the coning rotor are expounded in 3.8. As outlined in 3.9, the challenges to obtaining the benets of the coning rotor motivate the analysis and more detailed design work presented in Parts II and III respectively.

3.1 Chasing Function


Pahl et al. [38] presents design as a four step process. Although presented as a sequential process, in practice iteration between steps is carried out. This is either required as a result of errors in previous steps, or as a benecial part of the design process where the design process itself is suciently integrated and exible to accommodate it through design space exploration. The steps are: 1. Planning and clarifying the task 2. Conceptual design 3. Embodiment design 4. Detailed design The task here is clear, to produce a concept that cost-eectively provides electricity from the wind. The second and third stages typically involve the creativity 15

16

Chapter 3 Variations on a Theme

of the designer. These steps are crucial, as they determine at a fundamental level the eectiveness of the design. Unfortunately, they are also the most unquantiable and ill-dened steps. The nal detailed design step is relatively well handled by the application of quantitative engineering theory. To aid in the intermediate steps, a number of generalized design theories are useful in analytically deriving the design, by separating the functional and embodiment aspects of the design.

3.1.1 Design Theory


Engineering design can follow one of two approaches, intuitive or discursive, both of which may be complemented by conventional methods [38]. Conventional methods include: literature searches, bio-mimicry, reverse engineering, analogies, and model testing. Intuitive methods, such as brainstorming and other interactive activities, rely on the unconscious ashes of inspiration. They are thus highly designer dependent, and not useful in rigorous concept justication. Discursive methods follow a set of deliberate procedures to analyse the design, taking many forms from rigid and automated design catalogues [39], to knowledgecapture tools [40], through to generic frameworks with abstracted solution spaces [41]. The latter set of methods is useful here, to inform a generic discussion of wind energy converters. In particular, Axiomatic design and TRIZ are the most applicable amongst other possible options: bond-graphs [42], topological-spatialphysical decomposition software [43], and CAD-based software [44]. The key to both is separate consideration of the design in functional and physical spaces. 3.1.1.1 Axiomatic Design Axiomatic design theory is an attempt to set down rigorous rules (axioms) governing design [41, 45]. It does not set out specic steps to follow, but is rather a framework to work in, consisting of two domains. The rst is the functional domain, an abstract space containing the decomposed functionality of the design, from top-level (convert wind to electricity) through to minute detail (e.g. blade root connection). The second is the physical space, containing the actual parts and assemblies required to perform the specic functions. Design activity consists of mapping Functional Requirements (FRs) in the abstract space to Design Parameters (DPs) and Process Variables (PVs) in the physical space. Starting with the top-level functionality, the FRs are sub-divided into 210s of subordinate levels. In parallel, a set of DPs and PVs are co-evolved to aect the

3.1 Chasing Function

17

FRs [46]. Design constraints are not considered FRs, but impose limits in FRs, DPs and Design Variables (DVs). Two fundamental axioms are used (together with numerous corollaries) to inform the choice of the best design. These may be stated in word and numerical form as: Independence Axiom The FRs must be satised independently; i.e. variation in a DP or PV must only aect one FR at a time. Note that physical integration of parts is not precluded, as long as the functionality may be properly controlled. Mathematically this is dened as: F R = A DP DP = B P V C = AB F Ri Aij = DPj DPi Bij = P Vj (3.1.1a) (3.1.1b) (3.1.1c) (3.1.1d) (3.1.1e)

If the design is properly uncoupled, C will be diagonal or triangular. Information Axiom Minimise the information content I; i.e. keep the design as simple as possible to achieve a high probability of success p: I = log2 range 1 = log2 tolerance p (3.1.2)

German design theory of the Workshop-Design-Konstruktion (WDK) school proposes a level of abstraction between the physical and functional spaces: the organ domain [47]. An organ is a collection of Wirk elements, each a point, line surface or volume where a Wirkung (fullment of an FR) is performed. For example, for a tabletop in the physical space, the top surface is a Wirk element performing the function of holding up an object. Japanese General Design Theory (GDT) also centres on a decompositional approach [47]. Taguchi et al. [48] espouses robust design, achieved by incorporating the stochastic nature of the design to ensure that functionality is achieved in all circumstances. This is akin to the information axiom. 3.1.1.2 TRIZ Altshuller in Russia instigated an exhaustive patent search to extract principles common to all innovation, independent of the specic application [49]. The Theory

18

Chapter 3 Variations on a Theme

of Inventive Problem Solving (known by its Russian acronym TRIZ) and the computation tool ARIZ deriving from those eorts, is quite dogmatic and requires large investments of time and user skill. However, the generic aspects of the method are useful here. Eight lines of technical evolution were identied: life cycle, dynamization, multiplication cycle, transition from macro to micro level, synchronization, scaling up/down, uneven development of parts, and automation. Forty inventive principles were found (e.g. segmentation, taking out, asymmetry), useful in solving a contradiction. A contradiction in TRIZ is one of three types: administrative (e.g. lower cost and higher performance), technical (e.g. improving one parameter at the expense of another) or physical (e.g. requirement of multiple properties from same material). Parameters here are mass, length, temperature, etc. The concept of ideality gures prominently in the method, dened as: Ideality = Benets Expenses + Harms (3.1.3)

Technical systems evolve towards higher ideality by utilizing external and internal resources, the former frequently overlooked in a poor design (e.g. a refrigerator in a cold environment obviating the need for a refrigeration cycle). At innite ideality, the mechanism disappears leaving only function. 3.1.1.3 Common Elements An overarching theme common to many design theories is that of thinking in multiple domains, an approach followed in later sections for the wind turbine. Abstraction of function is found to be a powerful tool for breaking out of old ideas, by emphasizing the generality of the essential principles involved [38]. By postponing visualization of the means-to-an-end, the end itself can be concentrated upon, as it is the critical denominator of success. A complimentary theme is a multi-level approach in all domains, as the mind has limited capacity to eectively focus on multiple issues simultaneously. The base functionality must be broken down into sub-functions [38], analogous to parts and assemblies in the physical domain. Implicit in a number of the methods is the concept of lean design, which can be applied in three contexts: process, material and integration. Lean process refers to the streamlining of the design process itself, of focusing attention on the core issues. Material leanness is achieved by optimized form, using the minimal amount of the appropriate material required for the intended purpose. Integration deals with part

3.2 COE and CF

19

count issues and the overall simplicity of the design. The idea of function sharing has also been explored by Ulrich, Seering, Eppinger in their Functional Analysis System Technique (FAST)/Value Analysis [47]. The second axiom of axiomatic design makes lean design explicit, as does the concept of ideality in TRIZ. This is echoed by a number of other corollaries, elements of axiomatic design and TRIZ respectively, compared in Table 3.1. Table 3.1 Axiomatic/TRIZ design method similarities (Adapted from [50]) Axiomatic Design Corollary 2 Minimise the number of FRs TRIZ Ideal Final Result System imposes fee for realizing function, therefore minimize substance, energy and complexity Evolution Pattern 5 Increased complexity followed by simplication Evolution Line Mo-Bi-Poly Guidelines for reducing complication of a system

Corollary 3 Integration of physical parts Corollary 7 Uncoupled design with less Information

3.2 COE and CF


The primary function of a wind turbine is to eciently (economically and technically) convert wind power to electricity.1 Two metrics are useful in this discussion, Cost of Energy (COE) and Capacity Factor (CF): COE = CostAmortized capital + BOP + OM Energy CapturedActual Energy capturedActual CF = Energy capturedTheoretical operation at rated power (3.2.1a) (3.2.1b)

A related concept is availability, dened as the proportion of time the machine is available to produce power (i.e. excluding down-time due to maintenance, etc.). CF and availability are related but not synonymous. It is important to realize the wind turbine availability is on-par with conventional plants. The CF of wind is much lower, typically around 30%, and this is usually misconstrued that wind turbines only work 30% of the time. This is false; the CF is merely a reection of the economic decision of rated power, which is a trade-o between low-wind energy
Electricity production, rather than direct mechanical energy (e.g. traditional waterpumping) is the current focus.
1

20

Chapter 3 Variations on a Theme

capture and high-wind loading. Hypothetically, and extremely large rotor and very small generator would yield a CF of 100%, assuming 100% availability. The COE numerator is dominated by machine and balance of plant (BOP) capital cost, to nance a structure capable of sustaining the applied loading required to convert and transmit power from the wind through to the electrical connection point. The denominator represents the amount of energy captured over the economic lifetime of the machine. At each wind speed V , power coecient CP and area A = /4D2 at that point, the power captured is: P = 1/8V 3 CP D2 (3.2.2)

Note that CP is dened relative to the rotor area. Another common misconception is that for a CP of say, 0.5, a turbine is only capturing 50% of the available wind. In fact, this is a technical measure, fundamentally limited to the Betz limit of
16/ 27

= 0.593. It is akin to stating that a heat engine is only 15% ecient, relative

to a Carnot eciency of say 30%. Both numerator and denominator of COE are aected by the control strategy, inuencing both the loads and energy picked up from the wind. This is in turn constrained by the level of compliance in the drive train (i.e. xed or variable speed) and aerodynamic control (e.g. pitch actuation, active yaw, etc.). Some added cost (e.g. for a coning mechanism) is justied on the basis that the rotor cost is only approximately 10-20% of total cost, so that increase in energy capture can give an overall reduction in COE. Malcolm and Hansen [51, p. 29] found a 10% rotor cost change led to a 1% to 1.5% COE change. The typical FR of a wind turbine is a low delivered COE. The CF of a machine is typically seen as an artefact of that design process and the wind regime on-site. With remote load-centres, for example oshore or in remote areas, it may be benecial to consider a high CF as a FR in its own right. This would better utilize the expensive transmission cabling and infrastructure.

3.3 To Lift or Drag


The second-level choice for the functionality in a wind energy technology is to base it on either lift or drag forces. As will be seen in 3.3.2, the choice is critical.

3.3 To Lift or Drag

21

3.3.1 Governing Equations


In the most general case, a device of area A (frontal area for pure drag device, planform area for lift) travels at velocity Vd through wind speed V , both relative to xed ground. The two vectors subtend an angle , and their magnitude ratio is = Vd /V . The lift CL and drag CD coecients are the fully 3D values for the device. The energy extracting force F in the direction of motion is non-dimensionalized as: CF = F 1/2V 2 A = [cos (CL sin + CD cos ) + sin (CL cos CD sin )]

1 2 cos + 2 sin tan = 1 cos The power coecient is then simply CP = CF .

(3.3.1)

3.3.2 Lift-Drag Comparison


A simple caparison is shown in Fig. 3.1, taking CD = 2.0 and = 0 for the drag device, and = 90 , CL = 0.8 and CD = 0.1 for the lift device. These values are assumed constant throughout (i.e. trimmed to maintain the same Angle of Attack (AOA)). The drag device can never move faster than the wind, limiting its power capture. In fact, its prime functional advantage is a better CF for < 0.2. It is also possible to construct much simpler physical devices utilizing drag, so for high-force/low-speed applications these may have an advantage. Otherwise, the drag device is severely limited, especially if energy capture is the objective. Figure 3.2 shows the eect of the next decision, translation angle . Angles around = 90 have the highest peak force and power coecients. This is fortuitous, considering that the HAWT discussed next in 3.4 operates at = 90 . The alternative is either a VAWT, or a device the translates on rails at some angle to the wind. The latter involves considerable physical infrastructure.

3.3.3 Operational Conditions


Key non-dimensional parameters for aerodynamic performance are the Reynolds number Re = V l/, and some measure of surface roughness. Both characterise the conditions in which the boundary layer develops. With higher Re more energy is fed into the boundary layer, to overcome the surface roughness attempting to retard the ow. Leaving aside the laminar separation bubbles seen at very low Re (around

22
8 7 6 5 4 3 2 1 0 0.0 1.0 2.0 3.0 4.0

Chapter 3 Variations on a Theme


2.0

1.5

1.0 Lift C_P Drag C_P Lift C_F Drag C_F

CP

0.5

5.0

6.0

7.0

8.0

0.0

Figure 3.1 CP and CF for translating airfoils utilizing lift and drag

9 8 7 6 C_F 100 C_F 80

C_F 90

1.8 C_P 90 C_P 100 C_P 80 1.6 1.4 1.2

CP

4 3 2 1 0 0.0 1.0 2.0 C_F 20 C_P 20 3.0 4.0

0.8 0.6 0.4 0.2

5.0

6.0

7.0

8.0

0.0

Figure 3.2 Variation of

CP and

CF with translation direction (deg) for lift device

CF

1.0

CF

3.4 VAWT or HAWT

23

1x104 ), higher Re conditions and smoother surfaces tend to permit higher cl,max by avoiding separation. Critical to lift devices, the cl curve in stall may vary dramatically with Re and roughness, for the same airfoil. To set wind turbines in aerodynamic context with other applications, the wing sections on a large jet transport aircraft operate around 110x107 . A small wind turbine blade would see Re around 110x105 , and a large turbine around 110x106 . The range that wind turbines operate in, is the same range over which large dierences in cl behaviour appear. It is therefore essential that airfoil data is obtained for the operational Re and surface roughness.

3.4 VAWT or HAWT


The fundamental momentum balance derived in 4.4.3 predicts equal performance for a VAWT and HAWT (see Figs. 1.1 and 1.2). The limiting case for a VAWT is a gyromill, having vertical straight blades with continuous pitching to only shed vorticity (change lift) as the blades transit directly upstream/downstream. VAWTs have certain physical advantages including: generator location at ground-level; ability to operate in any wind direction; and no cyclic gravity loading. Mitigating against these are: cyclical aerodynamic loading (this led to fatigue failure of early aluminium blades); operation in the wake of the tower and other blades; usually close to ground loosing wind shear benet. For structural reasons, the blades are usually formed in a troposkein shape (see Fig. 1.2(b)) so that aerodynamically, the blades are constantly changing lift, even in ideal conditions. This fundamentally limits their CP , even with complicated pitching systems. For all of these reasons, VAWTs have been commercially abandoned, for the most part. The exceptions are a resurgent interest at small and very large scale. The former may have benets over a HAWT in built-up areas, with rapidly varying yaw angle.1 The latter is predicated on avoiding the cyclic gravity loads that are starting to drive the design of very large HAWT blades.2

3.5 Scale
The ideal scale of machine is very hard to derive analytically. At the most basic level,
COE would be expected to rise with scale, according to the square-cube law which
1 2

E.g. Quietrevolution from XCO2 E.g. Aerogenerator from Wind Power Ltd.

24

Chapter 3 Variations on a Theme

states that energy capture increases with diameter squared D2 , while the volume of material (cost) increases as D3 . This is of course overly simplistic, for a number of reasons, including falling installation and maintenance costs with fewer overall machines, and improving wind resource with tower height. Some authors have used component-wise scaling laws to arrive at curves predicting optimum machine size [16, 18], while others have pursued more detailed design study on components including blades [52] and balance-of-plant [53]. Jamieson [36] found that a multi-rotor concept would improve the area/volume relationship. Coulomb and Neuho [54] have studied the cost progression of machines, before vendors ceased disclosing list prices, from the perspective of learning curves. A technology typically exhibits cost reduction from the experience gained over time. In this case of wind turbines, when analysing cost data with size it was found important to include wind shear exposing machines to higher wind-speeds as they grow. With these considerations, learning has dropped costs by 12.7% with each doubling in installed capacity. Based solely on machine cost, 400-500 kW machines were found to be optimal, although this excluded Balance of Plant (BOP) factors. In general, optimal size predictions depend on myriad assumptions that are dicult to prove in the absence of real experience. Moving oshore changes the equation, shifting the cost centre away from the turbine to BOP. There is some nite upper-limit on machine size, as the machine cost certainly exceeds the D2 exponential progression. Grin [55] found in a scaling study a Dx cost exponent of 2.9, whereas the commercial average is 2.4. The largest inuence was design condition (see Table 2.1) making tailored machine/rotor design increasingly important. The TRIZ technical evolution trend of scaling up is clearly evident in the wind industry. Onshore the limit is practically around 2 MW, owing to transportation restrictions. Oshore it is less clear, with 5 MW prototypes currently being installed. Changing materials complicate trend analysis of blade weight (a proxy cost metric), shown in Fig. 3.3 compiled from vendor data sheets. Vestas is increasing the use of carbon bre to obtain stiness, while LM evolves their standard polyesterglass blades. High-performance materials (e.g. carbon bre) may deliver technical performance, however cost performance can be adversely aected if aerospace type composites are employed (Vestas is transitioning to pultruded carbon bre-wood composite blades to avoid this). The averaged curve t indicates a cost exponent with D of 2.1, for this data set. Figure 3.4 shows historical data compiled from WindStats magazine for a range

3.5 Scale
25000 LM Vestas PNE (Multibrid) Enercon Siemens Ecotecnia

25

Per Blade Mass (kg)

20000 15000 10000 5000 0

y = 0.7911x2.1055

20

40

60

Rotor Diameter (m)

80

100

120

140

Figure 3.3 Blade mass trend with rotor diameter of machines comprising full data sets over the year. No oshore data is included, as the wind resource would tend to skew general trends. There is large data scatter and some outlying data sets, with less variance in summer months possibly owing to less persistent storm activity. The seasonal variation is evident in the seasonal trend lines, with winter winds higher than those in summer. There is a general increase in average energy capture with machine size, but it is fairly gradual and only in evident in winter and spring.
Specific Energy Capture (kWh/m )
400 350 300 250 200 150 100 50 0 0 500 1000 1500 2000 Autumn Winter Spring Summer
2

Rating (kW)

Figure 3.4 Specic energy capture as a function of machine rating For the purposes of this thesis, a 1.5 MWe machine is used as a target scale. This size is representative of on-shore machines currently being installed. Data for

26

Chapter 3 Variations on a Theme

the REF-1500, a conventional machine (see C.3) was also available for this size machine.

3.6 Control Strategy


Wind turbine control synthesis is divided into two sequential stages: scheduling and controller design [56]. The latter provides the control loops and gains using control theory [57], but is subordinate to the rst task of specifying the control strategy and targets. Functionally, this is a complex and critical step (as will be shown in Chapter 6). The second step is therefore left to previous [56, 58, 59] and future work.

3.6.1 Goals and Methods


All wind turbines have two competing goals, to capture energy while avoiding loads. Figure 3.5 shows the three control regions: Region I below Vci , where the wind turbine is parked; Region II to optimally track the maximum power extraction point; and Region III above Vr to track the peak power of the generator, up to a maximum Vco where the rotor is parked. The choices of Vci and Vr are economic, as there is low energy content at low wind speeds (see 3.2). Likewise, Vco avoids extreme loads on the machine, at speeds that have very high energy content but very low frequency of occurrence.

Figure 3.5 Power curve regions The mechanism of control is important to both loads and energy capture. In Region II, variable speed operation permits optimal energy capture operation. This is usually done at xed pitch angle, ideally with maximal capture area. Region III

3.6 Control Strategy

27

must limit rotor power to that of the generator maximum, though one of a number of mechanisms. Yaw1 and coning both aect the gross capture area, via Eq. (3.2.2). More conventional control methods for Region III, in order of decreasing usage are: Pitch to Fine (PTF), Fixed Speed Stall (FSS), Pitch to Stall (PTS), and Variable Speed Stall (VSS). Each alters CP , either by changing the relative velocity vector (FSS, VSS) or AOA via pitch angle (PTF, PTS), as shown in Fig. 3.6. Notice that each maintains the same product of torque-producing force vector and rotational velocity r, to produce the same power. The resulting thrust force also varies between strategies.

Figure 3.6 Wind velocity and force vectors for various control strategies

From this list of choices, modern machines use almost exclusively PTF. A coning rotor combines coning control with VSS or PTS. The Proven machine (see Table D.1) adopts an even more complicated use of PTS coupled to rotation speed, and PTF coupled to cone angle. Adaptive exible coupling between exible blades (in bending or speed-linked extension) and torsion have been proposed by a number of authors [60, 61]. Veers et al. [62] provides a good overview of the static possibilities, and highlights the dynamic stability bounds and strict manufacturing tolerances required to practically execute such a strategy. Figure 3.7(a) shows the conventional control strategies considered from the nondimensional rotor perspective, with associated CP loss mechanisms either side of the peak CP,opt . Pitching strategies alter the CP curve directly, while VSS and
FSS both move along a nominally constant curve.2 Note that FSS, VSS and PTS

all operate in the left-hand stalling portion of the CP curve, while PTF avoids
1 2

E.g. Gamma 60 machine [16, p. 357] Re eects modify the curve somewhat.

28

Chapter 3 Variations on a Theme

stalling. The impacts of coning on this control map are discussed in 4.3, 4.7 and 6.3.

(a) CP

(b)

Figure 3.7 Power control in CP and planes

Figure 3.7(b) presents the control schedule as it is actually implemented. The wind speed cannot be accurately measured, so it is necessary to indirectly control based on an aerodynamic1 torque estimate and rotor speed . The generator will impose lower min and upper max bounds. After Vci , the rotor tracks from A to B in low winds. With sucient wind, the optimal CP,opt of the rotor is tracked by maintaining opt . This may be altered somewhat by the speed-dependent ineciencies in the drivetrain/generator. At some point C the upper speed limit is reached, sometimes prematurely to smoothly transition between Region II and Region III controllers. Pitching strategies PTF and PTS maintain a constant torque at Dpitch , while VSS must increase torque to reduce and stall the rotor towards point Dstall . Additional control inputs reduce somewhat the burden of analysis delity and aord performance enhancements. In particular, the stall behaviour of the rotor is subject to considerable uncertainty between prediction and measurement [64]. Provision of pitch action enables on-line tuning to account for modelling errors, and adjustment for air density variation and blade soiling. Active control of power in Region III (i.e. not FSS) removes the prime arguments for low-lift airfoils [65, 66], enabling the use of high-lift airfoils for more cost-eective blade designs.
Controlling on generator torque is sub-optimal, as the rotor inertia is involved rather than energy-producing aerodynamic torque [63].
1

3.6 Control Strategy

29

3.6.2 Dynamic Considerations


Although this work focuses on the steady-state facets of the control problem, a number of dynamic considerations must be kept in mind. All variable-speed concepts attempt to maintain optimal power capture in Region II. The slope either side of the CP peak in Fig. 3.7(a) eects the ability of the control system to accomplish this objective in unsteady winds. If the peak is sharp (usually a steep stalling front to the left), sharp drop-os in power will occur since the rotor speed response is limited by inertia. In particular, VSS rotors require a sharp peak to limit power in Region III, making optimal low-wind operation a competing design objective. Mercer [58], Homann [59] and Burton et al. [16, p. 476] all discuss the fundamentally poor power regulation of VSS concepts in Region III. VSS is better than FSS, in that steady-state control can maintain rated power and actively transition from the optimal cl /cd point to a stalled point. However, to do this, excess torque must be applied not only in steady-state (torque must rise to maintain power P = ), but also dynamic torque to alter the inertia of the rotor (via dyn = Id/dt). This lag provides greatly reduced power quality above rated. The negative slope of the Dpitch Dstall VSS control objective in Fig. 3.7(b) is unstable and demands a compensating controller. New grid demands such as LVRT (see 2.2.2) also make direct control of rotor input power, via pitch action, increasingly a control requirement.
PTF has been widely adopted because with fast actuators, it directly and quickly

limits input rotor power. It is usually preferred over VSS because of the diculty in analysing stalled rotors. The only large turbine to use PTS is the V-82,1 the perceived advantages being smaller/quicker pitch actions and reduced load variation associated with gust-slicing [16, p. 355]. The latter occurs because PTF operates in the linear-regime where much larger cl changes can occur with AOA than in stall.
PTF may also interact with tower vibrations in the following sequence: pitch action,

thrust decreases, tower moves upwind, relative velocity increase, more pitch action.
PTS operates in an opposite sense, reducing fatigue loading while increasing mean

loads, as discussed in 6.1.2. Practically, experience2 is that PTS does reduce fatigue loading relative to PTF, if the ill-damped edgewise vibrations found in large blades are adequately controlled
1

V82 1.65 MW machine (previously NM-82), Vestas website, http://www.vestas.com/ pdf/produkter/2006/V82_UK.pdf, July, 2006 2 Personal communication, Tomas Vronsky at Vestas Wind Systems A/S

30

Chapter 3 Variations on a Theme

(see 7.1.2.2). However, extreme loads are higher in extreme yaw error situations. In turn, Raben et al. [67] found higher blade ap fatigue for PTF versus FSS.

3.7 Adapting Structures


Structures incorporating Degrees of Freedom (DOF) are able to undergo conformational change. Properly designed, these changes are capable of reducing the applied loading and resulting structural stresses developed in the structure. The following sections explore the application of this concept to wind turbine design.

3.7.1 Dynamic Motion


Since Putnams original apping blade machine in the 1940s [8], motion of the blades has been proposed to alleviate dynamic system loads. Indeed, teetering of two-bladed machines is almost essential for viable performance [15, 68], in the same way as hinges are required for successful helicopter designs [69]. A number of researchers have examined two and three-bladed machines with individual discrete apping hinges [7, 51, 66, 7073] or combined exible hinging and teetering [74, 75]. However, only the Carter machine has achieved any widespread deployment in the past. The functioning of a exible structure is best explained by the fundamental dynamic equation of any structure, be it connected by discrete or exible elements: [M ] x + [Caero ] x + [Cstructure ] x + [K] x = F (t) (3.7.1)

where [M ], [Caero ], [Cstructure ], and [K] are the mass, aerodynamic and structural damping and stiness matrices, x the generalized displacements and F the applied forces. In a sti structure, displacements are relatively small; therefore forces are reacted almost exclusively by the stiness, which must be large. This has the eect that the ultimate stress/strain capabilities of the material are underutilized and the structure may be overly heavy and expensive. One way around this is to use light highmodulus materials, such as carbon bre composites, but this drives up the cost when used in large quantities. In a compliant structure, the velocities and accelerations are non-negligible, and the displacements are also larger. As a result, the stiness requirement is reduced, since the forces may be reduced by modifying the airow. What forces are imparted, will be reacted more by the mass and damping of the

3.7 Adapting Structures

31

structure, rather than stiness. The trend towards dynamization is one line of technical evolution in TRIZ [49]. Aerodynamic blade loading may be tailored with soft structures, most notably in the blades themselves, such as bend-twist coupling [60], apping ex-beams [9], or by discrete apping/teetering hinges [71]. All act to dynamically change the angle of attack at the blade sections by congurational change, so that negative feedback of load is achieved. Discrete hinges near the root of the blades avoid the complexity and stringent manufacturing tolerances of exible blades or hinges. Kelley et al. [76] in examining the Wind Eagle (a derivative of the Carter machine), and Quarton [75] from monitoring of a Carter 200/300, have highlighted the loads seen in practice from imbalance in exibility and mass between blades in a exible approach.

3.7.2 Load Matching


Any variably coned rotor also benets from a static matching of thrust and centrifugal loads. The steady out-of-plane bending moment carried along the blade is reduced, ideally leaving only an axial tension load. Lighter, cheaper blades then feed back to reduced edgewise gravity-dominated loads, further reducing blade weight and cost. Recent commercial eorts [70] and research studies [71, 77] have highlighted the eectiveness of apping blades in this respect. Obviously, any design with signicant apping or coning must have a downwind orientation in order to avoid tower strike. Kelley et al. [76] has noted that the Wind Eagle has reduced loading at high winds relative to conventional rotors, but the reverse in low winds. Evidently the bending and coning at high winds alleviates loading in high winds, but suers from a lack of pre-cone (from centrifugal opening) in low winds.

3.7.3 Soft Approach


The concept of an adaptable machine has been pursued in a number of machines and design approaches. In addition to the rotor itself, other aspects of the machine may adopt a sti or soft approach. For example, the fundamental mode of the tower may be tuned relative to the blade passing frequency B: ftower > B < ftower < B ftower < sti-sti sti-soft soft-soft (3.7.2a) (3.7.2b) (3.7.2c)

32

Chapter 3 Variations on a Theme

where B is the number of blades. Softer designs yield lower mass (cost) towers. The generator may also incorporate exibility, from direct-connected synchronous, to standard induction, variable resistance (slip) induction and doubly-fed induction through to fully-variable synchronous generators.

3.7.4 Coning Rotor Adaptation


The coning rotor concept at its heart employs ap-hinged blades, thereby inherently beneting from the static and dynamic load alleviation just mentioned. The coning rotor is further dierentiated from a apping rotor by two other operational characteristics. Firstly, extending the range of coning angles (gross coning up to 85 ) avoids storm loading when shut down. Secondly, by utilizing relatively long blades, the COEs denominator is increased by enhanced energy capture in partial load conditions. The longer blades are made possible by coning to appreciable angles (2035 ) as rated power is reached, to have similar loading to conventional rotors at that load-critical point. This creates an adaptive rotor with large area in low winds and smaller area in high winds. As the limits of aerodynamic performance (power coecient CP ) are reached by conventional designs, this adaptive rotor acts on the power/energy Eq. (3.2.2) in the most direct way, via the capture area /4D2 . Conventional PTF machine developments towards individual pitch control [78, 79] are aimed at load mitigation. With reduced operational loading, longer blades are possible [80]. The goal is the same as the coning rotor, to maximize energy capture from a larger rotor area. In contrast to coning rotors though, conventional rotors will always remain susceptible to 3D turbulence-induced limit loading while shutdown. The parasitic power loss and bearing life eects associated with aggressive continuous pitch actuation schemes have also yet to be quantied.

3.8 Function to Form


Having examined the functional aspects of an ideal wind turbine, evolution of the concept in the physical domain is explored. These choices lead to the concept explored in more detail in Part III.

3.8.1 Historical Context


Adaptable machines are typically much more dicult to design, given their dynamic nature. Indeed, it is partially the increased design challenge, associated with a lack

3.8 Function to Form

33

of adequate design tools, that has limited the success of a soft approach. Past experience is important, and is garnered here from a review of past machines in Table D.1, to yield the following principles: Blade Count The proclivity for two blades is a testament to the fact that soft designs require by their very nature more advanced analysis and design to be successful operationally. When a soft design approach is adopted, the more detailed analysis also seems to typically lead the designer towards a two bladed conguration. This is presumably for reasons of economy and simplicity when designing systems for load alleviation. Blade Articulation The only machine not employing articulation is a 3bladed one. This implies a clear requirement for the blades to teeter/ap so as to alleviate aero and structural loads, as previously discussed in this section. Coning Coning is usually used in a xed sense, relying on the inherent changes in eective coning angle resulting from blade exibility. The Ris Soft Rotor and Carter machines separated the coning and teetering functions. Both blade roots incorporated ex-beams to provide stiness to a pivoted joint. The MS-4 machine also employed ex-beams which turned out to be dicult components to design . The WTC machine employs independent hydraulically-damped discrete hinges. The Cone-450 concept machine was to use a central hydraulic cylinder to collectively cone the three blades, with dampers incorporated into the control links. It was the only one to use gross coning to adjust capture area for power control. Blade Flexibility The Htter-Allgaier machine opted for a low-solidity rou tor, so that the slender breglass blades were quite exible; as a result the prototype suered from utter. The Ris Soft Cone concept used no shear webs so as to be lightweight and oer the possibility of adaptable airfoil geometry. A common theme is to consider the bending moment supported by the blade. In order to ex (in bending) to any large degree requires large bending moments, which in itself is not desirable from a stress point of view. Downwind Orientation The majority of machines use a downwind conguration, although not universally, due to potential tower-shadow problems. Scale Early machines attempted to jump into the multi-megawatt scale in order to prove themselves on a conventional utility scale. This approach

34

Chapter 3 Variations on a Theme was an almost universal and unmitigated failure. The more successful machines to date have been much smaller. This is also echoed by the evolution of the Danish concept that has matured by growing in size, building on past experience.

3.8.2 Topology
The nominal layout of the rotor under consideration in this study is shown in Fig. 3.8. In the CONE-450 original work [66], the blades were hinged on a space-frame with the hinge axis signicantly away from the rotor axis. The lightweight carbon bre blades under consideration required this conguration to reduce the aerodynamic hinge moment and obtain reasonable free-coning angles ( 30 ). A more conventional compact cast hub is preferable from a complexity perspective, with conventional blade materials and mass tuning.

Figure 3.8 Coning rotor schematic layout (Only symmetric half of nacelle and inboard part of one blade shown)

A central hydraulic actuator was envisaged for the CONE-450, with links out to the blade roots that extended inboard of the hinge points, to impose equal moments about the apping hinge. The basic operational principle was for collective coning action, with the actuator active only during very light winds to hold the rotor more open with minimal bending moment, and during fully coned shut-down. The rest of the time the actuator was simply a means of applying damping (alleviating the loss of aerodynamic damping in stall) with relief values to assure free-coning from just below rated.

3.8 Function to Form

35

In the original coning work, the rigid links from the central actuator to the blades resulted in excessive dynamic overturning moments at the tower head. This was determined to be a result of the proximity of the 3P frequency range and the rst blade apwise mode (this is a fundamental result, see 4.6.3.2). To alleviate this loading, exible damped links were incorporated into the model, solving the problem in the simulations. The physical implementation of the dampers was never discussed in the reports. Experimentation with blade exibility was found eective in reducing blade loads, but individual apping was required to alleviate overturning moments. In the current work, the focus is on three independent actuators for three blades. This is a reection of the need for some individual blade motion found in previous work, and for a cleaner practical implementation. A more conventional compact hub with the actuators acting outboard of the hinge line on moderate weight, mass tuned blades is also adopted. A more integrated design approach is also being pursued, to closely couple a Permanent Magnet Generator (PMG) into a simple rotor, for enhanced reliability and cost eectiveness. This approach is being pursued on conventional machines with standard generator designs [81, 82] and also using extremely large-diameter bearings for the generator [83]. The two-bladed Wind Turbine Company (WTC) [84] prototype machines use independent dampers on individual ap hinges, achieving a type of teeter. Pierce [77] investigated a machine with an actuator between two otherwise freely apping blades. Both employ PTF (see 3.8.3) and therefore maintain small cone angles. A big problem with conventional teetering and these concepts is hitting stops, which re-introduces large loads. The coning rotor operates well away from any stops in a range of coning from 5 to 85 , instead of the WTC -5 to 15 . The failure of the WTC prototype from tower strike, presumably after stop impact, highlighted this risk. A three-bladed rotor is used in the present work, for a number of reasons: Public acceptance and aesthetic studies have indicated this preference [36] Lower optimal tip-speed ratios for on-shore siting issues Aerodynamic performance loss for two-bladed rotors is signicant [51] Two blades are not necessarily cheaper than three when size (solidity) and loads are accounted for1 Cyclic rotating imbalance-type hub loads [71] should be less in a three-bladed rotationally symmetric apping rotor
1

Personal communication with Peter Jamieson, Garrad Hassan and Partners (GH) 2004

36

Chapter 3 Variations on a Theme

The imbalance loads were not mentioned as signicant in the original coning rotor work but were in the two-bladed apping study [71]. The latter study also suggested high tip-speeds to maintain a relatively at rotor (to keep energy capture high) with the potential for attendant dynamic instability resulting from the low-solidity rotors required. Note that the coning rotor uses longer blades and hinge moment bias to eect even greater energy capture than a completely planar rotor.

3.8.3 Pitch Control


Chapter 6 will discuss the requirement for pitch control of some kind on a coning rotor. The CONE-450 used xed pitch and VSS. The attraction of VSS is elimination of a set of actuators and the possibility of integrating the hinges into the blades without requiring a circular root.1 Pitchable tips could be used instead, as is done for tip-brakes on some FSS machines. However, the mechanisms are dicult to integrate structurally and can pose maintenance issues being located at the blade extremities. An additional method of pitch control is possible by passively coupling pitch angle to ap angle by inclining the hinge axis by an angle 3 in the plane normal to the rotor rotation axis. The non-linear relation is: = cos tan 3 (3.8.1)

Helicopters and teetered rotors use +3 to pitch towards feather with ap angle, reducing loads. The coning rotor must positively cone with rising winds, so +3 is inappropriate. The rotor must stall (either VSS or PTS) to move away from the tower. While 3 would achieve this objective, dynamically it would exacerbate stall instabilities. At larger cone angles, gross in-plane (azimuthal) movement of the blade axis would further complicate matters (rotating imbalance, aerodynamic feedback). In any case, most pitch action occurs near = 0 , while it is only required at larger near and above rated power.

3.9 Coning Rotor Challenges and Opportunities


The preceding sections of this chapter have outlined the rationalization for the important functional and physical elements of the coning rotor. Using the decomAnother concept is an inner member with ap hinge allowing rotation of an outer shell, e.g. MS-4, WTC [85]
1

3.9 Coning Rotor Challenges and Opportunities

37

position approach of design theory, minimum COE has been identied as the core function of a wind turbine. Based on fundamental arguments, a lift-based, HAWT of around 1.5 MWe has been identied as the optimal basic approach. The qualitative steady state and dynamic ramications of control strategy have been discussed, as have the fundamental reasons for adopting stall-limited (VSS or PTS) high-wind operation of the coning rotor. The notion of an adaptive machine to tailor and mitigate loading, both in steady state and dynamically, has been identied as a key driver for load reduction. The key physical aspects of the coning rotor are apping hinges at the blade roots, allowing the blades to sweep out a cone and to park in the streamwise direction in high winds. This conguration aords reductions in both parked high-wind loads and operational blade bending moments. The coning rotor exploits these load reductions by employing relatively longer blades, with nominally constant cost, to yield lower COE relative to conventional machines. Although advanced conventional
PTF machines (for example with independent pitch actuation) may also increase

blade length by reducing loads, the non-apping blade roots of these concepts fundamentally remain more susceptible to high-wind parked loads. They also must resist larger steady bending moments during operation. The coning rotor therefore appears a viable alternate approach worthy of more detailed consideration in Part III, given the potentially large benet in reduced COE. As a conceptual approach to extracting energy from the wind, the coning rotor shares many elements with conventional machines. The analysis tools required will therefore be similar to those currently used and validated, but with important dierences, as covered in Part II. The primary area of uncertainty is in the aerodynamics, which is already complex in stalled unconed rotors. The theory on which almost all wind turbine design tools are founded, BEM, requires modication to handle the geometry of the coning rotor, as detailed in 4.4. BEM theory is extended here, rather than employing a higher order model, to facilitate dynamic simulations and optimization on a design relevant time-scale. The inclusion of non-linearities in the structural model (frequently linearised for conventional machines) are also found to be important when considering a coning rotor. The required downwind orientation of the coning rotor may lead to LFN, a negative eect not encountered with conventional machines but which must be mitigated for downwind machines. Finally, the solution of the optimization and control problem is complicated by the presence of the ap hinges, requiring modied

38

Chapter 3 Variations on a Theme

approaches and due consideration of the attached generator. Motivated by the qualitative potential benets of the coning rotor, the remainder of this thesis is focused on developing, validating, and applying design tools to examine the quantitative performance of such a rotor. The ultimate question to be answered is the permissible increase in blade length, which will in turn determine the COE advantage of the concept. In turn, this question may only be answered with any certainty after building condence in the design tools employed, with due appreciation for any shortcomings in the implemented analysis methods.

Part

II

Analysis

Chapter

Analytic Development
This chapter develops the theory validated in Chapter 5 and used later in Part III for design studies. The presentation is motivated by the unique conguration of the coning rotor described in Chapter 3, which violates some of the assumptions in standard models, most principally aerodynamic (4.4) and structures (4.6) related. Most of the models are implemented in a common software program, ExcelBEM, to facilitate later work.1 Throughout this work, emphasis is placed on extending engineering-level models, rather than more complex full-eld simulations, to facilitate rapid optimization and for incorporation into dynamic simulations while retaining reasonable runtimes. A number of analytical models are required, principally a novel BEM method capable of properly analysing a coned rotor (4.4), equally applicable to other highly exible concepts and extensible to yawed rotors and dynamic inow (4.4.14). The implementation of a low-frequency acoustic model is presented next in 4.5, required for evaluating the tower thump noise associated with downwind rotors. Structural modelling methods are presented next in 4.6, including sectional property computations (4.6.1), loads integration (4.6.2), dynamic EOM for the ap-hinged rotor (4.6.3), and derivation of a FEM approach to the centrifugally stiened beam problem in 4.6.4. Using these components, steady-state and dynamic control scheduling methods are presented in 4.7. Finally, a generator (4.8) and cost model (4.9) are presented. Before any of these models are presented, an industry-standard reference code is introduced (4.1), in addition to an overall coordinate system (4.2) and performance metrics for the coning rotor (4.3).

The name ExcelBEM derives from the fact that Microsoft Excel is used as a graphical user interface, with sheets to dene the blades, structure, etc. Matlab m-les and compiled C functions are used for computations and interaction with Excel.

41

42

Chapter 4 Analytic Development

4.1 Industry Standard Comparison (BLADEDTM )


Modern, industry standard, fully aeroelastic time-domain wind turbine simulation programs are quite complex. They have been developed over many years and certied by regulatory bodies. A large quantity of data must be properly input describing the machine, operation and conditions, including the blades, drive train, tower, controllers, and wind/wave inow. Wind turbines are inherently exible structures, making accurate coupling between aerodynamic loading and structural deformation critical for accurate performance prediction. These codes therefore adopt some modal or discretized description of the exible components of the system. Taken together, developing a full commercial-grade code is a large undertaking, well beyond the scope of an individual thesis. The BLADEDTM software from GH is used throughout this thesis as a comparison industry-level code. BLADEDTM is a program frequently used in industry for turbine design and certication, incorporating many useful features for detailed analysis and design. Preprocessor modules can create turbulent wind histories and compute modal properties. Steady-state aerodynamic and power curve analyses are possible, as are dynamic, fully aero-elastic computations including oshore wave loading and controllers. Post-processors are available for data reduction and visualization, including extreme and fatigue load computations, and state-space model linearisation.

4.2 Coordinate Systems


Prior to embarking on topical development, a common Coordinate System (CS) is developed to describe the machine. A set of DOF are shown in Fig. 4.1 that dene the position and orientation of the following components: Tower Top-head translational (xn , yn ) and rotational (nod , roll , yaw ) motion, in that order Nacelle Osets from the yaw-axis (o, h, l) Hub Shaft tilt t about y and rotation about x (positive rotation clockwise when looking downwind), and oset Rhinge of the coning hinge axis Blade Cone angle (positive coning downwind), distance from coning hinge to blade root shub and section distance along blade axis s. All section twist and pitch angles are about the zaero axis, positive towards feather. These DOF are sucient for the current purposes which focus on rigid-body motion,

4.2 Coordinate Systems

43

but which contain enough generality that extension to exible components should be straight forward.

(a) Earth-tower-nacelle

Figure 4.1 Coordinate systems

It is important to point out a further simplication that is made with respect to any exibility of the blades, and the blade pre-bend dened in Fig. 4.29, both of which move the section chordline o the pitch axis (zaero axis in Fig. 4.1(d)). For the purposes of aerodynamics, any o-axis position is ignored, as it will be small compared with other dimensions. Any structural velocity, however, resulting from exibility, is included in the aerodynamics. The structural calculations in 4.6.1.4 take into account the pre-bend, as do the load calculations in 4.6.2, as any mass oset from the pitch axis will have an important eect. The various CSs associated with the components are also detailed in Fig. 4.1. Throughout the following development, the transformation matrices dened in Appendix A are used to relate CSs. The most general transformation from blade

44

Chapter 4 Analytic Development

(b) Nacelle-shaft (downwind)

(c) Nacelle-shaft (upwind)

(d) Shaft-blade

Figure 4.1 Coordinate systems (cont.)

aerodynamic axes xaero yaero zaero to earth (xed) xe ye ze is: x y = T (xn , yn , ht ) R y (nod ) R x (roll ) R z (yaw ) T (0, l, h) . . . z 1 e 0 0 R y (t ) R x () T (0, 0, Rhinge ) R y () (4.2.1) shub + s 1 For anti-clockwise rotor rotation, the sign of is negative in Eq. (4.2.1). Also, for downwind rotors t is negative, while for upwind rotors o is negative. Unless otherwise stated, the tower is assumed rigid throughout this thesis (i.e. xn = yn = nod = roll = 0).

4.3 Coned Rotor Performance Metrics

45

4.3 Coned Rotor Performance Metrics


A variable-area rotor introduces an ambiguity into the conventional denitions for tip speed ratio , thrust CT and power CP coecients. Each may be based on either the actual R = Ractual (real, coned) or reference (unconed) tip radius R = Rref . For a coning rotor of blade length S, they are dened as: Ractual = Rhinge + S sin() Rref = Rhinge + S R = U P CP = 1/2U 3 R2 T CT = 1/2U 2 R2 (4.3.1a) (4.3.1b) (4.3.1c) (4.3.1d) (4.3.1e)

4.4 Aerodynamic Modelling


Clearly, aerodynamics lies at the heart of energy extraction from the wind. To properly design a blade, accurate predictions of the aerodynamic forces and derived quantities (e.g. rotor power, torque) are required, in both steady and dynamic conditions. Numerous tools have been developed based on three basic formulations: Blade Element Momentum (BEM), potential ow (vortex laments/lifting-line), and fulleld solutions of the Navier-Stokes (NS) equations. Advanced aerodynamic analysis techniques for wind turbines are in principle available to handle the aerodynamics of the coning rotor concept, including lifting-line [86, 87], Navier-Stokes actuator line and disk [88, 89] and full-eld simulations [90]. It should be noted at this point that an initial attempt was made by the author to modify a computationally ecient1 vortex code provided by Dr. Coton [87], to handle the coned rotor. This approach was abandoned for two reasons. The rst was that even in its unmodied original form, the analysis of a single steady case took on the order of minutes to compute. This rendered it unsuitable for later design work. Secondly, it became apparent that the underlying equations would require extensive reformulation to account for a coned rotor geometry.
This formulation splits the wake into a near and far part, with prescribed wake geometry. This formulation involved the least possible computations, principally evaluations of some form of the Biot-Savart law (see 4.4.5).
1

46

Chapter 4 Analytic Development

In any case, practical design codes retain, at their core, implementations of BEM theory. The BEM approach is much less computationally demanding than any other approach, a key requirement for analyses taking into account inherent and important coupled elastic behaviour, on a time scale relevant for design iterations. It has of course been acknowledged that more advanced aerodynamic models are required for better understanding of many wind turbine phenomena [91], but currently available resources prevent these approaches from being benecial in design practice and optimization. Bearing in mind these constraints and discussions of the inherent contradictions in BEM [92], it has nevertheless proven to be quite adept at adequately predicting performance for standard rotors. Highly coned rotors violate some of the underlying assumptions of standard BEM theory, although on closer examination and with the proper corrections, not as badly as might be expected. In light of BEMs relevance to wind turbine design, its underlying assumptions and history are elucidated in 4.4.1 and used to correct the model in later sections, to make it applicable to the analysis of coning rotor concepts. The revised BEM method is rst derived for the steady case to illustrate the fundamental concepts. The reader is led from a description of the ow eld in 4.4.2, through momentum equations in 4.4.3, then to blade element forces in 4.4.4, before introduction to the derived correction factors in 4.4.5 through 4.4.7. The postulate is that the critical limitation of BEM is the usual assumed relation between disc and far wake induction. Proper consideration of the relative placement of the wake, and inclusion of radially induced velocity, using potential ow theory, yields the corrections to the BEM method. The validation presented in 5.1 demonstrates that the main limitation of BEM theory in application to coned and yawed rotors is in the assumed wake position, not the planar disc nor the independence of streamtube assumptions discussed later. A model for centrifugal pumping is selected in 4.4.8, before synthesizing a spanwise ow stall delay model in 4.4.9. Finally, 4.4.10 presents the nal corrected BEM equations, with additional wake geometries in 4.4.11 and 4.4.12. Building on the same fundamental insights, the revised BEM method is extended in 4.4.14 to a time-domain formulation handling dynamic and yawed inow.

4.4 Aerodynamic Modelling

47

4.4.1 Back to Basics


The lineage of BEM theory dates back to Rankine [93] and Froude [94], who before 1900 developed uniform actuator disc models. These were based on linear momentum balances, and were extended by Betz [95] in the 1920s to include wake rotation. The idea of using blade elements to predict forces on the blades was originally developed by Froude [96] and Drzewiecki [97]. Glauert [98, 99] provides a useful discussion of the marrying of momentum theory to blade element theory by 1940 and the subsequent history of accounting for nite numbers of blades, which has remained virtually unchanged in modern BEM. It is this 60 year old presentation which has lent the insight required to extend BEM to account for coned rotors. The approach taken is somewhat analogous to recent extensions for skewed rotors [100], arising from work done in the 1940s and 1950s on helicopter rotors [101103]. In order to properly derive a modied BEM theory appropriate for coned rotors, it is necessary to go back and consider separately the momentum and blade element components. This can only be done after properly dening the ow eld through the rotor.

4.4.2 Flow Field Kinematics


In keeping with the standard set-up for BEM, the axisymmetric ow through a generalized coned rotor is shown in Fig. 4.2, from far up-stream conditions at station 0, passing through the rotor at station 1, to far down-stream at station 2.1 Inviscid ow is assumed throughout, except for the thin shear layer at the boundary of the wake. The ow is implicitly assumed to mix back to the free stream velocity in the very far wake downstream of station 2. The details of the far wake eld will be discussed further in 4.4.11 as they relate to modelling at high induction factors. The incident ow at station 0 is taken as constant at all radii. The eects of inow turbulence, yaw and wind shear are all ignored in the present steady, axisymmetric analysis until 4.4.14. The downstream velocity varies radially between streamtubes, as a function of the calculated loading. Both far-eld velocities, 2D vectors V0 and V2 in (, z ), are taken to have zero radial component, the formers magnitude r being an input variable and the latter an output of the simulation. The situation at the rotor for a single stream-tube is expanded in Fig. 4.3. The radial component of the ow (normally neglected) is shown, together with the aggregate rotor/blade
Note that this aerodynamic coordinate system diers from 4.2. The equations were originally derived for comparison to Mikkelsen [89] who used the CS used in this section.
1

48

Chapter 4 Analytic Development

Figure 4.2 Kinematics setup

generated forces acting on the ow through the individual stream-tube. The forces are either reactions of the lift and drag forces on the blades, or hypothetical idealized pressure forces. For a real rotor imparting azimuthal forces as well, there is also an azimuthal velocity component at the disc w and w,2 in the far eld, both pointing into the page.

Figure 4.3 Kinematics at rotor for single stream-tube Note that the radial velocity component modies Vn only for coned rotors, and aects the predicted velocity relative to the 2D airfoil. Critically, this is a description of the generalized velocity components of the true net ow eld through the rotor. Mikkelsen [89] presented a BEM method (in addition to the CFD results discussed in 5.1) which accounted for the radial velocity component, but as will be shown later, erroneously assumed wn to be parallel to Fn . The proposed rationalization was that BEM governing equations only consider axial momentum, and there could therefore be no induced velocity parallel to the rotor blades. In fact, there is radial ow and it is simply a modelling deciency that the radial ow is conventionally

4.4 Aerodynamic Modelling unaccounted for.

49

In the present method, the induced velocity components w are the orthogonal increments from the free stream velocity at the rotor, proportioned via the ow model to give the true net ow vector at the disc. The net ow eld can be rationalized by momentum considerations to be the result of the eects of the local loading of the disc (retardation of the ow in the stream-tube, which must result in ow expansion and hence radial ow), or as the induced eect of the vorticity bound on the blades and trailed in the wake. The latter approach will later prove quantitatively expedient in 4.4.5, yielding induced velocities u.1 The standard relations between the axial a and tangential a induction factors and induced velocities wz and w are retained in the present analysis for convenient comparison to the standard BEM equations: wz = aV0 w = a r Note that these are the induced velocities at the disc, station 1 in Fig. 4.2. (4.4.1a) (4.4.1b)

4.4.3 Momentum Balances


Momentum theory was the basis of Froudes early analysis of a uniformly loaded disc. Later, the disc was divided into annular tubes, that aord radially varying induction at the disc, when pressure and friction eects between stream-tubes are neglected.2 Pressure drop in the wake owing to rotation is also ignored in assuming a pressure of p0 at station 2 [16, p. 62]. Assuming wake rotation to be small compared to also greatly simplies the following momentum equations, as shown by Glauert [98]. All of these common simplications are adopted here. 4.4.3.1 Axial Momentum An axial momentum balance for a control volume enclosing a single stream-tube between station 0 and station 2 yields: BFz = V2 V0 m m = [(V0 wz ) cos wr sin ] 2rs (4.4.2) (4.4.3)

1 Note that for clarity w is used in the momentum equations, and u is used in reference to the results of vortex theory. 2 A variable area/expanding stream-tube will have non-zero integrated pressure forces on the streamwise sides, although Mikkelsen [89] showed these to be negligible.

50

Chapter 4 Analytic Development

Note that the mass ow is calculated as the dot product of V1 and area normal vector dA = 2rs. The standard BEM formulation assumes equivalently wr = 0 or stream-tube cross-sectional area of 2rs cos perpendicular to the total ow at station 1 (V0 wz ). Both formulations overestimate the mass ow which is relatively reduced for a non-zero , due to the radial inclination of the ow.

Figure 4.4 Flow relative to airfoil section

4.4.3.2 Tangential Momentum The tangential velocity imparted to the ow as a torque reaction is calculated from a tangential momentum balance: BrF = mr2 w,2 (4.4.4)

The notation w refers to azimuthal velocity at the disc and w,2 azimuthal velocities downstream. As with the rest of the method development, the basic theory essentially ignores wake expansion and assumes an innite number of blades. Considering the reference frame xed to the blade section shown in Fig. 4.4, the average tangential velocity w through the rotor is half the tangential exit velocity 2w , since there is no tangential velocity upstream of the rotor. With no wake expansion (r1 = r2 ), conservation of angular momentum yields:
2 2 2w r1 = w,2 r2 w,2 = 2w

(4.4.5)

4.4 Aerodynamic Modelling 4.4.3.3 Finite Length Blades

51

A nite number of blades B will produce distinct helical vortex sheets all moving downstream with the same velocity, rather than a continuous volume of vorticity. For nite numbers of blades, Prandtl [98] developed a simplied wake structure model idealized as semi-innite discs moving downstream at speed V2 , spacing the idealized discs at the pitch distance of the helical sheets. The average velocities in the wake V2 and w,2 are then predicted from complex-variable ow analysis around the discs, in the form of a correction factor F to the far-eld velocities: F = wz,2 2 B Rr cos1 exp 2 r sin() = F wz,2 (4.4.6a) (4.4.6b) (4.4.6c)

w,2 = F w,2

This accounts for the average momentum transferred to the bulk uid relative to the uniform ow produced by an innite number of blades. Thought of another way, the velocity at the blades (the velocity of the wake sheets) is higher than the average ow between the sheets near the edges. Some researchers have therefore also applied F to the induced velocity at station 1 in the mass ow term (as a circumferentially averaged quantity). This is not convention1 [104, p.55] or done here, because Prandtls derivation requires an innite series of discs which is not a good approximation in the near wake. Prantls model should strictly use the helix angle at the tip, tip . For the coned rotor, should be the wake helix angle in a plane tangent to the rotor axis, not normal to the blade pitch axis. Again, this detail is lost in the overall approximation, so at each section normal is used. 4.4.3.4 High Induction The BEM model breaks down for high a values (occurring equivalently at high and low V0 for xed ). In reality, unsteady turbulent mixing occurs with the ow outside the bounding streamline downstream of the rotor, to satisfy mass continuity (i.e. the velocity cannot be zero in the wake as predicted by basic momentum theory). To account for this eect, the stream-tube axial momentum balances are modied by empirically derived thrust coecients CT,model , determined experimentally from aggregate values for complete rotors.
BLADEDTM uses a not aF in the mass ow term. Note that some authors may also invert the denitions of a and aF by dening a as the averaged quantity.
1

52

Chapter 4 Analytic Development

Three existing models (referred to later as thrust models) that have been tted to experimental data are used here, when the induction factor in the plane of the rotor tip adisc exceeds ac : CT,Bladed (ac = 0.3539) = 0.5998 + 0.61adisc + 0.78a2 disc CT,Glauert (ac = 0.4) = (adisc 0.143)2 0.0203 0.6427 CT,Spera (ac = 0.2) = 4 a2 + (1 2ac ) adisc c + 0.889 (4.4.7a) (4.4.7b) (4.4.7c)

The rst CT,Bladed model, implemented in BLADEDTM , is quite similar to the Glauert model [17]. Results presented later obtained from runs of the commercial code are labelled BLADED. The same thrust model is used internally in the implementation of the current work, for which results are labelled CT Bladed, for comparison to the CT Spera results using the Spera model [15]. The behaviour of the three models is compared in Fig. 4.5, plotted with the original data Glauert used to derive his model [15], which will become relevant later.
3.0 2.5 2.0

CT

1.5 1.0 0.5 0.0 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

a
Momentum Glauert NACA MR No L5009 BRC R&M 885 BLADED Spera NACA TN 221

Figure 4.5 Thrust CT,model model behaviour Again, for the unconed rotor these models should strictly use adisc = aF not adisc = a to account for the average velocity at the disc, as the data was gathered based on an average velocity. As in 4.4.3.3, this is again not done to follow convention [16, p. 93]. However, one must be careful in comparing data as some authors choose a and others aF .

4.4 Aerodynamic Modelling

53

4.4.4 Blade Elements


Momentum theory alone cannot predict the forces imparted by a blade with particular airfoil cross-section proles. However, using the predicted induced velocities at the rotor, 2D airfoil data can be used to calculate the loading imparted to the stream-tube in the momentum equations. Prior to an understanding of the induced velocities produced from 3D wake structures, there was great debate as to the ow conditions at the rotor suitable for determining the aerodynamic forces [98]. Blade element methods which ignored any inow eects (i.e. assumed unaltered V0 velocity at the blades) were used with apparently heuristically derived airfoil datasets, each diering in the aspect ratio of the blade, in a misguided attempt to account for the actual 3D ow eects. Lifting line methods can now of course explicitly account for the velocities induced at the blade by the vorticity trailed downstream. The closure of this problem is left to the next section. For now, the generalized ow relative to a blade section shown in Fig. 4.4 is considered. Spanwise ow cannot be directly modelled by BEM. However, it will serve to further force the un-modelled 3D boundary layer and may exacerbate stall-delay eects from spanwise boundary layer migration [105] (see 4.4.9). The aerodynamic forces fz and f per unit length of blade are predicted from 2D airfoil test data (lift/drag coecients cl , cd ) dependent on AOA and Reynolds number Re:
2 Vrel c cos 2 V 2 f = c rel c 2 cn = cl cos + cd sin

fz = cn

(4.4.8a) (4.4.8b) (4.4.8c) (4.4.8d)

c = cl sin cd cos The total angle of the section tot is found from: tot = twist + set + pitch

(4.4.9)

with twist the blade twist, set the xed pitch oset, and pitch any active pitch angle. The AOA is then = tot . Lindenburg [104, p. 57] notes that the eective camberline of an airfoil section on a coned rotor will be modied, due to the curvature of the ow relative to the airfoil. The eects may be included by referencing the AOA to the 3/4c point on the section and modifying the AOAcm relation. This eect is ignored here however, given the spatial resolution of the BEM method.

54

Chapter 4 Analytic Development

4.4.5 Wake Analysis


The momentum equations must be closed by relating the far eld velocities to those at the disc. As previously mentioned, the detailed ow at the disc is required to compute the aerodynamic forces developed by the section. This is normally done via Bernoullis equation, relating the dierence in static pressure just before and after the rotor (equivalent to Fn /dA) to the upstream and downstream velocities. This again neglects wake rotation and expansion, but more importantly is the root of an inability to predict coning eects. Essentially, as has been noted in full ow eld studies [88, 106], standard BEM behaves as if the calculations are done in a radial plane through the blade tip. In essence they are, since information about the relative stream-wise position of the stations along the blade is absent. An alternative approach to closing the equations is to examine the wake structure in more detail using potential ow theory. It is of course possible to dispense with the momentum equations altogether, by modelling the blades and wake as bound and free vortex laments to determine the ow at the rotor (bound vortex), but this is overly computationally intensive for design purposes. To exploit the computational eciency of BEM, analogous with fundamental wake studies [107], yawed turbines [100] and helicopter rotors [102], the helical vortex sheets shed into the wake can be decomposed into rings and axial laments extending innitely far downstream. 4.4.5.1 Wake Decomposition The simplest case is obtained by restricting attention to the uniformly loaded rotor (constant pressure exerted on the ow by the disc) with an innite number of blades. Vorticity is continuously shed only at the blade tips to form a single sheet of vorticity, as shown in Fig. 4.6. The bound vortices on the disc also merge into a single axial root vortex joining from downstream at the blade roots. Note that in general as the loading along the blade changes for a real rotor, vortex rings and laments will be shed to ll the entire downstream volume with helical sheets. The Biot-Savart law is used to compute the induced velocities u for a vortex lament of strength and length dl with vector r to the point of interest: du = dl r 4 |r|3 (4.4.10)

Using this formula, it is shown later in 4.4.7 that the axial laments on the wake cylinder and at the root, together with the bound vortices on the disc, induce only second-order velocities. As in other works, the streamwise (axial) [100] and radial

4.4 Aerodynamic Modelling

55

Figure 4.6 Wake decomposition into rings and laments

induction eects are of primary importance here. The swirl is of course treated in the conventional manner, by the azimuthal momentum equation. Provision for further treatment is provided by the inclusion of

in the nal set of equations. The

correction factor is dened from the vortex lament results as:

u,2 2u

(4.4.11a)
vortexsheet

w,2 = 2a r

(4.4.11b)

In any case, swirl is irrelevant to the solution of the uniformly loaded disc, since only the axial momentum equation is solved. The axial and azimuthal equations are uncoupled and the latter uniformly zero, due to a lack of azimuthally applied pressure. For a real rotor, the loading (including an azimuthal loading) will of course depend on , since the lift and drag depend on the angle of attack which is determined by together with the blade twist and pitch. Again using lifting line theory, it can be shown that the bound vortices on the disc do not contribute to the induced velocities on the disc in the case of an unconed rotor. For a coned rotor, the axial and bound laments only induce azimuthal velocity, as for the unconed rotor. The net eect is that the azimuthal induction is changed less than 1% over the length of the blade for a coned rotor relative to an unconed rotor (see 4.4.7 for further details). In summary, inclusion of the swirl correction factor for coning

will have zero eect

for a uniformly loaded rotor and a second order eect for a real rotor.

4.4.5.2 Evaluation of Correction Factors Focusing back on just the vortex cylinder comprising the wake, the induced velocity from the cylindrical vortex sheet in Fig. 4.7 at point x0 , 0, z0 is found from integration of Eq. (4.4.10) over the surface of the cylinder with vortex strength per unit

56 area = /dzRd:

Chapter 4 Analytic Development

u=2
0

R 4
z0

z cos + z sin + (R x0 cos ) k i j x2 2Rx0 cos + R2 + z 2 + (c R)2 0


3 2

dzd

(4.4.12)

The point is x0 units radially o the centreline and z0 units upwind from the end of the vortex tube. The inner z portion of the integral is easily evaluated, however the azimuthal component requires either elliptical integral functions or as is done here, numerical integration. The z limits of integration may be set to to evaluate the induced velocity in the far wake. The vortex core addition to the formula, c R, avoids a discontinuity as the sheet is approached.1 The core size c R is taken as 0.1%R, with minimal eect on the results below this value.

Figure 4.7 Vortex ring geometry Correction factors in the far wake:
z

and

are calculated using the induced velocities from

Eq. (4.4.12), computed for each section radius at a point on the coned disc and

uz,2 2uz vortexsheet ur = uz vortexsheet

(4.4.13a) (4.4.13b)

The division operations cancel the vortex sheet strengths in Eqn. Eq. (4.4.12), which can be arbitrary here. It should be possible to bypass the use of the momentum
This is a numerical artefact of the mathematical model. Viscosity has important eects near the centre of the vortex, leading to various models to describe the induced velocities as the core is approached [108].
1

4.4 Aerodynamic Modelling

57

equations, by continuing the vortex analysis and computing the sheet strengths, rather than the induction factors. This method would amount to a prescribed wake vortex approach, and was not explored here as the purpose was to provide small corrections to well known and trusted BEM codes. Figure 4.8 shows the vortex sheet predictions for a unit radius rotor operating at a = 1/3. Only a single cylindrical vortex is shed at the tip . The vectors show the induced velocity from the vortex sheet, the grey streamlines the overall ow, and the ellipses centred on the tip are isolines of the induced radial velocity. An unconed rotor is shown, but the same diagram will exist for a unity radius rotor with coning (i.e. a longer blade). The axial correction factor radial correction factor
z

is greater than unity for points

upstream of the tip-plane, unity at the plane and less than unity downstream. The
r

is always greater than zero and bounded to approximately

4 near the wake sheet (depending on vortex core size).


2.0

1.5

1.0

0.5

0.0 1.0

0.5

0.0

-0.5 z

-1.0

-1.5

-2.0

Figure 4.8 Velocity eld around vortex cylinder

4.4.6 Relating Disc to Far Field


Using the vortex lament results (
z

and

r)

at each section along the blade, the far

eld velocity may now be related to that at the disc for closure of the momentum equations and denition of the induced velocities. The following equations dene

58 the induced velocities in terms of


z

Chapter 4 Analytic Development and


r: z z

wz,2 = 2wz = 2aV0 wr = wz


r

(4.4.14a) (4.4.14b)

For an unconed rotor, the vortex calculations yield a unity axial correction factor
z.

This is consistent with the usual assumed result from Bernoullis equation,

wz,2 = 2wz . The inclusion of a radial component in the ow modies the velocity normal to the blade in calculating the lift and drag coecients. The normal velocity seen by the blade section can now be dened using the radial correction factor: V1 = V0 (1 a)2 + (a r )2 Vn = V1 cos ( + ) = arctan a r 1a
1/2

(4.4.15a) (4.4.15b) (4.4.15c)

4.4.7 Azimuthal Induced Velocity


The axial laments on the wake cylinder sheet, the root streamwise vortex, and the bound vorticity on the blades shown in Fig. 4.9 are responsible for all azimuthally induced velocities. Qualitatively, the velocity vector increments at the blade induced by the laments can be worked out graphically, using the right-hand rule to determine the directionality of the induced velocities. Considering each of the three contributors in turn, it can be demonstrated that the net streamwise and radial induced velocities are zero for each set, independent of cone angle. The azimuthal contribution, however, is non-zero for the wake and root laments for an unconed rotor. Additionally, the bound vortices have a net contribution for a coned rotor. The error associated with ignoring the details of the azimuthal induction can be quantied using the Biot-Savart law Eq. (4.4.10). For any straight vortex lament of strength , dening r1 and r2 as vectors from the tail and head of the lament to the point of interest respectively, the induced velocity may be calculated from: u= (|r1 | + |r2 |) (r1 r2 ) 4 |r1 | |r2 | (|r1 | |r2 | + r1 r2 ) + 2 (4.4.16)

The factor again ensures that the formula converges near the vortex core. In order to quantify the induced velocities, the magnitude of must be determined for a given operating condition consisting of free-stream velocity V0 , axial induction

4.4 Aerodynamic Modelling

59

Figure 4.9 Axial, root and bound vortex laments

a and tip-speed ratio , as done by Burton et al. [16]. Considering the wake cylinder opened at, over one revolution of the wake the sum s of all the trailing vortex laments will be distributed over a distance 2Rsint perpendicular to the laments. For an innite number of blades, this gives an areal density of vortex strength of: w = s 2Rsint (4.4.17)

The angle t is the ow angle at the tip, which from geometry may be determined from: tan t = 1a (1 at ) (4.4.18)

The vortex density normal to the streamwise direction is a component of w from geometry, = w cos t . As will be shown later, in the far wake this induces a uniform axial velocity of k. This induction is equal to 2aV0 from classical theory, yielding: s = 4V02 a(1 a) (a + at ) (4.4.19)

where at is the tangential induction factor at the tip. For the unconed rotor, only the root vortex of strength = s contributes to the tangential velocity, so by denition of at and Eq. (4.4.16): at = 4R2 (4.4.20)

60

Chapter 4 Analytic Development

Combining Eqs. (4.4.19) and (4.4.20) we obtain a quadratic for at : at + at


2

a(1 a) =0 2

(4.4.21)

This may be solved for at for the given ow condition, and then substituted into Eq. (4.4.19) to yield . Given and using Eq. (4.4.16), the contributions to the net induced velocity at any point in the ow may be determined. The tangential induction factor at points along the blade at radius r can be computed from: at = u R rV0 (4.4.22)

where u is azimuthal induced velocity and R the tip radius. The results shown in Fig. 4.10 are for a typical case of a = 0.3 and = 6, with = 0.001R. 100 blades are used for the numerical calculations. The tip radius is the same for the coned and unconed cases. For these results, the above relations for are assumed valid for a coned rotor. As will be shown, Eq. (4.4.20) is minimally in error due to the additional inuence of the bound vortices on a coned rotor. As expected, only the root vortex has any inuence at the disc for = 0 . The dip at the tip is due to the vortex core model. For the downwind coned rotor, the bound vortex has an inuence in opposition to that created by the root vortex. However, the inuence of the root vortex is increased, as the blade now sees more of the vortex. For the upwind coned rotor, the eects are reversed. The net eect in both cases is that the azimuthal induction is changed less than 1% between the coned and unconed cases. These results lends credence to the decision to ignore the details of the azimuthal induction in the coning BEM theory.

4.4.8 Stall Delay and Centrifugal Pumping


It has long been recognized that propeller and wind turbine blades behave as if sections near the root of the blades have delayed stall, to higher cl (up to double) and stall , than would be predicted by 2D section data. It has also been noted that a much more benign and opposite eect appears to exist near the tips. Lindenburg [104, 4.2.2] provided an excellent review of the relevant studies that have examined this eect. As the eect has only been observed on rotating blades, it is variously referred to as centrifugal pumping, spanwise pumping and radial pumping. Section 4.4.8.1 rst gives an overview of the experimental challenges in studying stall delay, followed by a comparison of the various theories that have been proposed to

4.4 Aerodynamic Modelling


4 3 2 bnd cyl root total 4

61

1 0 -1 0 0.2 0.4 0.6 0.8 1

a
1 0 -1 0 0.2 0.4 0.6 0.8 1

r/R
(a) = 0
4

r/R
(b) = 40

a
1 0 -1 0 0.2 0.4 0.6 0.8 1

r/R
(c) = 40

Figure 4.10 Azimuthal induction along blade

explain the phenomenon in 4.4.8.2. Based on this background, 4.4.8.3 selects and presents the best existing model for use in this thesis. 4.4.8.1 Experimental Investigation Post-processing experimental data to validate the occurrence of the phenomenon is complicated [109]. Measurements of aggregate shaft torque/power mask the details of any spanwise variation. A common method of investigation is to employ pressure taps at sections along the blade. Unfortunately, the spatially resolution is fairly coarse, and the angle of attack must be computed in some manner for decomposing

62

Chapter 4 Analytic Development

the normal cn and tangent ct force coecients into conventional aerodynamic cl and cd ones. To avoid the latter ambiguity, some data reduction (and models in 4.4.8.2) present the data in terms of cn [104]. Others make use of full vortex analysis models to back-out the angles of attack and thereby decompose the measured pressures into lift/drag coecients. This has produced conicting results, particularly as some studies note an increase in cd while others nd a decrease. There is no doubt, however, that the eect is real and particularly pronounced at the root.

4.4.8.2 Mechanisms for Centrifugal Pumping Stall delay is clearly a boundary layer eect, whereby gross separation is delayed and stall avoided. Based on this, various researchers have employed the NS equations in an attempt to tease out the underlying mechanism of stall delay. The earliest collection of eorts [110] reduced the equations to their laminar boundary layer form. In fact, at its most reduced, no stall delay is predicted for c << r. More recent eorts, as detailed and expanded on by Corten [105], have emphasised the use of order-of-magnitude analysis for eliminating terms in the NS equations. This reference should be consulted for more information on the following discussion. Snel [111] employed this technique [112], embedding the equations into X-Foil1 to solve for enhanced lift characteristics. The theoretical examination of the boundary layer equations predicted a trend with (c/r)2/3 , however the calculations found a relation of (c/r)2 . Corten [105] disagreed with a number of assumptions in the reduction, fundamentally whether the use of the boundary layer equations was justied in stall. A revised analysis was framed in a cylindrical CS rotating with the blade, without the boundary layer assumption, but neglecting viscosity (reduction to the Euler equations). Emphasis was placed on a separation point existing on the blade surface, so that the separated boundary layer is thick and velocities negligible. The nal result is a prediction that stall delay eects are proportional to c/r. At issue between the two methods is the underlying mechanism. In the former, it is a centrifugal pressure gradient owing to the rotation, which sets up a radial ow. In turn, the radial ow creates a positive chordwise pressure gradient, owing to Coriolis acceleration in the rotating reference frame. This reduces the adverse pressure gradient and hence delays stall. Corten [105] on the other hand, nds that
1

2D coupled inviscid-viscous boundary layer code by Dr. Drela at MIT

4.4 Aerodynamic Modelling

63

radial convection is the underlying mechanism. A related phenomenon predicted by the latter approach is a helical ow outwards in the separated region near the TE. It seems somewhat duplicitous to simultaneously assume a separated region of ow and delay of stall, since separation is the mechanism of airfoil stall. If a stall region extends from the TE, then a drag increment should be more evident in the results. This would arise both from a chordwise pressure dierential, and energy input to the large volume of separated ow. Ultimate resolution of the theoretical discrepancy will depend on more experimental data becoming available. 4.4.8.3 Modelling Centrifugal Pumping In an attempt to match analysis to experimental data for the National Renewable Energy Laboratory (NREL) Unsteady Aerodynamics Experiment (UAE) experiment (see C.2), Lindenburg [104] compared the available models based on the preceding analyses. A new model was also proposed by Lindenburg [113], based on the separation point at the TE, together with a set of assumptions about fully-developed spanwise ow at the TE creating a Coriolis acceleration. The lift increment is then taken as proportional to the chordwise extent of the separation region. The other models propose modications to cl , cd and cn in various forms. Based on the accuracy of the predictions in that report, relative to the experimental data, the modied form of Snels model is used in the current work. As outlined in 4.4.8.2, the Snel model also appears the most physically based. It is expressed as a cl increment from the static non-rotating coecient cl,nonrot : cl,rot = cl,nonrot + 3.1 cos2 c r
2

(2( 0 ) cl,nonrot )

(4.4.23)

The factor of 3.1 was obtained empirically. The cos2 local tip-speed ratio r dependency was introduced after the original derivation, to enable analysis during start-up and shut-down. BLADEDTM can only accept static corrections to airfoil tables computed a priori, so this model cannot be used in BLADED runs. A number of additional features of the model implementation are given in the following list: The maximum lift increment is limited so that cl,rot never exceeds 2( 0 ), the linear lift curve. A linear reduction between r = 0.75 . . . 0.8 to zero limits the spanwise extent of the model to be consistent with experiments.1
The linear reduction (rather than a sharp cut-o) is required to avoid numerical diculties in optimization as sections pass the cut-o radius under DV manipulation.
1

64

Chapter 4 Analytic Development The model is only applied for > 0 . It is also linearly tapered to zero over 30 < < 50 . This ensures stall occurs at some point over that range, consistent with experimental evidence.

4.4.9 Spanwise Flow


By virtue of its geometry, a coning rotor will encounter a velocity component Vs along the spanwise axis, as shown in Fig. 4.11. Any rotor operating in yaw or ow non-parallel to the shaft will also experience spanwise ow. The eective sweep angle over the blade is shown in Fig. 4.12, for a range of coning angles at zero yaw. Two operating points are shown, one representative of optimal operation (Fig. 4.12(a)) and the other power limiting (Fig. 4.12(b)).

Figure 4.11 Skewed ow velocity vectors and boundary layer proles The induced velocities (axial and radial) have been computed using the correction factors developed in 4.4.5.2. The induction factor corrections have been computed for an innite number of uniformly loaded blades, so the only limit on induced velocity close to the tip is the vortex core size. Consequently the eective sweep angle is found to rise sharply towards the tips. In reality, a nite number of blades will reduce this large increase seen over the last 5% of the radius. Due to the ow expansion predicted by the vortex model, even an unconed blade will experience some degree of spanwise ow. In this case though, the angles are quite small, even

4.4 Aerodynamic Modelling


40 30 (deg) (deg) 20 10
1 10 2
8

65

40 30 20 10
6 18 1 6 14 12 1 0 8 6

0 40 20 (deg)

4
2

0 40 0.4 r 0.6 0.8 1 20 (deg) 0

0.2

0.8 1 0.4 0.6 0.2 r

(a) a = 1/3, = 7

(b) a = 2/3, = 3

Figure 4.12 Eective sweep angle with coning for typical operating conditions

in the stalling condition Fig. 4.12(b), and so would not have much of an eect based on the experimental evidence to follow. Conversely, the increasingly large seen over the blade length, particularily towards the root as the cone angle increases, can be expected to inuence the stalling behaviour of the rotor. This section rst examines the common assumptions regarding spanwise ow in 4.4.9.1, before referencing experimental evidence in 4.4.9.2 to show that the common assumptions break down in stall. To deal with the conditions of the coned rotor, a model is developed in 4.4.9.3 that is used to modify the 2D airfoil characteristics to account for spanwise ow. 4.4.9.1 Independence Principle Almost universally, the spanwise ow component is ignored when computing the aerodynamic properties of the section. It is assumed that the lift and drag of the section may be predicted solely from Vrel and , computed as V cos . This approach is justied by the independence principle [114], also applied to swept wings and helicopter rotors, and which may be rationalized by the following thought experiment. Consider an innitely long wing (unbounded zaero ), of constant cross section, immersed in a frictionless ow. If the ow is at right angles to the wing axis ( = 0), then the ow around the wing is uniquely determined by the cross-sectional prole normal to the spanwise axis. If the wing is now assumed to translate along its own axis, the lack of friction means that the external ow is unaltered from the original

66

Chapter 4 Analytic Development

condition. It is also equivalent to an eective yaw angle being experienced by the wing. The thought experiment of course breaks down when the viscosity of the ow becomes important, i.e. when the boundary layer is no-longer thin and separation (stall) begins to occur. The laminar boundary layer proles ([see 110]) before stall are illustrated in the upper right corner of Fig. 4.11, with a greatly exaggerated dimension normal to the airfoil surface. Simple wing sweep theory simply relates angle of attack , lift curve slope cl and lift coecient cl, for a section in the streamwise direction, to the analogous parameters in the section normal to the span axis, , cl and cl : = cos cl = cl cos cl, = cl cos2 (4.4.24a) (4.4.24b) (4.4.24c)

These formulae can be derived from the geometry of Fig. 4.11, decomposing the rotation vector lying along the spanwise axis into a component normal to the freestream, and recognising that c s = c cos b/ cos = c b, i.e. the non-dimensional areas are equal for both un-swept (chord c, length s) and swept (chord c , span b perpendicular to free-stream) denitions. 4.4.9.2 Experimental Evidence Experimentally, the validity of the independence principle has been conrmed for real wings, at least before 2D results would predict stall. In the set of swept wing test data to follow, sweep angle and yaw angle (in an aircraft sense) are synonymous with each other. Figure 4.13 shows a typical set of 2D lift curve data (i.e. for sections perpendicular to spanwise axis), derived from a 3D yawed wing test. Another set of 2D lift data shown in Fig. 4.14 has been measured from pressure taps on two swept wings, one without twist and employing a symmetric airfoil, and the other with twist and a non-symmetric airfoil [115]. The various lift curves are labelled by the section percentage location along the span. The data is somewhat complicated by the induced velocities from the nite wing length.1 The eect is to alter the linear lift-curve slope dierentially along the span, as well as shift the zero-lift AOA for the non-symmetric wing.
Secondarily, a fuselage was also present at the root of the wing, and the pressure taps at the most inboard and outboard locations were orientated in the free-stream direction. The rest of the mid-span pressure taps were located along chords perpendicular to the swept wing axis. The tests were also carried out for Re 108 .
1

4.4 Aerodynamic Modelling


2.4 2.0 1.6 1.2 0.8 0.4 0.0 0 0 35 30 15 40 75 60 45

67

cl

10

20

30

40

50

AOA (deg)

Figure 4.13 Experimental 2D lift curves for varying yaw angle (data from Harris [110])

Experimental 2D lift curves for = 0, shown as the dashed lines in Fig. 4.14, where used to compute estimated lift curves based on the simple sweep theory given in Eq. (4.4.24). Note that AOA is the overall wing angle of attack for this dataset, not the sectional angle of attack . A Weissigner lifting-line method was then used to compute linear lift-curve slopes which were then used to adjust the 2D curves. The predictions are not shown in Fig. 4.14 for simplicity, but suce it to say that the linear portions matched well, but the predicted stall behaviour was akin to the 2D curve in terms of stall AOA and cl,max . The measured data therefore shows signicant stall delay, increasing in eect inboard. Based on data such as this, it has been found that a yawed wing will follow the 2D linear lift curve well past the 2D un-swept stall angle and lift coecient cl,max ( = 0), delaying stall to much higher lift coecients cl,max () and stall
AOAs. Data points taken from a number of studies for cl,max and stall are given in

Fig. 4.15 illustrating this behaviour.1 The proposed mechanism for the observed stall delay is a thinning and energization of the boundary layer by the spanwise ow. The boundary layer is therefore able to continue to navigate the adverse pressure gradient without separation for much higher 2D AOAs . Furthermore, it has been found experimentally that the 2D drag and moment coecient curves are basically unaected by . Of course, energy is
The additional data in Fig. 4.15(a) for the NACA study was obtained from full-wing CL data, scaled using Eq. (4.4.24c).
1

68
1.6 1.4 1.2 1.0 0.8 0.6 0.4 0.2 0.0 0 10 20 2D 0.815 0.167 0.383

Chapter 4 Analytic Development


1.6 1.4 1.2 1.0 0.8 0.6 0.4 0.2 -0.2 -5 0.0 5 15 25 0.383

0.545

0.167 0.545 0.707 2D 0.815 0.924

cl

0.924

0.707

cl

AOA (deg)
(a) Plain wing

AOA (deg)

(b) Twisted wing

Figure 4.14 Wing section lift curves for 45 swept wing (data from [115])

being put into the boundary layer to avoid 2D stall, and so increased drag is present, but in the spanwise direction.
3.0 2.8 2.6 Harris Harris 2 NACA cos^1.5 cos^1 cos^0.75 45 40 35

cl,max()/cl,max =0

2.4 2.2 2.0 1.8 1.6 1.4 1.2 1.0 0 10

20

30

40

50

60

70

AOAstall

30 25 20 15 10 0 10 20 30 40 50 60 70

Sweep angle (deg)

Sweep angle (deg)

(a) Maximum lift

(b) Stall angle

Figure 4.15 Stall delay behaviour for 2D sections in yawed ow (data from [110, 116])

It should be noted that this spanwise drag component is quite important for helicopter rotors [110, 117], since it translates into an additional power requirement for forward ight. Helicopter blades experience continually varying yaw with azimuth

4.4 Aerodynamic Modelling

69

angle, owing to the relative component of the free-stream in forward ight (greater angles towards advancing side, [see 110, Figure 2]). In aircraft applications [118, pg. 252], stall is generally avoided, so only the linear theory (below stall) is presented. Further comments are usually directed towards the mitigation of tip stall (washout, vanes), resulting from outboard migration of the boundary layer on swept wings, which precipitate dangerous pitch-up and loss of roll control. This migration results from a spanwise pressure gradient developed between sections normal to the free-stream [116], since the airfoils in those planes are displaced relative to each other in a streamwise sense. Assuming the pressure proles are identical, there will therefore be a pressure dierential between sections. Another important eect with relevance to aircraft is vortex lift. As noted by Furlong and McHugh [116] in an overview report of spanwise ow, much larger cl improvements are found for airfoils with sharp Leading Edges (LEs), which stall from the LE, rather than TE for airfoils with more rounded LE. The mechanism is a stable vortex [118] created over the upper surface, near the LE. The lift increase associated with this ow structure was included as a safety factor in the original CONE-450 study [66] for parked calculations. It is not clear whether or not such enhanced vortex lift would in fact develop, given the negative sweep angle formed by the thinning chordwise distribution.1 No prediction of eects from spanwise ow in operation was included in the original study. For wind turbines, it is the enhanced lift and delayed stall eects that are critical. Predictions in yaw above stall remain within the realm of experiment, defying an easy analytic solution [110]. Boundary layer theory can be applied to the chordwisespanwise coupled problem, but is only valid to the lower limit of stall initiation2 and available in closed form only for laminar ow. Most data obtained for aircraft design has looked at wings with relatively low aspect ratio (average turbine blades (approaching

A = 5) [116].

More experimental eort to explore these eects for the very high aspect ratio wind

A = 25) is required for validation. It is expected that

the eects may be even more pronounced, given the larger span absent of tip-eects, over which the spanwise ow can develop.

1 This is opposite to the conventional delta wing that increases in span (wind turbine blade chord) in the ow direction. 2 The same is true for predicting the stall delay eects of centrifugal pumping discussed in 4.4.8.

70 4.4.9.3 Spanwise Flow Model

Chapter 4 Analytic Development

Below the point of 2D stall, sectional data may be used in the BEM method with some condence. There is clearly no fully satisfying analytic method available for predicting the 2D properties beyond the 2D stall point when spanwise ow is present. In the face of this uncertainty, Harris [110] originally proposed a simple approach, beginning with the assumption that a section in the streamwise direction will stall at approximately the same ,stall as a 2D section without sweep stall . This was rationalized from transforming Fig. 4.13 to a free-stream denition, and noting that the cl,max values were somewhat close (0.85 . . . 0.95 up to = 45 ) and the lift curves linear to 11 . . . 12 . From Fig. 4.15(a), curves were t through the single data point marked with a triangle (1/ cos0.75 and 1/ cos1 relations). The rst was postulated as relationship for cl,max () for turbulent conditions, and the latter for laminar conditions, both being somewhat consistent with the original assumption. The theory is then left for further investigation, and based on the sparse dataset in that study the curve ts were not overly accurate. Moreover, what is required is some prediction of the airfoil behaviour over the complete range of found in a BEM simulation. To that end, new curve ts for 1/ cos1.5 are shown in Fig. 4.15, for both lift coecient and 2D stall angle. This relation ts the available data better, even if no better theoretical basis can be oered. By scaling both cl and by the same ratio, the linear portion of the lift curve is left un-altered. This is important, as it is known that performance in this region is in fact unaltered by spanwise ow. The scaling for spanwise ow is therefore implemented as given by Eq. (4.4.25): () = 0 + ( 0 )sf cl () = cl sf sf = r r + 1 Rtip Rtip cosn (4.4.25a) (4.4.25b) (4.4.25c)

The exponent n is taken as 1.5 based on the experimental data presented in 4.4.9.2. The eect is tapered based on radius towards zero at the tip. Using the experimental data for = 0 given in Fig. 4.13 and Eq. (4.4.25), corrected data up to = 45 is shown as dotted lines in Fig. 4.16. The full set of experimental data from = 0 is also shown as solid lines. The scaling seems reasonable, at least up to 30 . Beyond this angle, the delay in stall angle prediction degrades, although cl,max is predicted well. A much more involved model would be

4.4 Aerodynamic Modelling

71

required for a better prediction, but with such a limited dataset would be hard to synthesize in a generic and computationally simple manner.
2.0 1.6 1.2 45

cl

40 35 0 15 30

0.8 0.4 0.0

10

15

20

25

30

AOA (deg)

Figure 4.16 Sectional data corrected for spanwise ow

With no direct experiments available, it is unclear whether it is valid to simultaneously apply spanwise ow models for both sweep eects and the centrifugal pumping discussed in 4.4.8. Both produce similar eects, one from a direct velocity component and the other owing to a rotating reference frame. Since the underlying mechanisms both energize the boundary layer, albeit from dierent sources, it stands to reason that there is at least some degree of synergy and additive eect. Furthermore, it is conservative from both an overall force and power limiting capability perspectives to analyse the eect of considering both simultaneously. 4.4.9.4 Dynamic Spanwise Flow It is important to remember that helicopter rotors and wind turbine blades experiencing yawed ow will experience dynamically changing sweep and AOA (see 4.4.14). These dynamically changing ow conditions can be expected to deviate from the steady-state behaviour just discussed [117]. Based on some experimental evidence, Leishman [119] concludes that the normal force coecient cn for a swept wing in oscillating pitch does not attain a larger maximum, but does produce a higher average value with the eventual dynamic stall delayed to larger AOAs, compared to an un-swept wing. The net eect is an increase in rotor thrust with sweep angle.

72

Chapter 4 Analytic Development

Leishman [117] concludes that although applying the static corrections just discussed is predicated on dierent fundamental mechanisms, the net result is representative of the experimentally observed behaviour. Fortuitously, purely coned rotors in a uniform un-yawed ow experience a constant sweep angle and so the assumption of constant sweep angle is more accurate. Of course, the spanwise velocity will vary along the span and is thus dierent than the swept wing case.

4.4.10 Closed Equations


Combining the developments of previous sections, the momentum-derived thrust coecient CT,mom (from Eq. (4.4.2)) and the coecient obtained from blade element forces, CT,loc may be derived: CT,mom = 4aF CT,loc [(1 a) a r tan ] Vn = cn 2 2 V0 sin
z

(4.4.26a) (4.4.26b) (4.4.26c)

= cn

cos2 ( + ) (1 a)2 + (a r )2 sin2

The manipulation from Eq. (4.4.26b) to Eq. (4.4.26c) is possible by vector geometry visible in Fig. 4.4. Together with the tangential momentum balance Eq. (4.4.4), the nal set of equations is: anew = c1 (1 + c2 ) + c3 1 + c1 cn H cos2 ( + ) c1 = sin2 4F z c2 c3 (a r )2 = 1a a2 r tan = (1 a) Bc = 2r 1 = 4a(1 a) 4a2 z CT,model (4.4.27a) (4.4.27b) (4.4.27c) (4.4.27d) (4.4.27e) adisc ac
r

tan

adisc > ac

(4.4.27f)

anew =

2 Vrel c 4F V0 [(1 anew ) cos anew

sin ] r

(4.4.28)

The set of equations is now closed and allows an iterative solution to be obtained using xed-point iteration to determine anew = f (a, a ) and anew = f (a, anew , a )

4.4 Aerodynamic Modelling

73

from the current estimates a and a .1 To handle parked and very low induction cases, a and a are blended to zero over local tip speed ratio (r) = 1.5 . . . 0.5, after limiting a and a to 1 . . . 1. The deviations from the standard BEM are in the
z

and

terms calculated

a priori with vortex theory, and the inclusion of proper induced velocities in the various equations. As mentioned before, the details of swirl as predicted by vortex theory are ignored in the present work, so

is set to 1.

Some authors [16] argue that the cd term should be zero in computing the induction factors. The rationale is that as only pressure drag contributes to the velocity decit in the overall ow, the velocity decit owing to shear on the airfoil is conned to the wake sheet. This argument breaks down for the stalled airfoil, and so cd is generally included in modern codes and ExcelBEM. The velocity in the tip-plane adisc =
za

derives from the vortex results for the

wake geometry discussed thus far, and is assumed valid in the modications considered in the following sections. The induction in the tip-plane is used as the criteria, rather than the induction on the blade, to be consistent with the denition of the original experimental data. As the CT,model equations no longer intersect the momentum-derived CT,mom function curve, a smooth blending function over a = 0.05 around ac is used to transition between the momentum and experimental CT models. It was found in some cases (equivalently high a, CT or ) that convergence of the xed point iteration was problematic. This was found to be a result of Eq. (4.4.14), which ties radial to axial induction and in these specic cases computed unrealistically large ow angles . The ow angle will never physically reach extremely large angles, as the vortex wake state will be entered, instead of the ever expanding wake diameter predicted by a simple non-recirculating mass ux analysis. For this reason and to aid convergence, has therefore been limited to 60 by modifying condition only ever found near the tip in any case.
r,

4.4.11 Expanding Wake


In reality, any induction must expand the ow through the rotor to satisfy continuity. Figure 4.8 demonstrates that the expansion is non-negligible for the common Betz limit operating point (a = 1/3) and standard assumed wake geometry. The vortex
Moves are limited to half that predicted by Eqs. (4.4.27) and (4.4.28) to facilitate convergence.
1

74

Chapter 4 Analytic Development

wake tube should strictly follow the tip streamline, since free vortices must strictly convect with the ow. To approximate this condition, alternate induced velocity equations were derived for a cone shaped vortex sheet (i.e. radius R a linear function of z in Eqn. Eq. (4.4.12)). The wake could then be approximated as a series of N cones with the mid-side of each cone (dots in gure) aligned with the local ow, as shown in Fig. 4.17. The coincident edges of the cones are positioned in a cosine distribution downstream, to put more points in the high-curvature near-eld region, joining a cylinder wake at 5Rtip . The strength of the sheets i are constant over each individual cone.

Figure 4.17 Expanded wake prole

The wake is initialized to have constant radius Ri = Rtip and i = cyl . After each
BEM solution, each wake segment is updated sequentially downstream by comput-

ing the induced velocity at the mid-point and modifying the slope of the segment to align it with the ow. After all segments are updated,
z

and

are recomputed

based on the deformed wake geometry and a new BEM solution is obtained. The convergence condition was a movement of the wake end-points between wake iterations of less than 0.001Rtip . Convergence was obtained after approximately 1020 wake iterations for the cases presented in 5.1. The wake sheet strength cyl of the far-wake cylinder is determined from mass continuity and vortex theory. The former is an integration of the mass ow in Eq. (4.4.3) and the latter the result that at the centre of an innitely long cylinder, the induced velocity is uniform across the cross-section and equal to the vortex sheet strength. Rearranging this relationship yields: cyl = V0 m/ 2 Rf ar (4.4.29)

4.4 Aerodynamic Modelling

75

The wake strength is not strictly constant (as assumed in the classic result wz,2 = 2wz ) for either the cylindrical or conical wake geometries, since it must convect with the ow. Note that is an area density of vorticity, varying only in the streamwise direction due to the bunching up of vortex laments as they slow down. The vorticity density must therefore increase, so that = (Vz ). Circumferentially, the vortex lines simply stretch as the local wake radius increases, changing the core size, but maintaining the same path integral (circulation) around them. The strength of each wake segment i is then simply: i = cyl Vz,f ar Vz,i (4.4.30)

The wake length used (2.5 D) is a commonly used value in wind farm wake models [120], as the point at which expansion is considered complete. Variation of the length had minimal eect on the results compared in 5.1 for the expanded (Ex) and unexpanded (Unex) wake geometries. Likewise, the results are shown for N = 10 wake segments, as increasing N had little eect on accuracy.

4.4.12 Additional Wake Geometries


Two additional wake geometries are developed for comparison. For any real blade, the loading along the blades varies; therefore, vorticity must be continually shed along the span. To crudely model this, additional stream-tubes joining points along the blade from downstream are considered. The circumferential direction of the vorticity of these sheets is opposite to the one shed at the tip to reect this directionality. The rst geometry is similar to that in [100], consisting of a single tube at some fractional radius towards the root. Here the tube is shed at 20% of the tip radius, approximately where the cu of the blade prole greatly reduces the lift on the blade (large change in vorticity). The second wake geometry consists of equal strength vorticity tubes joining the blade at 10 equally spaced radial stations, with the cumulative vorticity shed at the tip. In all cases, a straight cylindrical sheet is assumed. The correction factor
z

is

calculated from the vortex system at each radial position. It would be possible to include a second induction factor iteration loop, to adjust the vortex sheet strengths according to fed back changes in loading along the blade using the multiple vortex tube model. This would entail dramatically more computational eort and is not pursued here.

76

Chapter 4 Analytic Development

4.4.13 Aerodynamic Analytic Optimum


An analytic optimum blade shape (chord and twist distributions) can be derived from the BEM equations. To dene the optimum, the airfoil must be selected, in particular the AOA clcd,max for maximum lift-to-drag cl /cd ratio. An analytic optimum is only possible for xed values of and . For conventional variablespeed rotors operating at a xed , the optimum is well-dened by the analytic solution. For xed-speed and coning rotors, which sweep through and , the analytic optimum can only serve as an initial blade design. This is also true if the airfoil varies along the blade, producing non-smooth geometry. Likewise, for a real blade, structural and other considerations (circular blade root) will require further optimization to produce a smooth and realistic blade shape. This problem is treated in 6.4. To derive an analytic optimum, it is assumed that cd = 0 and F = F (, ) in Eq. (4.4.6a) for convenience. It will turn out that CT is below the critical point, so the equations equating momentum and element forces can be non-dimensionalized as:
2 BVrel ccl cos cos = [(1 a) cos a r sin ] 8F a z RV02 2 BVrel ccl sin = [(1 a) cos a r sin ] 8F a 2 RV02

(4.4.31a) (4.4.31b)

for axial and azimuthal momentum respectively, using = R/V0 and = r/R to follow Burton et al. [16]. Likewise with Eq. (4.4.6a): 2 B 1 3 F = cos1 exp 1+ 2 2 (4.4.32)

where it is assumed from later calculations that a = 1/3. Next, the two balance equations are divided and combined with the geometric formula for tan to relate a and a : ((1 a) a r tan ) a z (4.4.33) 2 2 (1 + a ) To nd maximal power output, the torque must be maximized, so the stationary a = point of the right-hand-side of Eq. (4.4.31b) is found from: C = 0 = a ( cos a
r

sin ) +

a [(1 a) cos a r sin ] a

(4.4.34)

Using Eq. (4.4.33) with (1 + a ) 1 in the above yields: = a= 1 (1 + 3


r

tan )

(4.4.35)

4.4 Aerodynamic Modelling which agrees with other analyses for = 0 [16].

77

Finally, a blade parameter cl may be found from the blade element half of Eq. (4.4.31b): cl = 4F a 2 2 ( (1 + a )) + ((1 a) cos a r sin )
2 2

(4.4.36)

The blade parameter may be thought of as a non-dimensional chord. Given a set of operating conditions (, ), section location , and best cl at clcd,max , Eq. (4.4.35), Eq. (4.4.33) and Eq. (4.4.36) dene the optimal blade shape with twist angle from: twist = clcd,max (4.4.37)

The new set of optimum equations include the correction factors of the revised BEM theory, indicating a rotor optimized for non-zero will dier from that optimized for a planar rotor.

4.4.14 Dynamic BEM


In reality, wind turbines never operate in a uniform inow. Wind shear, tower shadow, temporal variation, and skewed ow all alter the conditions at the blades and in the wake. To account for these eects, ve main elements must be considered: The induction eects of a skewed wake Proper projections of the incident velocities The 2D airfoil properties must include dynamic AOA changes The transport equations method of dealing with yaw Some accounting of the time-delay associated with temporal changes in induction These are treated sequentially in the following sections, to derive a unied BEM applicable to a generalized yawed, coned rotor in unsteady ow. A nal solution iteration scheme is outlined last. Note that the following derivation leads to dynamic equations that do not require azimuthal iteration [16] or assumed induction [100] for yawed rotors. The current method thus contrasts to previous treatments of the subject that are not able to compute a general unsteady solution. 4.4.14.1 Skewed Wake Analysis The wake may be skewed relative to the rotor shaft of the rotor by the aggregate eects of structural deection (yaw, tilt, etc.) and the free-stream wind velocity.

78

Chapter 4 Analytic Development

The wake vortices may be decomposed as before in Fig. 4.6, in this case assuming the plane of the ring elements remains parallel to the tip-plane plane of the rotor as they migrate downstream. Extending the analysis of 4.4.7 to include skew, it is found that the axial laments will induce both azimuthal and axial velocities that vary with skew and cone angles. For = 0 , the root lament is responsible [16], while for = 0 the bound vortices also contribute. Nevertheless, the axial laments are again ignored by setting

= 1, assuming again second order eects [100]. As the

blades sweep through a yawed ow, the AOA and hence lift and shed vortex strength will continually change with azimuth. The assumption adopted here, as before, that the constant strength tip-wake cylinder dominates, is therefore not strictly valid. The simplication from varying strength helical sheets to tip cylinder is however consistent with the previous analysis. The integral for the skewed wake induction is set up with reference to Fig. 4.18. The coordinate system is the shaft-aligned x y z axes from Fig. 4.1(d). Point p is the point of interest, located a vector r away from the integration point on the vortex cylinder. The wake is assumed to develop as a skewed elliptical cylinder of strength and constant radius R. The vortex rings lie in planes parallel to the y z plane with their common centreline at a skew angle from the x axis. The wake axis tracks the point xw , yw , zw , related to the tangent of the skew angles csy and csz projected in the x y and x z planes respectively: yw = csy xw zw = csz xw (4.4.38a) (4.4.38b)

The Biot-Savart integral of Eq. (4.4.10) can be evaluated by rst dening the following quantities: dl = rring id = R 0, cos , sin d r = p R 0, sin , cos + xw , yw , zw = px xw , py R sin yw , pz R cos zw = rx , r y , r z dl r = R rz cos + ry sin , rx sin , rx cos d
2 2 2 |r|3 = (rx + ry + rz + (c R)2 )3/2

(4.4.39a) (4.4.39b) (4.4.39c) (4.4.39d) (4.4.39e) (4.4.39f) (4.4.39g)

4.4 Aerodynamic Modelling

79

Figure 4.18 Skewed wake integral Equation (4.4.39g) adds the same cut-o model as used previously in Eq. (4.4.12). The nal integral that must be computed for the wake from xw = 0 . . . is:
2

u= 4
0 0

dl r dxw |r|3

(4.4.40)

When this integral is expanded, dxw (as a result of Eq. (4.4.38), yw and zw are a function of xw ) and d become the variables of integration. As before, the axial dxw portion of the integral may be analytically solved, while the d part of the integral is solved numerically. 4.4.14.2 Correction Factors The induced velocity u now has 3 components ux , uy , uz . Evaluating at points f (p) u and f (xw = 1000R) uf ar (at same radial distance from skew axis as p) yields the following correction factor relations:1
z uy uz
1

ux,f ar 2ux uy = ux uz = ux =

(4.4.41a) (4.4.41b) (4.4.41c)

Note that the z and x axes are swapped for the skewed wake formulation relative to the simple coned formulation in 4.4.5.2, to be consistent with the global CS of 4.2. The correction factor z denition is kept as before.

80

Chapter 4 Analytic Development

Note that the previous result that |uf ar | = is not correct for the skewed wake. For p = xp , r cos , r sin , the radial and azimuthal velocity components are: ur = uy sin + uz cos u = uy cos + uz sin The previous radial factor
r

(4.4.42a) (4.4.42b)

(see Eq. (4.4.13)) is then:


r

ur ux

(4.4.43)

A new azimuthal correction factor laments as for


):

(owing to the vortex rings, not the axial u ux

(4.4.44)

These correction factors are used later, as in 4.4.6, to relate the induction a at the section to the far-eld induction, and to the radially and tangentially induced velocity components. To reduce the on-line computational burden of evaluating the integrals at each time-step, a 4D table of axial, radial and azimuthal induced velocities was computed. The points are referenced by [tab , rtab , xtab , tab ], as shown in Fig. 4.19. Utilizing symmetry, the table need only cover tab = 0 . . . /2, rtab = 0 . . . 1 (R = 1), tab = /2 . . . /2 and xtab = tan max . . . tan max . The angle max is set to 60 , and is the maximal cone angle with Rhinge = 0 the table can completely cover (outside the axial range the edge value is used). The maximal value of tab is set to 80 as a practical limit, and rtab is bounded by 0.98 to avoid numerical diculties with the vortex singularity.1 The points are distributed using cosine spacing, to place more near the rotor tip axially and near /2 in the regions of maximum u gradient. The table is computed with the skew axis in the x y plane. In order to use the table, the coordinates of the actual points on the blade sections ps must be transformed into the table coordinate system. As shown in Fig. 4.20, the skew axis in general lies in a plane rotated = arctan csz /csy around the x (shaft) axis of Fig. 4.1(d). The skew axis is assumed to pass through the shaft axis a distance xw0 away from the y z plane, corresponding to the tip-plane.2 Using the azimuthal angle s and local radius rs , the table may be entered with the
An excessive numerical quadrature tolerance was required to accurately compute u past this radius. 2 The tip-plane for each blade is computed independently, varying with .
1

4.4 Aerodynamic Modelling

81

0.8 0.6

z
0.4 0.2 0 -0.5 0 0.5 0.5 0 -1 0 1

y
(a) yz grid points

x
(b) xr grid points

0.8

0.6

y
0.4 0.2 0 0

0.2

0.4

0.6

0.8

x
(c) Skew axis () range

Figure 4.19 Grids for following coordinates: xtab = s sin xw0 rtab = Rhinge + s cos tab = s tab = arctan

table

(4.4.45a) (4.4.45b) (4.4.45c)

c2 + c2 sz sy

(4.4.45d)

Note that is signed, allowing the table to only cover positive values. The angle tab is bounded to by adding/subtracting 2 as appropriate, and nally computed as: tab tab = tab tab /2 tab /2 tab > /2 tab < /2
z

(4.4.46)

Finally, the factors are computed from the interpolated induced velocities. The

is the only non-symmetric component, requiring a change of sign for |tab > /2|.

82

Chapter 4 Analytic Development

Figure 4.20 Skewed wake table transformation is practically bounded to 10. Storing u instead of
z

The axial factor

was preferand
2 ,

able for accurate interpolation. As ux 0, relatively coarse grid of numerical issues. 4.4.14.3 Wind Vectors

, making interpolation on a
1

relatively less accurate. The other factors,

introduced later in 4.4.14.6, are computed and stored directly as they avoid these

The aerodynamic velocity vectors for the fully dynamic set of equations must be carefully composed from their components, as only a subset are aected by the induction factors. The total velocity seen by the blade section is given by Eq. (4.4.47), including the induced velocities Vind , wind Vwind , rotational velocity Vrot = r, and structural velocity Vstr owing to the dynamic motion of the section. These vectors are all in the global earth CS of Fig. 4.1(a). Vtot = Vind + Vwind + Vrot + Vstr (4.4.47)

The induced velocities are left to 4.4.14.4. The wind velocity given in Eq. (4.4.48) is composed of the turbulent wind at the section location, Vturb (t, x, y, z), modied by the velocity increments owing to wind shear Vshear , tower shadow Vshadow ,

4.4 Aerodynamic Modelling and velocity decits or turbulence from up-stream turbines Vupwind
turb .

83

Vwind = Vturb (t, x, y, z) + Vshear + Vshadow + Vupwind

turb

(4.4.48)

The upwind turbine contribution Vupwind work.

turb

is not explored further in the current

The fully turbulent wind Vturb (t, x, y, z) is in general obtained from a probability function [121]. A fully 3D block of wind vectors is synthesized over the plane of the rotor, for a region of space extending in the streamwise direction a length equal to the mean wind speed times total simulation time. This wind eld then passes through the rotor over the run of the simulation, dening interpolated velocity vectors for any position in the eld corresponding to the coned and rotating blade sections (or tower sections). The presence of the rotor is assumed to have no eect on this volume of velocities. The wind shear increment Vshear is dened according to the standard exponential model of Eqs. (2.1.1) and (4.4.49). Only the components of wind in the xe ye plane of Fig. 4.1(a) (i.e. parallel to ground plane) are scaled, according to Vshear (x, y) = (1 fws )Vturb (t, x, y) with h = z in Eq. (4.4.49). V = Vref h href
ws

= fws

(4.4.49)

The tower shadow model accounts for the retardation (or acceleration) of the ow due to the presence of the tower. The model smoothly blends a potential ow model upstream of the tower [122], with a wake decit model downstream [123125]. To apply the model, the wind vector Vturb,tow (t, xtow , ytow , ztow ) at the centreline of the tower at time t and height z is found. Again, the model is only applied to the velocity components in a plane parallel to the ground plane. The local angle is therefore found from arctan(Vturb,tow,y /Vturb,tow,x ) with the wind magnitude Vw,tow =
2 2 Vturb,tow,y + Vturb,tow,x . The point of interest (i.e. x, y position of the

blade section) is translated to the (x y ) CS with R z (). Dropping the primes on the transformed coordinate notation, the velocity increment is calculated from: Vshadow = Vw,tow fv (u + v i j)

(4.4.50)

84

Chapter 4 Analytic Development

Figure 4.21 Tower shadow model

The velocity increments u and v are computed separately upwind uu , vu : uu =


2 Rt (x2 y 2 ) cd Rt x + 4 r 2 r2 2 xy R cd Rt y vu = 2 t 4 + r 2 r2 r = x2 + y 2

(4.4.51a) (4.4.51b) (4.4.51c)

and downwind ud , vd : cos2 y w x ud = 0 vd = 0 x x= 2lw Rt w = 2ww Rt x r d= 2 Rt |y| w/2 |y| > w/2 (4.4.52b) (4.4.52c) (4.4.52d) (4.4.52e)

(4.4.52a)

The upwind equations are the solution for inviscid ow around a cylinder, with a drag correction term. The downwind equations are empirical, based on a cosine curve distribution for the velocity decit. The wake decit is non-dimensionalized by the free-stream. The wake width ww and downstream distance lw dene the location and extent of the specied decit, both non-dimensionalized by the tower diameter 2Rt . An alternate formulation for ud that only depends on cd (with less investigative

4.4 Aerodynamic Modelling exibility) has been proposed [122]: c d cos2 d ud = 0 d= r2 2 Rt

85

y 2Rt d

|y| |y| >

d d (4.4.53b) (4.4.53a)

Although not used directly here, this formulation provides an estimate of ww and , based on cd . The upwind and downwind models are smoothly blended in the x y plane using a cosine function: x0 uu 0 < x Rt u = fa uu + (1 fa )ud ud x > Rt x0 vu 0 < x Rt v = fa vu 0 x > Rt fa = 1 2 cos x Rt +1

(4.4.54a)

(4.4.54b)

(4.4.54c)

An example complete ow eld is shown in Fig. 4.22, for typical parameters (see 5.2). The tower cylinder is superimposed on the eld. Note that the inclusion of a drag-dependent term in the inviscid upstream formulation produces a stagnation region upstream of the tower. The inuence of the tower on upwind rotors is typically assumed minimal, while it is recognized as critical to downwind rotors. Figure 4.22 shows that within Rt of the tower surface, the ow disturbance is of similar magnitude, indicating its signicance to all rotor orientations. It is also necessary to blend the tower wake model smoothly from the top of the tower.1 For points above the tower top, the blend can either be azimuthal or vertical; the latter was chosen and implemented as: 1 1 (z ztop ) fv = cos +1 lblend 2 0

z ztop ztop < z ztop + lblend z > ztop + lblend (4.4.55)

The blending length lblend is taken as the diameter of the tower at the tower top.
It was discovered that BLADEDTM does not do this blending which resulted in severe derivatives when performing noise calculations (see 4.5.3).
1

86
3

Chapter 4 Analytic Development

y
1 0 -3

-2

-1

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.1

Figure 4.22 Example total velocity tower shadow ow eld (1 + u)2 + v 2 (Rt = 1, cd = 1, = 0.3, ww = 2.5, lw = 3.25)

4.4.14.4 Wind Vector Decomposition Having dened the various velocity components, a number of components can be computed with reference to Fig. 4.23. The rst is the out-of-plane velocity Vw,OP = Vwind OP . This vector corresponds to V0 in the steady BEM formulation in Fig. 4.2. It is eectively the shaft-aligned free-stream velocity, normal to the unconed plane of the rotor.

Figure 4.23 Dynamic wind vector decomposition

The normal, tangential and spanwise sectional velocities Vw,n , Vw,t , Vw,s are computed as the dot product of Vwind with n, t and s respectively. These are the wind velocities that the section sees without any induction. The computation of the

4.4 Aerodynamic Modelling

87

structural velocity term Vstr is dened later in 4.6.3 and decomposed to normal and tangential components Vstr,n and Vstr,t respectively. The total velocity that the 2D section sees, Vrel is dened with reference to Fig. 4.24 and Eq. (4.4.56). The aerodynamic force vectors are the same as in Fig. 4.4, however the axes and velocity vectors are modied. The induced components Vind,n and Vind,t shown in Fig. 4.24 are components of the axial, radial and tangential induction velocities. These represent the proper induced velocities, acting on the out-of-plane VOP and in-plane r component velocities respectively. The distinction between out-of-plane/in-plane and normal/tangential velocities, and which velocities are induced constitutes one source of discrepancy between analysis codes.1

Vn = Vw,n aVw,OP (cos +

sin ) + Vstr,n
z

(4.4.56a) (4.4.56b) (4.4.56c)

V = Vw,t + r 1 + a + aVw,OP Vs = Vw,s aVw,OP (sin


r

+ Vstr,t

cos ) + Vstr,s

Figure 4.24 Dynamic ow relative to airfoil section (velocity vectors scaled for clarity, not indicative of relative magnitudes)

The nal element is the computation of , the generalized skew angle. The Vturb,tt (t, x, y, z) vector used for computing the wake skew vector is determined at
BLADEDTM computes induction based on total section velocity, whereas NRELs AeroDyn is similar to the present derivation acting only on the wind vectors.
1

88

Chapter 4 Analytic Development

the centreline of the tower-top, with wind shear but without tower shadow. It is assumed that the wake convects downstream in-line with this vector. In reality, the wake will self-induct and tend to increase the skew angle [16], especially near the rotor. Thought of another way, the thrust force (directed normal to the rotor) on the ow is directed upwind and in the direction of aerodynamic yaw yaw , thereby increasing the skew angle. Burton et al. [16] provides a formula approximating the proper skew angle corr derived from vortex theory: corr = (0.6a + 1) yaw (4.4.57)

where yaw is the geometric skew angle without induction. This formula assumes that the wake convects far downstream along the velocity vector at the centre of the disk, which is also not strictly true. It also implies a dependency on a requiring further solution iteration. For these reasons, no correction for wake self-induction is made to in the current formulation. 4.4.14.5 Unsteady Sectional Aerodynamics There are two basic approaches to modelling the 2D aerodynamics of an airfoil in dynamic situations. The simplest method is to adopt the quasi-steady model, which assumes that the airfoil behaves according to steady 2D lift curves. More advanced theories recognize that changes in lift and drag do not develop instantaneously. There is therefore some lag introduced in the tracking of the steady AOA-cl /cd /cm curves. Two DOF are important: pitching and heaving. Both dynamically alter the angle of attack and the latter the incident velocity as well. In order for the aerodynamic force system to change, the ow around the airfoil must change over a nite period of time. The eects are separated into lift-dependent (circulatory) and added-mass (non-circulatory) eects. The former is further dierentiated into an inner and outer problem, owing to vorticity on the airfoil and in the wake respectively [126]. Outside of the linear lift regime, dynamic stall behaviour has been observed experimentally. Governed by the rapidity of AOA variation, the airfoil can achieve much higher lift coecients (approaching the linear 2) well past AOAs for which stall would have occurred in a steady ow. This is possible because the boundary layer only changes relatively slowly, so that separation (leading to stall) takes time to develop. Eventually the airfoil does progress to full-stall, and does so in a much

4.4 Aerodynamic Modelling

89

more violent manner (and over a smaller AOA range) than in the steady case. Once the airfoil has fully stalled, a hysteresis eect is observed, wherein the fully stalled condition is maintained as the AOA is reduced. The dynamic airfoil behaviour only returns to the steady characteristic once the airfoil is well back into the linear regime. The nature of the stall (from LE or TE) is largely a function of thickness. A number of models have been proposed to account for the dynamic stall phenomenon [89, 126, 127]. The model adopted here was chosen for its relative simplicity and for comparison with Computational Fluid Dynamics (CFD) results [89] in 5.1. Accurate stall prediction is dicult even for much more advanced models [126], which also incorporate better dynamic eects below stall. The present method, presented by Mikkelsen [89], only considers changes in cl , and essentially maps the airfoil between two curves as the AOA changes. The upper curve is the linear one, dened by: cl,lin = 2( 0 ) and the lower curve is the fully-separated curve: cl,sep = Gcl,st A0 G = (1 A0 ) + [1 + tanh (B0 (| 0 | 1 ))] 2 (4.4.59a) (4.4.59b) (4.4.58)

where the suggested constant values are A0 = 0.3 . . . 0.5, B0 = 5 . . . 15 and 1 is the rst stall angle of the steady lift curve data dened by cl,st (). The zero lift angle of the airfoil is 0 . The tanh function smoothly blends the assumed fully-separated cl,sep curve1 to the cl,st curve (G 1 away from = 1 ). The three separate curves are given in Fig. 4.25 for an example airfoil, with the parameters A0 and B0 varied for cl,sep . The mid values for the t constants (0.3 and 10) are used in the remainder of the current work. The movement between the curves is determined by a fractional function fst (): fst () = cl,st () cl,sep () cl,lin () cl,sep () (4.4.60)

Clearly, this equation denes the steady cl,st curve for fst = 0.5. For each time step i, the value of fst () is computed with Eq. (4.4.60). A simple lag equation is then
i used to compute the fraction fst to use for the current step: i1 i1 i fst = fst + fst () fst
1

t dyn stall

(4.4.61)

This curve is a rough approximation, with approximately half the cl,st curve slope near 0 and joining the cl,st curve around 30 .

90
1.5

Chapter 4 Analytic Development

cl

0.5

c c

l,lin l,st l,sep,mid l,sep,low l,sep,high

c c

-0.5

c -30 -20 -10 0 10 20

-40

30

40

(deg)

Figure 4.25 Dynamic stall parameters (low to high values for A0 and B0 are [0.3, 0.4, 0.5] and [5, 10, 15] respectively)

where t is the time since the last computation, and dyn stall the lag time step. The suggested value of 4c/Vrel is used, changing locally with chord c but using a xed average Vrel of 30 m/s.1 Finally, the current lift coecient is computed by interpolation:
i i cl () = fst cl,lin () + (1 fst )cl,sep ()

(4.4.62)

To test the implementation, an example hysteresis loop was run, is shown in Fig. 4.26. The model displays a smooth behaviour and predicts a lift increment into stall, as desired. 4.4.14.6 Governing Equations Implicit in the following discussion is the continued validity of the stream tube independence assumption and to varying degrees azimuthal averaged eects. For the yawed rotor, these assumptions are even less rigorous than the steady, un-yawed case. They are rst adopted pragmatically and the results compared to other predictions and measurement in 5.1 to ascertain the valdity of the assumptions. There are three possible departure points for deriving the momentum equations, based again on considering averaged conditions over an idealized uniform disc [16]. The rst, termed Axial Momentum Theory (AMT), assumes the thrust (pressure) force of the disc creates average induction a at the disc and 2a in the far-eld, both
1

It was found that varying Vrel with operating condition led to numerical diculties.

4.4 Aerodynamic Modelling


1.8 1.7 1.6 1.5 1.4 c c

91

l,lin l,st

cl,sep c
l

cl

1.3 1.2 1.1 1 0.9 0.8 4 6 8 10 12 14 16 18 20 22

(deg)

Figure 4.26 Dynamic stall hysteresis loop for cyclic ( = [5 . . . 20 ] at 1 Hz)

perpendicular to the disc. The axial momentum equation is formulated in the same perpendicular direction, ignoring unbalanced pressure forces that also contribute to the wake skew. Glauert proposed the second theory, by comparing the autogiro downwash system to that of a highly yawed wind turbine rotor. The key dierence to AMT is that the transport velocity through the disc area is the total velocity at the disc, not that perpendicular to the disc. This allows a component of the thrust force to create an eective lift force on the rotor, analogous to a wing or autogiro rotor at an AOA. Burton et al. [16] proposed a third model based on averaged vortex theory. The induced velocities at rotor centre and downstream are used in the Bernoulli equation to derive the thrust equation for the disc. It is observed that AMT will predict power most accurately (lowest CP ()), and Glauert the thrust (highest CP ()) [16]. Notice that Glauert attributes some lift force to thrust, and as this is perpendicular to the velocity, cannot extract work from the ow. The vortex theory produces results intermediate to these theories for CP . BLADEDTM and most other codes uses AMT, and so this theory is employed here with those caveats.

92

Chapter 4 Analytic Development

The axial (shaft-aligned, x y z CS of Fig. 4.1) momentum equation is derived as before in Eq. (4.4.2), with V0 replaced by Vw,OP : (momentum) = V2 Vw,OP m = (2F a z Vw,OP )m
A

(4.4.63a) (4.4.63b) (4.4.63c)

m =
0

V1 d A

The denition of dA varies with the AMT or Glauert assumptions: dAAM T = dA cos + sin sin cos k i j dAGlauert = dA V1 V1 The transport velocity at the section V1 includes the section wind velocity Vwind = Vw,x , Vw,y , Vw,z and the induced velocities. Adopting AMT, the mass-ow term becomes, with dA = dsrd and from Fig. 4.1(d):
s 2

(4.4.64a) (4.4.64b)

m = Vw,OP
0 0

cos (1 a) + Vw,y + Vw,OP sin Vw,z + Vw,OP

sin

uy a

uz a

cos rdsd (4.4.65)

The tilde over a and

are used to emphasize that these quantities vary during


za

integration with . Since the induction in the wake af ar = 2F a z Vw,OP is constant over , the local a and a may be related as =
z a.

Now

= f (), whereas

( z , a) = f () in the velocity change term of Eq. (4.4.63a). For a temporally and spatially uniform inow (assumed approximately valid in dynamic ow as well), Vw,y and Vw,z will also be invariant in . Performing the integrals while noting the components that evaluate to zero over = 0 . . . 2 yields: m = Vw,OP rs2 [cos (1 a where
2 1 z 1 )

2 sin ]

(4.4.66)

1 = 2
0 2

1
z

d
2 uz z

(4.4.67a)

1 = 2
0

cos

uy z

1 sin d = 2
0

r z

(4.4.67b)

4.4 Aerodynamic Modelling

93

These factors are evaluated with trapezoidal integration from the table quantities stored in 4.4.14.2. For = 0, these equations reduce to the un-yawed ones in 4.4.10. As before in 4.4.3.4, empirical CT,model values are used for adisc = a
z

> ac to

introduce an H factor into the momentum equation for CT,mom , non-dimensionalized


2 by 1/2Vw,OP 2rs cos . The blade element CT,loc equation remains unchanged

as: CT,loc =

2 V2 Bc cz Vrel = cn 2rel 2 2r cos Vw,OP Vw,OP

(4.4.68)

It must be assumed that CT,loc applies to the whole annulus, even though it will vary with as the conditions at the blade change (i.e. Vrel ). A steady code could iterate on this to achieve uniform a
z

over by using the averaged element forces,

however a dynamic code does not have this luxury. Temporally varying inow can be handled in ve ways [16, 89, 125, 128, 129], in increasing order of complexity: Frozen wake This method assumes that the wake is frozen after the rst time step. The induced velocities aVw,OP and a r do not change thereafter. Equilibrium wake This method assumes that the wake changes instantaneously with loading changes. The solution is computed anew from the steady equations, at every time step. Dynamic inow The most realistic assumption is that the wake will evolve with some time-lag relative to loading changes. Vortex laments The wake is evolved as time-dependent vortex strengths and paths. CFD The full-eld simulation of the Euler or NS equations simply include temporal eects when implemented in their unsteady forms. Notice that it is the induced velocities aVw,OP , not induction factors a, that change. The dynamic inow is preferred here, for comparison to BLADEDTM . The solutions may either be formulated generally by including multiple assumed functions for the pressure distribution on the disc, or in a reduced form containing terms for aggregate thrust and moment. The former uses acceleration potential methods based on asymptotic solutions of the Euler equations to model the ow. Wake rotation is not explicitly included in the theory, modications are required for the high a levels of wind turbines, and coning cannot be accounted for [122, 130]. BLADEDTM adopts the latter, most primary and simple assumption, that only thrust is aected.

94

Chapter 4 Analytic Development

Essentially, the forces applied by the blades on the ow must not only balance the steady momentum change of the ow, but also apply a force to temporally change the induction velocity of the ow. This is handled by an additional term in the momentum equation: CT = CT,mom + CT,dyn where CT,dyn mA z,f ar dt = 2 1/2Vw,OP dA
z d(w /2)

(4.4.69)

(4.4.70)

The far-eld induction term wz,f ar /2 = aVw,OP far wake induction that changes. BLADEDTM

is used to emphasize that it is the

and standard theories can only use

the value at the blade aVw,OP , as no information is available about the change in induction from relative wake position with . Neglecting this subtlety would tend to over-predict the dynamic inow changes. From potential theory, the appropriate added mass term mA for a disc of radius R is 8/3R3 [16, 125]. Therefore, for an annulus from r1 to r2 : CT,dyn = k d (aVw,OP z ) dt 3 3 16 r2 r1 k= 2 2 3 r2 r1 (4.4.71a) (4.4.71b)

BLADEDTM simply sets k = Rtip .1 Each blade is computed independently (i.e. dierent , a solutions, etc.) assuming the equations valid for the entire annulus. To initialize the dynamic inow equations, and for the frozen and equilibrium wake cases, the axial momentum equation is rearranged for xed-point iteration as:

anew = f (a) =

c1 1 ac2 V2 H + a2 c1 = cn 2rel Vw,OP 4F z c2 =


z 1

(4.4.72a)
2 z

tan

(4.4.72b) (4.4.72c)

The geometric relations used to derive Eq. (4.4.26c) are no longer applicable, however this turns out to be fortuitous. In xed-point iteration, it the iterate equation with minimal f (a) that is the fastest [131]. The choice of iterate equation is constrained by the condition that |f (a)| < 1 to converge. Equation (4.4.72) violates
1

The dierence was found to be minimal, and Eq. (4.4.71b) is used here.

4.4 Aerodynamic Modelling

95

this near a = 1/c2 where f (a) . Solutions above this limit are not viable in any case, so the updates are limited as: anew = 1/c2 anew c2 a 1 < or anew > 1/c2 anew 1/c2 (4.4.73)

The oset is set to 0.01 to avoid the discontinuity. It was evident from the decreased solution iterations required that f (a) for this formulation is smaller than that of Eq. (4.4.27). The tangential equation uses the dynamic mass-ow term of Eq. (4.4.66) in Eq. (4.4.4) to yield: anew = c 4F
rVw,OP 2 Vrel (cos (1 a z 1 )

2 z

sin )

(4.4.74)

Including the added mass term in CT for the dynamic inow model, the axial iterate equation is: d (aVw,OP ) 1 = dt K z
2 cn Vrel

H 2 4 z Vw,OP a (1 ac2 ) a2 F

z 2 tan

(4.4.75)

A simple Euler scheme is used to advance the solution and obtain the xed-point iteration equation for the dynamic wake: anew = (aVw,OP )old,1 + d (aVw,OP ) 1 (t told,1 ) Vw,OP dt (4.4.76)

The estimate of anew is bounded as in Eq. (4.4.73). New solutions are only computed when t told,1 exceeds a set dynamic wake time step, typically 0.02 s for a large machine [125]. The new value of aVw,OP is then stored as (aVw,OP )old,1 , after updating the previous value of (aVw,OP )old,2 (aVw,OP )old,1 . The induction factor is nally linearly interpolated based on t between the values at told,1 and told,2 , so that the wake evolves smoothly over the next wake time step, thereby introducing aerodynamic lag. The tangential update equation is unchanged from Eq. (4.4.74), but is lagged from its dependency on Eq. (4.4.75). 4.4.14.7 Computational Procedure For clarity, the overall ordering of computations in the general case is given in the following procedural list. All steps are carried out for each section on each blade in turn, with coupling to the structural model of 4.6.3. 1. Structural positions and velocities are obtained either deterministically, or computed as part of the time-marching solution of structural equations (see 4.6.3).

96

Chapter 4 Analytic Development 2. The wind vectors, as described in 4.4.14.3 and 4.4.14.4 are computed based on the position and orientation of the blade sections. 3. Induction factors a and a are initialized according to the unsteady induction model: Initial Timestep/New Dynamic Solution a = 0.3 and a = 0 Frozen Previous values Equilibrium Default values Dynamic If (t told,1 ) dyn , a new dynamic solution is computed. If (t told,1 ) < dyn , the induction values are interpolated: t told,1 fdyn = (4.4.77a) told,1 told,2 (aV0 )old,2 + fdyn ((aV0 )old,1 (aV0 )old,2 ) (4.4.77b) a= Vw,OP (a r)old,2 + fdyn ((a r)old,1 (a r)old,2 ) a = (4.4.77c) r The induced velocities (aVw,OP )old,1 and (a r)old,1 are the most recent estimates computed at time told,1 . (aVw,OP )old,2 and (a r)old,2 are the previous estimates computed at time told,2 . The interpolation function fdyn therefore allows the new induction to fully develop over the time span dyn , introducing the aerodynamic lag. 4. The full dynamic ow velocities are computed from the relations in 4.4.14.4 and the induction factors. 5. The aerodynamic forces are computed, including the eects of stall delay (see 4.4.8 and 4.4.9) and dynamic stall (see 4.4.14.5). The tip-loss factor F and turbulent wake H parameters are also computed. 6. Induction factors are updated if required. This is done in all cases for the rst time-step and for the equilibrium wake model using Eq. (4.4.72). The dynamic wake model uses Eq. (4.4.75) for all subsequent time-steps when new inductions factors must be computed ((t told,1 ) dyn ). Steps 46 are repeated to convergence of the induction factors. 7. The various aerodynamic values are recorded and loads computed (see 4.6.2). The dynamic stall fractions fst are updated, as are the dynamic wake induced velocities (aVw,OP )old,i and (a r)old,i .

4.5 Acoustic Modelling


The noise created by wind turbines can be separated into two main areas: aerodynamic and structural. Structural noise can arise from the following sources [132]:

4.5 Acoustic Modelling Transmission (gear meshing, bearings) Coolant systems Electrical components

97

Structural noise on modern wind turbines has largely been eliminated by careful design of components, acoustic treatment of the nacelle, facilitated by relatively well understood physical mechanisms. The focus of the noise analysis presented currently is therefore on aerodynamically generated noise. Section 4.5.1 rst introduces the metrics and mechanisms of noise. Next, 4.5.2 overviews the modelling approaches and mathematical background of noise prediction. Section 4.5.3 then introduces
LFN, the primary mechanism of concern for the coning rotor. Fortuitously, theory is

available in the literature to model LFN from rst principles, presented in 4.5.3.2. The nuances of building the theory into an accurate prediction code are however not present in the literature in a unied manner, and so are discussed last in 4.5.3.4 and 4.5.4.

4.5.1 Aerodynamic Noise


At a fundamental level, aerodynamic sounds (noise if an undesired sound) are acoustic waves travelling through the air. Physically, sound is equivalently a pressure or density (condensations and rarefactions) wave travelling at a characteristic speed c0 , depending on the ambient uid temperature and composition (343 m/s at 20 C and 1 atm for air). Humans experience sound as pressure uctuations acting on the ear drum. The frequency of the wave is experienced as tone, and the amplitude as volume level. Since the ear is a dynamic mechanical system, it has a dened frequency/amplitude response that inuences the perception of sound.

4.5.1.1 Intensity To account for the human response to noise, Sound Pressure Level (SPL) Lp is measured on a logarithmic scale: Lp = 10 log10 p2 prms rms = 20 log10 2 pref pref (4.5.1)

where pref is a standardized pressure of 20 P a corresponding to the limits of human hearing. The units of Lp are decibels (dB). The sound power level LP and intensity

98 LI may also be dened:

Chapter 4 Analytic Development

LP = 10 log10 LI = 10 log10

P Pref I Iref

(4.5.2a) (4.5.2b)

where Pref is 1012 W and Iref is 1012 Wm2 . The power is a constant function of the acoustic source, while the pressure and intensity both decrease with distance r from the source. In free-space, I is proportional to p2 ; LI and Lp will decay as rms r2 as the sound is spread over a spherical surface centred at the source. 4.5.1.2 Frequency Spectrum At the most basic level of characterization, SPL may be computed by sampling the entire frequency spectrum. To account for the frequency response of the human ear, three weighted sounds levels (A, B, C) are used. Each denes a lter function over the range of human hearing (20 Hz20 kHz) to attenuate sound at the high and low ends of this range. Commonly referred to as dBA sound levels, this denotes a sampling with the A lter suitable for most situations. The B and C weightings are more appropriate for low-frequency sounds [132]. Two methods are used to resolve specic frequency content of sound, octave and Fast Fourier Transform (FFT) analysis. Octave analysis divides the frequency spectrum into bins dened by: fc,i+1 = 21/n fc,i fl,i = 2(1/2n)fc,i fu,i = 2(1/2n)fc,i (4.5.3a) (4.5.3b) (4.5.3c)

where fc , fl and fu are centre, lower and upper frequencies for each bin i. By international standard, 125 Hz is a centre frequency (the others may be derived from the above equations). Most analysis is done for n equal to 1 or 3, termed 1/1 and 1/3 octaves respectively. In practice, sound meters use multiple analogue lters tuned to these frequency ranges to sample the sound and obtain Lp for each bin. To obtain ner frequency resolution, FFT analysis can be performed on data discretely sampled from the continuous pressure signal (see 4.5.4). It is possible to convert from FFT results to octave bins Lp,bin by summing the FFT Lp,j in each bin:
b

Lp,bin = 10 log10
j=a

10Lp,j /10

(4.5.4)

4.5 Acoustic Modelling

99

where a and b are the lower and upper indices of the FFT results in that bin. Note that is it the sound intensity I that must be summed, not the Lp values. The above equation is only valid for incoherent (dierent frequency and phase) noise sources.

4.5.1.3 Noise Mechanisms The minimisation of aerodynamic noise from wind turbines remains a challenge, but by careful design can be minimised. The rst step is to identify the individual aerodynamic noise mechanisms [133]:
TE - Turbulent Boundary Layer The turbulent boundary owing over

the trailing edge of an airfoil will radiate noise from both suction and pressure surface TEs. Separated Flow Progressive stall from the TE increases the boundary layer thickness and size of coherent eddies passing over the TE. In full stall, noise is radiated from the entire suction surface as as result of these large eddies and can dominate the TE source.
TE - Laminar Boundary Layer This tonal noise is created by amplica-

tion of Tollmien-Schlicting waves via vortices shed at the TE, into coherent vortex shedding. The Reynolds number must be small enough for laminar ow to develop and so this source is more important for small machines. Blunt TE Depending on geometry, especially for thick and blunt TEs, coherent vortices are shed creating tonal noise. The phenomenon is analogous to the von Karman vortex street behind blu bodies. Turbulent Inow The eddies present in the turbulent inow incident on the rotor interact with the LE of the airfoils to radiate noise. The magnitude of the noise is larger when the ratio of turbulence length scale to LE radius is large, the former ranging in size from 1 mm to 100 m. Tip Vortex Unlike all of the other noise sources which are predominantly 2D phenomena, this origins of this source are 3D. The interaction between the shed tip-vortex and the blade surfaces create an important highfrequency noise source. The strength of the vortex is determined by the shape of the spanwise load distribution and can be mitigated from this perspective. Also, it has been found that parabolic tips, and others with cut-away TEs reduce tip noise. The former tailors the spanwise loading, while the latter also moves the tip surfaces upwind of the tip-vortex spiral.

100

Chapter 4 Analytic Development

Low-Frequency Noise (LFN) Described further in 4.5.3, this source is typically at or below the threshold of human hearing, but nevertheless is important as it is felt. LFN is created by the blades sweeping out a volume of air and also uctuating sectional aerodynamic forces acting on the uid and moving relative to the observer. The primary source of LFN is from the blade passing through the tower wake creating rapidly uctuating forces. The noise heard by an observer will be modied by two factors, directionality and propagation. Directionality modies the sound perceived by the observer owing to the relative position (directivity of the sound) and motion of the sound source, the time-lag due to propagation from spatially distributed sources, and convection of the noise with the free-stream. The propagation of the sound from source to observer will also introduce artefacts owing to attenuation, reection o terrain and buildings, and the non-uniform properties of the atmosphere. All but the last of these noise sources can be alleviated by clean airfoils, sharp trailing edges and tip-shape selection. Operationally, stalled rotors and those with high tip-speeds produce relatively more noise. The latter is a design choice, typically limited to 70 m/s on-shore. O-shore however, where noise is less of a constraint, tip-speeds may increase substantially in the future, leading to cost-reduction through reduction of torque/forces and hence material usage. LFN will be discussed further in 4.5.3.

4.5.2 Modelling Approaches


Three classes of model may be used to model aerodynamic noise [132]. The rst Class I models are empirically based on measurements from existing machines. They typically are parametrised in terms of generic quantities such as rotor diameter, rated power and tip speed to give an overall dBA sound level. They can give an overall indication of noise production, but not detailed spectral information or accurate quantitative predictions. Class II models are the current state-of-the-art, such as those in the NREL FAST code [134], and rely on more detailed information about the machine. They consider the noise mechanisms separately to synthesize the overall acoustic signature. These models are based on fundamental theory, either in terms of deterministic forces or scaling laws from non-dimensional 2D airfoil analyses. They are also supported and validated by experimental work [135]. Both Class II and III models are capable of

4.5 Acoustic Modelling

101

providing time-domain predictions of acoustic pressure, and hence frequency content and overall noise levels. Class III models rely on detailed description of the geometry within the context of a full-eld time-domain simulation. These approaches typically solve the
NS equations in either compressible or incompressible form (see [136] for a dis-

cussion of the relative merits). The predictions must be accurate for separated ow, so Reynolds-Averaged Navier-Stokes (RANS) (steady solution) and Unsteady Reynolds-Averaged Navier-Stokes (URANS) (excessive numerical dissipation) are inadequate; Large Eddy Simulation (LES), Detached Eddy Simulation (DES) or Direct Numerical Simulation (DNS) simulations are required. Clearly the underlying NS solution is quite expensive [137], and due to cost is not used to any great extent in practice. This can be reduced somewhat by employing the acoustic analogy discussed next. 4.5.2.1 Acoustic Analogies Mathematically, acoustic sound is governed by the NS equations. Together with the continuity equation, these equations fully describe the generation and prorogation of acoustic signals, i.e. aerodynamic noise. Unfortunately, the full set of equations requires ne spatial and temporal resolution for accurate predictions. This precludes the use of ne grids extending from a wind turbine to observer. To overcome this limitation, Lighthill developed his acoustic analogy [138]. Without delving into the mathematical details, it is essentially a re-arrangement and combination of the continuity and momentum equation resulting in a single compact equation: 2 c2 0 t2 where Tij , the Lighthill stress tensor is: Tij = ui uj + (p c2 )ij + ij 0 (4.5.6)
2

2 Tij xi xj

(4.5.5)

where ij is the Kronecker delta function, xi and ui are position and velocity components, and ij are the viscous stresses (usually ignored as second order). This equation is a hyperbolic partial dierential describing a wave travelling at speed c0 in a quiescent medium (left-half), forced by source terms (right-half). This inhomogeneous wave equation (acoustics) is the analogy to the full uid-dynamics equations. Equation (4.5.5) can be solved for the small disturbance density (x, t) = (x, t) 0 , and hence pressure via the isentropic relation p = a2 at the observer position 0

102

Chapter 4 Analytic Development

x. The nal solution involves an integral of the stress tensor through a volume bounded by a xed surface enclosing the noise generating regions. It also includes an integral of pressure over the same bounding surface. Using this mathematical result, the process of computing noise is separated into two distinct tasks. First, a detailed computation is made of the ow conditions within a volume bounding the important noise generating mechanisms. Using these results, the far-eld noise signature is then obtained by integrating the detailed solution, as just described, at each time-step. The ow-eld conditions are therefore know a priori for the second step. Lighthill originally developed the theory for jet noise, using a stationary boundary. Ffowcs Williams-Hawkings extended the analogy to include moving boundary surfaces, introducing two additional terms into the solution [136]:
2

p (x, t) =

{[0 vn + (un vn )] (f )} t 2 {[Pij nj + ui (un vn )] (f )} + [Tij H(f )] (4.5.7) xi xi xj

where un and vn are the uid and boundary surface velocities normal to the boundary surface f = 0, (f ) is the Dirac delta function, and Pij is the compressive stress tensor jump across the surface. The addition of a moving boundary allows the bounding surface to be a real (on the blade surface) or an imaginary surface located in the ow eld but conveniently co-rotating with the blades. 4.5.2.2 Mathematical and Physical Noise Sources The three source terms in Eq. (4.5.7) are mathematically monopole, dipole and quadrupole elementary acoustic sources [132]. The abstract and physical nature of each source, and examples for each (in brackets) are as follows: Monopole A point source with radial symmetry (radially pulsating sphere surface); uctuating sources of mass (siren); moving volumes (propeller blades) Dipole Fluctuating and/or translating forces exerted on uid (oscillating sphere); varying lift (airfoil angle of attack change); moving lift force (rotating blade) Quadrupole A pair of dipoles (small deforming sphere); turbulent ows, uctuating Reynolds stresses (eddies from stalled ow over wing)

4.5 Acoustic Modelling 4.5.2.3 Advanced Analysis

103

Although the Ffowcs Williams-Hawkings (FW-H) equations are conveniently compact, they assume sound radiation into an unbounded domain. Wind turbines are typically placed in a complex environment. To perform a detailed simulation of atmospheric absorption and terrain eects, the linearised Euler equations may be used with clever computation of the Greens function [136]. However, this level of complexity is beyond the scope of the current work.

4.5.3 Low-Frequency Noise


This section presents the relevance (4.5.3.1), theory (4.5.3.2 and 4.5.3.3), and implementation nuances (4.5.3.4) for analysing the LFN of the coning rotor.

4.5.3.1 Relevance A noise study was commissioned during the course of the original work on the coning rotor concept [139]. That study only included two noise mechanisms, turbulent inow and TE - turbulent boundary layer interaction. Mention was also made of using cut-away TEs tips to limit tip-noise and sharp trailing edges (< 3 mm) to avoid vortex shedding noise. The overall ndings were that the coning rotor concept did not create noise levels much in excess of a conventional machine. The modest increase that was observed, derived mainly from the increased TE length of the relatively longer blades. The variable-speed coning rotor presumably also better matched noise production to background levels compared to the constant speed reference machine. Missing from this earlier study was a treatment of LFN. While this noise source does not have a large component of A-weighted noise levels, it is very important for public acceptance. This is based on practical experience in both the US with the MOD-1 [15, 140] and in Sweden with the WECS Maglarp machine [141]. Both were large ( 80 m rotors) prototype downwind machines with 2 bladed rotors. The MOD-1 had a heavy-weight truss-type tower with 4 major structural elements creating a downstream wake with 24 distinct velocity decits (depending on yaw) and a broad wake. The rotor was also completely rigid. The WECS machine had a more conventional cylindrical tower. Both experimental eorts generated complaints from neighbours due to the characteristic tower thump (LFN described in 4.5.1) heard (and felt) at the blade passing frequency.

104

Chapter 4 Analytic Development

Clearly then, LFN is an important aspect of the coning rotor design, as the rotor must be downwind of the tower. Owing to the fact that LFN can travel long distances before being attenuated, it will be important for both on-shore and near to mediumoshore concepts. LFN is not necessarily a fundamental constraint on the concept, but must be carefully considered if the coning rotor is to be publicly acceptable. Some comfort may be taken from other downwind concepts that have been built and did not suer unduly from LFN: Htters machines of the 1950/60s; the MS-4 u machine in the UK; and the Wind Turbine Companys prototypes.

4.5.3.2 Theory Utilizing a number of simplifying assumptions, LFN can be treated relatively easily using a development of Eq. (4.5.7). This development is outlined by Wagner et al. [132, p. 98], citing detailed work by Succi and Farassat and is described here. For wind turbines operating at low Mach numbers, the driving mechanisms of LFN are monopole and dipole sources (see 4.5.2.2). The monopole source is created by the moving volume of the blades as they move through the air. This can be visualized as the blade temporarily displacing a volume of air in a given region as the blade sweeps through it, and it termed the thickness noise. The dipole source can be divided into two distinct mechanisms. Both are created by the aerodynamic forces (lift and drag) developed by the blade, vectorially opposite to those experienced by the blade, as they are exerted on the uid. The rst source is created by a constant magnitude force moving relative to the observer and is termed near-eld noise. The second source is developed by a temporally varying force vector (magnitude and direction) and termed far-eld noise. The solutions to Eq. (4.5.7) typically contain inconvenient spatial derivatives that may be re-cast as time derivatives. They also contain integrals over a bounding volume and surface. The volume integral may be ignored, as no quadrupole sources (turbulence) are important for LFN. The surface integrals may be conveniently considered as acoustically compact for frequencies below 50 Hz. Using this assumption, the monopole and dipole sources may be evaluated as discrete sources on the blade, rather than requiring integration over the actual blade surface. This approach is commensurate with the BEM method (see 4.4), which provides position and loading output for sections along the blade. It is assumed here that the forces are located on the pitch axes of the blades.

4.5 Acoustic Modelling

105

The nal formulas governing the time dependent pressures created at observer location x produced by the three sources are as follows: pthk,i (, yi ) = pnear,i (, yi ) = pf ar,i (, yi ) = 1 V 0 4 r(1 Mr ) r2 (1 1 Mr )2 1 1 Mr 1 1 Mr (4.5.8a) (4.5.8b) (4.5.8c)

1 4

ri fi ri

1 Mi M i fi Mi 1 Mr ri Mi

1 4

1 c0 r(1 Mr )2

fi ri fi + 1 Mr

These equations are applied at each spanwise element of each blade, using the timedependent aerodynamic force vector fi in units of force per unit-length of blade. Likewise, the sectional volume V0 , thickness pthk,i (, yi ), near pnear,i (, yi ), and far pf ar,i (, yi ) pressures are per unit span (denoted by the overbar). The position of each blade section is yi ( ). Note that all input quantities are known a priori from the BEM simulation for all i = 1 . . . ns stations on ib = 1 . . . B blades. The other parameters in Eq. (4.5.8) are the unit vector from source to observer ri and the distance between source and observer ri : ri = |x yi | x yi ri = ri the source Mach vector and its derivative: 1 yi c0 Mi 1 2 yi = c0 2 Mi = and the relative Mach number: M r = r i Mi (4.5.11) (4.5.10a) (4.5.10b) (4.5.9a) (4.5.9b)

For simplicity, the sectional volume is computed with the following formula: t V0 = tpc + (1 p)c 2 3 (4.5.12)

where t is the section thickness, c the chord length and p 0 . . . 1 a parameter determining the fractional chord wise location of maximum thickness (taken as 0.35 throughout this work). This formula approximates the airfoil as a half-ellipse for the rst pc length of blade, and a triangle for the remaining length.

106

Chapter 4 Analytic Development

Returning to Eq. (4.5.8), the three components are rst evaluated for each section in retarded time (the original simulation time corresponding to the time when the noise is generated). The observer time ti (time at which the pressure disturbance reaches the observer) may then be computed for each segment from: ti ( ) = + ri ( ) c0 (4.5.13)

Implicit in this equation is a disregarding of the free-stream velocity, which will modify the transmission velocity from source to observer. For wind turbines, this velocity dierential is less than 0.1c0 and is typically ignored. The pressures in retarded time psrc,i ( ) are therefore converted to observer pressures psrc,i (t). The next step is to properly assemble the pressure signals from each source and location i. This is done by sorting each pressure psrc,i (t) according to arrival time at the observer. Practically, this is accomplished in the implementation used for this thesis by linearly interpolating each psrc,i (t) value to a common tobs time computed as: t0 = max [min [ti (0)]] obs tn+1 = tn + dtobs t min [max [ti (max )]] (4.5.14) (4.5.15) (4.5.16)

The choice of dtobs will become important for proper FFT analysis (see 4.5.4). The penultimate step is to integrate over the span of the blade(s) to obtain physical pressures psrc,j (t, yj ) from pressure per unit span psrc,i (t, yi ): psrc,j (t, yj ) = (sj+1 sj )(src,j+1 psrc,j ) j 1 . . . ns 1 p (4.5.17)

These pressure sources therefore eectively originate at the mid-point of the spanwise segments located at yj ( ). The nal step to compute the overall pressure timehistory at x is to sum the contributions from each source and panel:
B ns 1

p(x, t) =
ib =1 j=1

pthk,j (t, yj ) + pnear,j (t, yj ) + pf ar,j (t, yj )

(4.5.18)

4.5.3.3 Simplified Theory It should be noted that an even more simplied approach is possible [132]. First, the acoustically compact approach is carried one step further, by assuming that the forces on the blades can be lumped at one spanwise point (e.g. 7275% radius). The

4.5 Acoustic Modelling

107

loading is then assumed to move in a circle, any dynamic blade motion is ignored, and the rotor is assumed to rotate at constant speed. Using these additional assumptions, an analytic formula for the nth harmonic of RMS pressure at the observer location may be derived. The formula depends only on the Fourier coecients of overall rotor thrust and torque (derived over one rotor rotation period). It has been applied in the study of noise from rotating forces (Gutin propeller noise) and wind turbines [141]. The current work uses the more accurate and complete formula of Eq. (4.5.8). 4.5.3.4 Nuances Equation (4.5.8) is relatively compact and straight-forward to implement, but care must be taken with the input data. Specically, as the equations are dependent on rst and second order derivatives of position, loading, and derived quantities, the numerical procedure for obtaining these derivatives becomes important. The timestep used in the original simulation must also be suciently small to accurately determine the derivatives, as well as for issues associated with post-processing (see 4.5.4). Ordinarily, derivatives are avoided if possible in data-gathering. Data is gathered for derivative values directly (e.g. acceleration) and then integrated to yield lower-order quantities (e.g. velocity and position). This procedure avoids the noise amplication associated with computing numerical derivatives of data.1 When only zeroth degree data is available, as is the case for Eq. (4.5.8), various nite-dierence formulas with varying degree f (n) , location of the centre point, number of included points, and error O(h) may be derived from either interpolating polynomials or Taylor series expansion. Typical formulae for the rst and second derivatives of a function f (x) with error terms are [142]: f (xi+1 ) f (xi1 ) 2h f (xi1 ) 2f (xi ) + f (xi+1 ) f (2) (xi ) = h2 f (1) (xi ) = 1 + h2 f (3) () 6 1 h2 f (4) () 12 (4.5.19a) (4.5.19b)

Unfortunately, sampled data invariably contains some noise component, either physical artefacts of sampling a real system, or numerical noise in simulations. Moreover, data samples only possesses a certain level of accuracy. Consequently, these formulas may depend on very small dierences between large numbers, that may in fact be noise rather than signal. This leads to cancellation and rounding errors and in
1

It does of course introduce the new problem of integrator drift error.

108

Chapter 4 Analytic Development

turn to highly inaccurate derivative estimates. In application to Eq. (4.5.8a), the problem is even worse owing to a double-use of rst derivative formulas. One approach to this problem is to construct a moving average estimate of the derivative [143]. Using linear regression, the sampled data f (x) = y(x) may be assumed to follow a linear relationship y = mx + b over a limited interval. For derivative computation, only the slope m is important, yielding the following formula for the derivative: mi = n( xi yi ) ( xi ) ( yi ) n x2 ( xi )2 i (4.5.20)

where mi is the derivative estimate at data point (xi , yi ). The summations are carried out over n data points around xi . In fact, Eq. (4.5.20) is a type of low-pass digital lter, more generally dened by:
nR

fi =
n=nL

cn f (xi+n )

(4.5.21)

where fi is the smoothed estimate, the cn a set of weights, applied to a set of points nL and nR to the left and right of the current data point xi [144, Chap. 14.8]. In the case of simple data-smoothing, the most basic constant coecient would be cn = 1/(nL + nR + 1) (moving window averaging). For constant or linearly varying data, this would produce an exact result, but for higher order variation would be in error. Mathematically, depending on the weights, the various moments1 of the data are preserved or not. To analyse chemical spectra, Savitzky and Golay [145] developed a set of lters referred to as Savitzky-Golay Smoothing Filters (SGSF)2 that preserve the higher moments up to some degree. To build the lter, a polynomial function of a given order is t to the data, in a least-squares optimal sense. The parameters for the t include: the total number of points to use (ns = nL + nR + 1); centring of the other data points relative to the current one (i.e. nL = nR potentially for data near start and end of sampling); and the degree of the polynomial n. Since a least-squares t is a linear operation, the t procedure only has to be done once and thereafter applied to a moving set of data as a digital lter. The mathematical details of the lter derivation are left to the interested reader. The implementation of the lter weights used in the current work was obtained
Moment of order e e dened as e = i xe f (xi ). The zeroth moment is the area, 1st i moment location in x, etc. 2 Also referred to as a least-squares smoothing and dierentiating lter.
1

4.5 Acoustic Modelling

109

as the Matlab function sgsdf_gram_poly.1 The code came pre-validated against other data [146], implemented in an ecient recursive algorithm and containing parameters for polynomial degree, derivative degree, and number and centring of data points. The polynomial interpolation approach readily yields a set of weights that directly lead to the derivatives of the sampled data. The order of the derivatives s must of course be less than or equal to n and a sucient number of data points used. The choice of of n and ns eects the characteristics of the smoothing. More points and higher polynomial degree better preserve data peaks, but at the expense of decreased noise rejection. Based on guidance given by Press et al. [144] and parametric variation, the SGSF lters used to evaluate the acoustic equations use n = 3 and ns = 20. One-sided lters are used for the rst and last ns data points.

4.5.4 Post-Processing
The nal output of the analysis procedure outlined above is a pressure-time history at an observer position. In order to evaluate the noise characteristics, the pressure signal p(x, t) is transferred to the frequency domain using Fast Fourier Transform (FFT) analysis. This topic is covered in the literature [147, 148], but elucidation of the nuances involved can be hard to nd [149]. In order to obtain accurate quantitative results, there are a number of aspects to consider: proper units, sample rate, windowing, and binning. 4.5.4.1 FFT Units The Fast Fourier Transform Xk is the discretely-sampled (and computationally efcient) analogue of the continuous-time Fourier transform X(j), dened here by the following equations:
+

X(j) =
N

x(t)ejt dt

(4.5.22a)

Xk =
i=1

xi e

2j (i1)(k1) N

(4.5.22b)

where j is here

1, k = 1 . . . N . The xi are the N sampled points, uniform spaced

in time at dt = 1/fs , corresponding to a sampling frequency of fs . Essentially, the


Source code from Jianwen Luo, PhD Candidate, Department of Biomedical Engineering, Tsinghua University, Beijing 100084, P. R. China, e-mail: luojw@bme.tsinghua.edu.cn, luojw@ieee.org
1

110

Chapter 4 Analytic Development

FFT computes the frequency components of a signal, returning the amplitude and

phase of each component as a complex quantity.1 The theory derives from the Fourier series, whereby any periodic signal may be decomposed into an innite summation of sines and cosines of increasing frequency, plus a constant term. The Fourier transform (integral) extends this to non-periodic signals. The Discrete Fourier Transform (DFT) replaces the continuous integrals with discrete summations. The FFT is simply an ecient way of computing the
DFT. In order for the FFT algorithm to work, N must be a power of two N = 2n

(typically 256, 1024). The FFT may be thought of as frequency bins, with each Xk acting as a lter of bandwidth f . The achievable frequency resolution and range both depend on the original sample frequency and number of sample points according to: f = fmax fs N fs = f 2 (4.5.23a) (4.5.23b)

Matlabs fft function was used here, which implements the above equations and returns a two-sided spectrum2 of complex values. To convert from the computed spectrum values to amplitude Ak for the positive frequency components, the following relation is used: |X1 | k=1 N Ak, peak = (4.5.24) |Xk | 2 k = 2 . . . N/2 N Since the FFT amplitudes represent sinusoidal components of the signal, the RMS values of the (non-zero frequency) components may then be found from: Ak,
RMS

Ak, peak 2

k>1

(4.5.25)

The power spectrum is dened as: S k = A2 k,


RMS

(4.5.26)

This allows denition of the signal frequency components on the Decibel scale as: dB = 10 log10 Ak, RMS Sk = 20 log10 Sref Aref (4.5.27)

Additionally, phase information is available from the complex vectors Xk .


1 Note that some canned functions may return the power spectrum, not amplitude spectrum. 2 Both frequencies are returned, zero frequency component rst followed by positive then negative frequency components.

4.5 Acoustic Modelling 4.5.4.2 Sample Rate and Aliasing

111

The sample rate fs is important not only for determining the spectral range in Eq. (4.5.23), but also in accurate decomposition of the signal. It turns out [148] that a digitally sampled signal, where fs < 2f , will suer from an eect called aliasing. The quantity f is the Nyquist frequency, and denes the permissible lower bound on the sampling frequency, given a maximum frequency content f in the sampled signal. As an explanation, the digital sampling is not occurring frequently enough over the period of the signal 1/f . The peaks and troughs of the signal are sampled at points spaced too far apart which then appear as an entirely dierent frequency waveform. This in eect creates a false signal at an alias frequency fa = (f mfs ) if f fs /2, for m = 1, 2, . . .. In physical measurements, this problem is avoided by employing a high sample frequency. If the analog signal cannot be assured to have frequency components below fs /2, anti-aliasing lters are used before the analog-to-digital converter to clean the signal. This approach is not possible for numerical simulated systems, such as the current LFN computations. There are essentially three types of numerical simulation that may be encountered in the present context: Partial Dierential Equations (PDEs) Equations1 solved using a eld method (e.g. nite dierence/element) to yield solutions in position and time. The temporal discretization method must satisfy von Neumann stability criteria, and the spatial discretization the Courant limit. An example is the solution of the NS equations. Ordinary Dierential Equations (ODEs) Solved to yield positional and force quantities over time. The numerical integration schemes (e.g. RungeKutta) must satisfy the von Neumann stability criteria. Examples include the usual coupled dynamic equations for a BEM-structural simulation of a exible wind turbine, or prescribed airfoil section (or blade) motion to solve dynamic aerodynamic equations (e.g dynamic inow, dynamic stall). In the latter case, the time-step may be discretionary (i.e. stability easily satised) but still requires a suciently small time-step to attain reasonable accuracy.2
1 2

Hyperbolic or parabolic in this context. E.g. simple rst-order Euler equation used for dynamic inow equations has poor accuracy due to local truncation error.

112

Chapter 4 Analytic Development

Prescribed Time A deterministic system of equations with known analytic solution for all t is sampled at discrete time intervals. In each of the cases, satisfaction of the stability criteria, or sucient prescribed temporal resolution, will ensure that the computational sample frequency fs is sufciently fast to capture the frequencies for the relevant phenomenon. The Nyquist criteria will therefore be satised, as no higher frequency component can exist in the signal, as it is not present in the numerical simulation. 4.5.4.3 Windowing Implicit in the FFT algorithm is that the signal is sampled for an integral number of periods, and that the signal is innitely repeated in time. Unless the sampling is triggered to be synchronous with the process (enabling capture of an integral number of cycles), or the signal decays to zero over the sample time, this condition will never be true. This sampling-introduced artefact is termed spectral leakage. In order to mitigate these eects, the sampled data is windowed [150]. A window is a digital lter that modies the tail regions of the data sample, gradually reducing them to zero. This avoids sampling non-integral numbers of periods. The denitions and characteristics of a number of common windowing functions are given in Table 4.1, where fi = (i 1)/(N 1) is the fractional position of each data point i in the sample. Table 4.1 Windowing functions Window Uniform Hann Hamming Point Weight 1
1cos 2fi 2

Coherent Gain 1.0 0.50 0.54

Noise Power Bandwidth 1.00 1.50 1.36

Worst-Case Amplitude (dB) 3.92 1.42 1.75

0.54 0.46 cos 2fi

It is actually impossible not to use a window; without specically applying one, the Uniform window is applied. To understand the artefacts introduced by windowing, it is useful to consider the frequency spectrum of the window. Every window has a peak centre lobe (0 dB) with side lobes of decreasing magnitude (towards dB). The centre lobes are located at the frequency components of the true signal content. The application of the window to the signal can be viewed as a convolution of the signal frequencies with the frequency characteristic of the window. As an example, a signal with a single frequency, sampled for integral number of periods, will produce

4.5 Acoustic Modelling

113

an exact FFT result. If sampled for a non-integral number of cycles, the centre lobe of the window is shifted away from the true signal content. The side lobes therefore appear in the FFT results. Viewed another way, the FFT process views the convolved spectrum through a set of slits at the frequencies dened by Eq. (4.5.23); this is called the picket fence eect. The width and magnitudes of the lobes determine the spreading of frequency content between adjacent Xk , important for resolving closely spaced frequency content. However, the minimum width of the centre lobe is limited by increased spectral leakage as the energy content of the side lobes increases with decreasing main lobe width. The Hann window is used here, as it has good frequency resolution with minimal spectral leakage. Once a window is applied, it introduces two eects. The rst is to reduce the overall amplitudes of the signal, obvious from the applying the point weights of Table 4.1 (all < 1). To account for this, the nal Xk are divided by the coherent gain of the window. The second eect is that the windows spread the frequency content, eectively increasing the bandwidth f of each FFT bin. The factor increase is given in Table 4.1 as the noise power bandwidth. For estimating the true peak frequency for a component of a non-continuous spectrum,1 a frequency-weighted average of power over 3 adjacent peaks can be used. The true peak amplitude is then the sum of the adjacent powers, divided by the noise power bandwidth. 4.5.4.4 Ensemble Averaging To improve accuracy, multiple FFTs may be performed on a given signal, if the signals length is long enough (N < Ns ). In these cases, sections of signal, possible overlapping from 0100%, are ensemble average to yield a better estimate of the Xk . The last sample may be padded with zeros to ll a complete set of N samples. Of course, since the starting point of each sample is dierent, and in general not contemporaneous with an integral number of periods, the phase information embodied in the Xk is lost. 4.5.4.5 Binning Deriving octave binned frequency spectra from FFT results is accomplished with Eq. (4.5.4). Doing so provides a more coarse, but concise and standardized, representation of the noise. The total overall noise level may also me computed by the
1

This procedure is not valid for continuous spectra

114

Chapter 4 Analytic Development

same formula, either for the entire spectrum or a specic sub-set. The steady X0 component is disregarded throughout these calculations.

4.6 Structural Modelling


A wind turbine is an inherently aeroelastic structure, especially for large scale multiMW machines. There is therefore a need for an accurate model of the structure for combination with the aerodynamic models discussed in 4.4. In this section, sequential aspects of the structural modelling are developed: a general beam-section model (4.6.1); distributed load integration (4.6.2); a rigid-body model of an independently hinged rotor (4.6.3); and nally a FEM method for the blade modes(4.6.4). Only the EOM for the rigid rotor are derived here, leaving exible body dynamics to the commercial code (BLADEDTM ) with proper modes from the FEM method. Commensurate with the scope of this thesis, the primary emphasis is on the modelling of the rotor (blades). It is recognized that the overall dynamics of the complete machine, including drive-train (shaft(s), transmission and generator), yaw mechanism and tower are all critically important for a nal complete analysis in future studies.

4.6.1 Sectional Modelling


Wind turbine blades (and towers) are typically treated as linear beams. This is standard practice, even though it is arguable in some cases whether blade crosssections are well approximated as slender beams [151]. The primary quantities of interest for beam theory are the mass/unit length mps and area moments of inertia I, which must be computed for sections made up of many composite layers. The approach adopted here is similar to the one the author has used in the past [152]. The denition of the composite layup is dierent however, obtained here from direct user-entry rather than being CAD-based. The section properties are used in later derivations as inputs to describe the dynamic loads and motion of the blades. 4.6.1.1 Layup Definition For each airfoil dened for a blade, two input sheets are required in ExcelBEM. The rst contains the aerodynamic properties, and the other the airfoil proles. These proles are dened at various percent thicknesses covering those used in the blade for (linear) interpolation. Internally, the proles are interpolated when loaded to yield proles with a uniform number and location of chordwise spacing, to speed

4.6 Structural Modelling

115

computation. The same chordwise cosine spacing is used for the upper and lower surfaces, which are input separately to accommodate a range of data sources. The layup of blade sections is dened on a separate sheet. Every station may be dened, or only a subset which is kept constant for intermediate sections. The input sheet for the denition of an example single section is shown in Fig. 4.27, and the resulting section shown in the left-half of Fig. 4.28.
1 Upper Surface c (%) 0.00 t (m) 0.0005 0.0004 0.0009 0.0050 0.0009 0.0005 0.0004 0.0009 0.0100 0.0009 0.0005 0.0004 0.0009 0.0100 0.0009 0.00051 0.00038 0.00089 0.0005 0.0004 0.002 0.0009 t (m) Mat'l Gelcoat Random mat CDB340 Balsa CDB340 Gelcoat Random mat CDB340 Spar cap mix CDB340 Gelcoat Random mat CDB340 Balsa CDB340 Gelcoat Random mat CDB340 Gelcoat Random mat TE spline CDB340 Mat'l Gelcoat Random mat CDB340 Balsa CDB340 Gelcoat Random mat CDB340 Spar cap mix CDB340 Gelcoat Random mat CDB340 Balsa CDB340 Gelcoat Random mat CDB340 Mat'l CDB340 Balsa CDB340 CDB340 Balsa CDB340

0.18

0.53

0.85 0.99

Lower Surface c (%)

0.00

0.12

0.47

0.85

0.0005 0.0004 0.00089 0.005 0.00089 0.00051 0.0004 0.00089 0.01 0.0009 0.00051 0.0004 0.00089 0.01 0.0009 0.0005 0.0004 0.0009

Web(s)

c (%) 0.18:0.12 0.53:0.47

t (m)

0.0015 0.0100 0.0015 0.0015 0.0100 0.0015

Figure 4.27 Section denition

The denition is split into three, one for each of the upper and lower surfaces, and a third to dene the webs. For the upper and lower surfaces, fractional chordwise section divisions are input, starting with 0. The denition is assumed to continue to the TE, at c(%) = 1, so a minimum of one chordwise region must be dened

116

Chapter 4 Analytic Development

for both skins. These dene the locations of lay-up transitions along the skins. In this example, there are 5 lay-up denitions for the LE, spar cap, balsa-core TE, skinned TE and nally a TE spline. The last two columns dene the lay-up for each section, dened by layer thickness and material. The material denitions are named and dened on a separate sheet (density, moduli, cost, etc.). The web section may be blank, or contain an unlimited number of shear web denitions. To allow for angled webs, the chordwise locations may be dened as colon separted pairs, dening the fraction chordwise locations of the web intersection with the upper and lower surfaces respectively. The layup dention for the web(s) is identical to the other surfaces. It is assumed that the thickness of the web projects towards the TE. 4.6.1.2 Section Computations Using the layup of the sections structure, together with the airfoil proles, the relevant quantities may be computed as follows: 1. The relevant airfoil proles are interpolated and scaled to the proper chord length. For thicknesses outside the dened range, simple scaling of thickness is used.1 2. The normals ni shown in Fig. 4.28 are computed for each point on the airfoil surfaces using: ni = ni,upper ni,lower 1 yi1 + yi+1 2 xi+1 xi1 ni = |ni | ni = |ni | (4.6.1a) (4.6.1b) (4.6.1c)

where x, y are the airfoil prole coordinates. 3. Points and normals are linearly interpolated to place points exactly on chordwise layup transition points. 4. The upper and lower surfaces are processed by dening each layup thickness block as two triangles, and processing each triangle. The derivation of section properties for the triangular elements is included in Appendix B. The computation proceeds along each interval of the surfaces, and through the layup thickness along the normals. 5. The web is processed in a similar manner, using normals dened by the x, y coordinates of the web termination points on the airfoil surfaces.
More properly, the camberline should be extracted from the thickness prole, the thickness prole scaled and then recombined with the camberline [52].
1

4.6 Structural Modelling

117

The formulas for overall section properties (xCG , mps , EI, etc.) from the triangular element aggregatrations are also included in Appendix B.

Figure 4.28 Blade section triangulation

4.6.1.3 Common Confusions and Assumptions It is important at this point to clarify two issues, as they are commonly misconstrued. The rst is that the torsional constant GJe is not in general the same as the shear modulus weighted polar moment of area, computed here as GJ about xEA . They are equal only for axisymmetric circular bars [151]. The computation of the eective polar moment of area Je is more involved than computing J, utilizing either thin shell theory (shear ow) or more generally Prantls membrane analogy. These two quantities are however often used interchangeably, but may be in signicant (nonconservative) error, especially near the root [153]. Related to this is the denition of the shear centre, which is not in general co-located with the Neutral Axis (NA) (xEA ), although it is commonly assumed so. Likewise, the centre-of-mass (xCG ) and
NA are not in general co-located (the circle and cross in Fig. 4.28 respectively).

In any case, the pure torsional modes of blades are not typically used in wind turbine analysis. Notwithstanding a wealth of research [61, 154], there has never been any evidence of utter1 in practice. Flutter can be avoided in any case by ensuring the section xCG lies between the shear centre and the LE. Deliberate aptwist coupling has also been investigated, and so provision for input of these coupling values is provided in commercial codes (e.g. BLADEDTM ). The investigations are usually performed in terms of either an arbitrary coupling coecient [61], or derived from more detailed FEM analysis [155]. Another confusion surrounds the denition of apwise and edgewise/chordwise, the relation to Out-of-Plane (OP) and In-Plane (IP) and structural twist [80]. This confusion also extends into aerodynamic modelling. To clarify the denitions, a
1

Flutter is caused by the coupling of bending and torsional modes

118

Chapter 4 Analytic Development

general blade section with various coordinate systems is shown in Fig. 4.29. The section is viewed looking in inboard along the blade (i.e. towards z direction). The blade is translating to the right of the page, and the wind vector is nominally up the page, modied by the cosine of the cone angle.

Figure 4.29 Denition of section coordinate systems

At the LE is the x, y CS used for structural computations.1 The next coordinate system is on the chordline, at the star symbol c1/4 from the LE. This is the point where aerodynamic coecients (cl , cd , cm ) are dened. The next coordinate system is lcx away from the LE on the chordline. The inclined set of vectors dene the apwise and edgewise bending moment directions. The horizontal-vertical axis located pb (pre-bend) o the chordline denes the pitch axis location. All other analysis (aerodynamic and structural) compute forces decomposed into the (xaero , yaero ) directions.2 These are the OP and IP directions, for = 0 .

4.6.1.4 Stress/Strain Computation Wind turbine blades are usually governed structurally by direct normal stresses, rather than shear forces; the latter are therefore ignored in the stress calculations. The location of maximum stress (or strain) is usually checked at a number of points around the periphery of the section. In general, given combined loading, the curved
1 Standard airfoil coordinates take the LE as 1, 0 and TE as 0, 0 , so these must be reversed in x here. 2 The aerodynamic axes usually fall on the chordline (pb = 0). The pre-bend is included here to be comparable to BLADEDTM load results.

4.6 Structural Modelling

119

geometry and variable layup material properties (both stiness and strength allowables), the location of extreme stress is not known a priori. Therefore, the current analysis checks all grid-points computed in 4.6.1.2, using the following formula: yp (Mx EIy + My EIxy ) xp (Mx EIxy + My EIx ) Fz + 2 EA EIx EIy EIxy

(4.6.2)

The strain formulation is used to accommodate heterogeneous stinesses of the layers. The stress is of course found from = E using the layers Youngs modulus E, although composites are frequently characterised in terms of strain rather than stress. The loads (Fz , Mx and My ) must be translated from their denition on pitch axis (from other calculations) to xEA . The test points xp , yp are dened in the x, y LE coordinate system of Fig. 4.29 as vectors from xEA to the test point in the layup. Fatigue calculations (e.g. rainow counting) may proceed from the resulting stress or stain time-history, in the case of dynamic loading. Although fatigue tends to dominate many wind turbine components, static loads are used throughout this thesis to provide initial structural designs for later dynamic simulations of fatigue loading. Skin buckling could also be incorporated in the current framework, and should also be checked in a nal design [153, 156].

4.6.2 Kinetostatics
The key elements of aerodynamic force prediction and the mass properties of the blades have been developed in 4.4 and 4.6.1.2 respectively. The current task is twofold: to compute the moments required for equilibrium operation (see 4.7.1) and power production, and secondly to compute the loading in the blade for dynamic operation (4.6.3) and as input for structural load calculations (4.6.1.4). The loads are computed on the pitch axis (zaero in Fig. 4.1(d)), for each section on the blade lying in the xaero yaero plane. Three sources of loading are considered, aerodynamic (faero , maero ), gravitational (fgrav , mgrav ) and inertial (facc , macc ), comprising the total forces f and moments m per unit length. The loading is developed with reference to Fig. 4.30, which is consistent with Fig. 4.29. DAlemberts method is used to handle the inertial accelerations.

120

Chapter 4 Analytic Development

Figure 4.30 Section layout for kinetostatics

Figure 4.31 Kinetostatic oset vectors The lengths in Fig. 4.31 dene the position of P , which represents either the aerodynamic (da ) or mass centre (dm ). da = c (pA 1/4) dm = c (pA CG ) lx = pb d sin tot ly = d cos tot where lx and ly apply for either the aerodynamic or mass centre with subscript a or m respectively. The applied forces will then generate moments according to: mof f = (lx + ly f i j) (4.6.4) (4.6.3)

from this oset. The sectional mass is assumed concentrated at the CG , so no direct moments are produced from gravitational and inertial forces, only the mof f moments. Point masses are included as increments to the dened mass per unit length on the blade (mps ), computed to preserve the 1st moment about the ap hinge, by distributing the mass over two adjacent sections. The sectional forces and

4.6 Structural Modelling

121

moments (per unit span) are now outlined in 4.6.2.1 through 4.6.2.3, followed by the integrated quantities in 4.6.2.4. 4.6.2.1 Aerodynamic Loading The aerodynamic forces follow directly from the aerodynamic characteristics presented in 4.4.4. It only remains to decompose the vectors appropriately and include the eect of the oset section in the moment calculations. The 2D aerodynamic applied forces are:
2 q = 1/2Vrel c

(4.6.5a) (4.6.5b) (4.6.5c) (4.6.5d)

fx,aero = q (cl cos + cd sin ) fy,aero = q (cl sin cd cos ) faero = fx,aero + fy,aero i j

As discussed in 4.4.9, any shear force normal to the section, owing to the viscosity of the spanwise ow, is considered insignicant as a direct force. The pure moment contribution is found from: maero = (qccm ) k which will be in addition to the moment found from Eq. (4.6.4). 4.6.2.2 Gravitational Loading The gravitational force on the section is easily found from fgrav,e = mps g k. In order to compute the loading in the section CS, the unit vectors uaero,i describing the axes xaero yaero zaero are rst found from: uaero,i = R z (yaw) R y (t ) R x () R y ()ui (4.6.7) (4.6.6)

where ui are the unit vectors k, in the column notation of Appendix A. The i, j, section loading along each axis is then: fgrav,i = fgrav,e uaero,i (4.6.8)

4.6.2.3 Inertial Loading The inertial loading is determined by the acceleration vector from facc = mps a. The determination of the acceleration vector a is found from the kinetics of the

122

Chapter 4 Analytic Development

active DOF discussed in 4.6.3, the azimuth and cone angle . The position vector x of the section (on the pitch axis) relative to the centre of the hub (i.e. in x y z
CS of Fig. 4.1(b)) is found from expanding the following:

lx,m ly,m x = R x () T (0, 0, Rhinge ) R y () s 1

(4.6.9)

where s = shub + s. The full form of this equation, and the derivatives x and x are not included here for brevity. A set of reduced and easily veriable equations is obtained with lx = ly = Rhinge = 0: x=s j cos + sin sin cos cos i sin cos cos sin k (4.6.10a)

x=s

cos 2 sin i i+

cos cos + 2 sin cos + 2 cos sin + 2 sin cos + sin sin j+ sin cos 2 cos cos + 2 sin sin 2 cos cos cos sin k (4.6.10b) Equation (4.6.10a) contains the velocity terms owing to azimuthal rotation and apping, decomposed onto the coordinate axes. Equation (4.6.10b) contains the direct acceleration motions in azimuth and ap, the Coriolis accelerations associated with the rotating reference frames (), and the centrifugal acceleration terms (2 and 2 ), again decomposed through and . 4.6.2.4 Integrated Loading The loading from each source may now be integrated to yield shear loads and bending moments along the blade, for each section. Each contribution is computed separately, as the dynamic model in 4.6.3 requires only the integrated aerodynamic hinge moments. The properties of the blade are assumed to vary linearly between sections, as are the sectional forces. Figure 4.32 illustrates the blade as a beam, where section i is of interest, located si from the blade root and extending by s = (si+1 si ) outboard to section si+1 . The sectional applied forces f and moments m per unit length in 4.6.2.1, 4.6.2.2 and 4.6.2.3 are assumed to vary linearly with s. Both shear and moment loading are integrated from tip (constraint and force free) to hub.

4.6 Structural Modelling

123

Figure 4.32 Integrated loading

The orthogonal axial and shear forces Vi is easily obtained from integration over the section: Vi = Vi+1 + fi + fi+1 s 2 (4.6.11)

The moment components of Mi are also found by integration and from outboard shear forces as: Mi,x = Mi+1,x + Mi,y = Mi+1,y + Mi,z = Mi+1,z + mi+1,x + mi,x Vi+1,y s 2 mi+1,y + mi,y + Vi+1,x s + 2 mi+1,z + mi,z s 2 fi,y fi+1,y + 6 3 fi,x fi+1,x + 6 3 s2 s2 (4.6.12a) (4.6.12b) (4.6.12c)

4.6.2.5 Transformed Loading The nal step is to transform the shear and bending load from the most inboard section to the ap hinge and hub centre. The important hinge loads are found from:

Mx,hinge = M1,x shub V1,y My,hinge = M1,y + shub V1,x Mz,hinge = M1,z

(4.6.13a) (4.6.13b) (4.6.13c)

My,hinge is the critical coning moment, either summed over all sources, or just the aerodynamic forces (for 4.6.3). The hub loads in the x y z CS of Fig. 4.1(c)

124 are:

Chapter 4 Analytic Development

Mx,hub = Mx,hinge cos + Mz,hinge sin Rhinge V1,y Vx,hub = V1,x cos + V1,z sin Summing the hub loads for each blade, the driving torque is = P = and the rotor thrust T = apwise and edgewise directions: Vf lap = Vx cos tot Vy sin tot Vedge = Vx sin tot Vy cos tot Mf lap = Mx sin tot + My cos tot Medge = Mx cos tot My sin tot
B B

(4.6.14a) (4.6.14b) Mx,hub , power

Vx,hub .

The moments and shears for each section can also be easily transformed into the

(4.6.15a) (4.6.15b) (4.6.15c) (4.6.15d)

4.6.3 Dynamic Modelling


Wind turbines are described by multiple DOF, both rigid and exible (blades and tower). The EOM may be derived by numerous methods, all deriving fundamentally from a Newtonian force or energy approach (Hamiltons principle). Newtonian methods tend to be more laborious, as they require explicit vectorial consideration of each loading mechanism. Energy-based methods permit more general DOF and a scalar formulation to realize more compact EOM. To reduce the DOF to be included at run-time, modal representations of the exible structures are typically used. An alternative is an FEM model, either as beam elements or lumped mass/force elds. Both serve to reduce the partial dierential equations to ordinary dierential equations amenable to solution, the modal method being preferred for compactness. The modal approaches use either true normal modes, or approximate modes consistent with imposed boundary conditions and with reasonably accurate mode shapes (e.g. 4.6.4) [69]. In the former case, the PDEs are directly converted to ODEs by substitution of the modes (usually up to second or third modes). In the latter case, a Galerkin method may be used to reduce the PDE from a Newtonian formulation. Alternatively, Lagranges equation (Eq. (4.6.16)) or Kanes method may be used to directly synthesize the EOM in terms of assumed modes. The EOM may be further reduced by substituting the exible component mass and stiness (including rotational stiness) terms with modal shape and frequency

4.6 Structural Modelling

125

terms.1 This is only correct for normal mode shapes (i.e. no coupling of modes). Even then, the rotation speed must be constant and the stiness linear (it will be non-linear for large continuous bending or apping deections). Some authors adopt this simplication, assuming the last non-linear considerations small, but as will be seen in 5.4.3, this can lead to large errors. In most cases, apwise/edgewise mode shapes are used, so that with twist inplane/out-of-plane motion is coupled but the modes themselves are normal and invariant with pitch angle (to rst order). Others, for example BLADEDTM , use in-plane and out-of-plane assumed mode shapes, which vary with pitch angle.

4.6.3.1 Equations of Motion A full and detailed accounting of the all DOF in Fig. 4.1 is quite involved, especially when exible elements are incorporated. Numerous couplings are introduced by the various rotating and translating reference frames, yielding obtuse sets of equations. The post-processing of loads (see 4.6.2) further burdens the process of writing a full simulation code. BLADEDTM was therefore targeted as an eventual platform for full, exible body simulations. Unfortunately, commercial time constraints have hindered implementation of the aerodynamic renements appropriate to the coning rotor. Moreover, as illuminated in 5.4.3, it appears the structural model internal to BLADEDTM will require modication as well. In short, a full simulation has had to be relegated to future work (see 8.3). Here, only a rigid body set of EOM is developed, with individual blade coning j and rotor azimuth as the DOF.2 This will allow examination of the fundamental apping motion of the coning rotor. It is also the same number of DOF adopted by other authors studying apping blades [77]. Lagranges equations are used to dene the EOM [69]. The method followed is similar to that in the original CONE-450 report [66], but including more discussion of non-linearities, gravity, tilt angle, and independent apping. The total kinetic T and potential U energy of the system are obtained from kinematic analysis of the various DOF xi . The generalized forces on the system are
1 The terms of the homogeneous PDEs used to nd the modes are also present in the full non-homogeneous PDEs (see [69, p. 387]). 2 Note that the blade-dependent azimuth angle j is oset from the rotor azimuth angle by (j 1)2/B.

126

Chapter 4 Analytic Development

contained in Fi (forces for linear DOF and moments for rotational DOF). L=T U t L xi L = Fi xi (4.6.16a) (4.6.16b)

The resulting equations are equivalent to those obtained by kinetostatic analysis (Newtons Law) or Kanes equations. Lagranges approach was chosen as it is quite straightforward to apply and generates reasonable compact equations, although any internal forces must be obtained by post-processing. Applying the method to the reduced set of DOF used for dynamic analysis, the following expressions are obtained for T and U : 1 T = IH 2 + 2
B B j=1

1 2 1 2 Irj j + I j 2 2

(4.6.17a)

U=
j=1

Mb ghj

(4.6.17b)

The blade moment of inertia about the hinge axis I , blade mass Mb , and centre of mass distance sCG are assumed constant and equal for all blades (no imbalance). It is also implicit in the dynamics formula that the masses are on the pitch axis (pb = 0 and pA = CG in Fig. 4.30). IH is the inertial moment of the hub about the rotation axis. The moment of inertia about the centre of mass ICG of an individual blade yields I = ICG + Mb s2 . The moment of inertia of the blades about the CG shaft varies with cone angle j as:
S

Irj =
s0

2 r2 dm = Rh Mb + 2Rh Mb sCG cos j + I cos2 j

(4.6.18)

The potential energy term U needs a height hj , dened relative to the hub centre: hj = sCG ( sin(t ) sin j + cos(t ) cos j cos j ) + Rh cos(t ) cos j (4.6.19)

The tilt angle t here is positive for an upwind rotor, and negative for a downwind rotor, as dened in Fig. 4.1(b). The nal set of 1 + B equations is derived as:
B B

IH +
j=1

Irj
j=1 B

j (2Rh Mb sCG sin j + I sin 2j ) =

Qaero Qr
j=1

[Mb g (Rh cos(t ) sin j + sCG cos(t ) sin j cos j )] (4.6.20a)

4.6 Structural Modelling 2 I j + HactD j + (2Rh Mb sCG sin j + I sin 2j ) = 2 Haero Hact + Mb gsCG (cos(t ) cos j sin j + sin(t ) cos j )

127

for j = 1 . . . B (4.6.20b) Equation (4.6.20a) represents the azimuthal DOF and Eq. (4.6.20b) the B independent (in general) apping DOF. The aerodynamic moments, Qaero and Haero about the rotor and ap axes respectively, are dependent on the state variables , , , j , j , thereby introducing spring and damping action dependent on the aerodynamics. A damping element HactD and actuator moment Hact are also introduced. The second term in Eq. (4.6.20a) is a damper developed as a result of apping motion, which in the rotating reference frame of the rotor is manifest as a torque from Coriolis acceleration. The moment Qr is the reaction torque from the generator, assumed to transmit its torque through an innitely sti structure.1 The second term in Eq. (4.6.20b) contributes the centrifugal stiening of the rotor. In both equations, the gravity terms introduce cyclical forces with /2 phasing between the DOF. Equation (4.6.20) is rst rewritten in state-space form. The model is then solved with Matlabs ode45 solver, implementing a 4th or 5th order Runge-Kutta method with variable step size. Without a controller, is set to zero for a constant rotation speed. The state variables are time-stepped according to the internal solver algorithm. The aerodynamic calculations are only run at specied time steps, after the EOM equations have converged to sucient accuracy based on the previous aerodynamic values, using the OutputFcn facility of ode45. The time steps are specied in terms of azimuth angle, ranging from 0.5 5 . 4.6.3.2 Linearised Equations of Motion Equation (4.6.20) represents the full non-linear set of equations that describe the system. A degree of further insight may be gained by linearising Eq. (4.6.20b) about j = 0, taking t = 0, ignoring gravity and hinge moments (both aerodynamic and actuator): Rh Mb sCG 2 + 1+ j = 0 (4.6.21) I Recall that the standard denitions of critical damping cc , damping ratio , natural frequency n and damped natural frequency nd for an under-damped and un-forced
If a exible shaft(s) is used with a gearbox, this additional compliance can be very important, however a closely coupled direct-drive machine will be less sensitive to this assumption.
1

128

Chapter 4 Analytic Development

dynamic system described by displacement x are: m + cx + kx = 0 x n = k m cc = 2mn c = cc 1 2 (4.6.22a) (4.6.22b) (4.6.22c) (4.6.22d) (4.6.22e)

nd = n

A non-dimensional number relevant to the current discussion is the Locke number. As the change in aerodynamic force is proportional to cl , the ratio of aerodynamic to inertial forces is termed the Locke number and dened as: Lk = ccl R4 I (4.6.23)

Finally, examination of Eq. (4.6.21) yields the following insights: The centrifugal stiening eect is increased with Rh , which is a partial reason for why the CONE-450 employed a space-frame, since it employed very light carbon bre blades. Lk may also be dened to include the full stiness term in Eq. (4.6.21) and is a metric of equilibrium coning angle. The un-damped natural frequency of the system is found to be n = 1+ Rh Mb sCG I (4.6.24)

indicating that Rh again increases the stiness and hence frequency. 4.6.3.3 System Damping The sources of damping in the system are the embedded hinge damper, HactD , active damping imposed by active control of Hact , and any aerodynamic damping that may be present in Haero (developed by relative AOA changes with apping motion ). In the absence of any direct hinge damping, aerodynamic damping becomes critical (see 7.1.2.1). In the absence of an active actuator moment, the damping of the system may be estimated by further assuming that Haero = Haero , , V0 sef f cos [66]. The length sef f along the blade is an eective moment arm where the aerodynamic forces are assumed to act. This accounts for aerodynamic apping mo ments created by , without requiring a separate map of Haero with . A value

4.6 Structural Modelling

129

of sef f = 0.7S has been found appropriate [66]. With this linearising assumption, Eq. (4.6.21) can be expanded to include aerodynamic forces as: 1 Haero 2 j + HactD + sef f j + n j = 0 I V0 by incorporating Haero Haero Haero + j + V0 sef f j (4.6.26) V0 and taking the steady-state case where = V0 = 0 and ignoring the added Haero = stiness term
Haero

(4.6.25)

in n . The damping ratio is then computed as: = HactD + sef f Haero V0 2I n (4.6.27)

To achieve a certain minimum positive damping factor , the actuator damping must be positive to balance any negatively sloping aerodynamic hinge moment curve. This is typically associated with negative Lk (negative cl ) found in stalled conditions. The non-linear damping of the system may be studied using the full simulation code as well. In this case, damping may be computed via measurement of the logarithmic decrement [157, p. 137] in free vibration from a non-equilibrium start point. Other possible approaches include utilizing the Hilbert transform [158], generalized system identication methods [125], or work equivalence1 [159], however these methods are more complex than is required here. The logarithmic decrement d is based on the following equation: d = 1 x0 ln = N xN 2 1 2 (4.6.28)

which derives from the analytic solution to Eq. (4.6.22a), where x0 and xN are peak positive displacements N cycles apart. Critically, this equation depends on an equilibrium value of xinf = 0. The oset can be explicitly removed from the data (if known) [158], or dealt with implicitly by computing: yi = ln |xi xi+1 | (4.6.29)

where xi are now the Np extrema (positive and negative), located in the displacement time history with a peak-nding algorithm. A good estimate of d is then obtained as d = 2m, where m is the least-squares t to the yi points at indices i = 1 . . . Np 1. Finally the damping ratio is obtained from Eq. (4.6.28) and damping constant from Eq. (4.6.22), if n and m are known for the mode.
The work equivalence method is quite useful in determining the damping of individual exible-body modes, but requires accurate n .
1

130

Chapter 4 Analytic Development

4.6.4 Centrifugally Stiffened Beam


In order to include apping in BLADEDTM which incorporates exible blades, a method was required to compute the mode shapes of hinged blades. BLADEDTM takes as input a dened number of in-plane and out-of-plane modes, described by their mode shapes, natural frequencies (stationary and rotating), and assumed damping factors for each mode. The modes are computed by transforming the structural stiness through the section twist and pitch angles, using Eq. (B.8). There is no built-in facility in BLADEDTM for computing freely hinged modes. The approach taken to this problem is to model the blades with an FEM representation. The centrifugal stiening and linear variation of beam properties are somewhat non-standard and are therefore developed here. BLADEDTM uses some form of iterative technique [157] on the beam PDE for conventional xed-root blades.

4.6.4.1 Finite Element Formulation The specic approach taken to this derivation is the minimization of potential energy p = U + , following the general approach and notation of Logan [160, p. 56] where standard beam equations are presented (constant properties, without centrifugal stiening). A beam element is shown in Fig. 4.33. Linear theory is used throughout this derivation, both in displacements and material properties (i.e. = E ).

Figure 4.33 Beam element coordinate system (), displacement () and nodal x y v displacements (d), rotations (), forces (f ) and moments (m)

The shape functions dene the displacement v as: d1y 1 d2y 2

v = N d = N1 N2 N3 N4

(4.6.30a)

4.6 Structural Modelling 1 23 32 L + L3 x x L3 1 x x N3 = 3 23 + 32 L L N1 = 1 x3 L 22 L2 + xL3 x L3 1 N4 = 3 x3 L x2 L2 L N2 =

131

(4.6.30b)

The strain may be dened from its denition and the deformed geometry of the beam with axial extension u (see Logan [160]): u = y x x (, y ) = d v d x (4.6.31a) (4.6.31b)

d u d2 v = 2 y d x d x

Using the shape functions N , the strain is found as: x x (, y ) = B d y (4.6.32)

where B is the second derivative of N with respect to x: B = 1 12 6L, 6L 4L2 , 12 + 6L, 6L 2L2 x x x x L3 (4.6.33)

The stress is simply dened as: x = D The strain energy U is found to be: U=
V x

= E(, y ) x

(4.6.34)

1 x x dV 2

(4.6.35)

by neglecting shear stress (i.e. Euler-Bernoulli beam, not Timoshenko beam). Direct axial strain (owing to displacement u) is also ignored in this derivation. Substituting the above constituent quantities yields:
L L

1 U= 2
0 A

1 E(, y ) d B B dAdd = x y x 2
T 2 T 0

T EI()d B T B dd x x

(4.6.36)

where the usual integral for the moment of inertia is used: EI() = x
A

E(, y )2 dA x y

(4.6.37)

and the product EI in general varies over the length of the beam: EI() = (EI)1 + x (EI)2 (EI)1 x L (4.6.38)

132

Chapter 4 Analytic Development

Evaluating Eq. (4.6.36) and dierentiating for each DOF of the four beam variables and setting each equal to zero (minimum energy condition), yields the elemental stiness matrix k stif f :
L

k stif f =
0

EI() B T B d x x

(4.6.39)

the expanded matrix form of which is omitted here for brevity. It was found that using constant EI values (average of end section values) produced signicantly dierent results to this exact equation. The same was found for the mass matrix, dened consistently1 as: m =
V L

N T N dV

(4.6.40a)

=
0

mps () N T N d x x mps,2 mps,1 x L

(4.6.40b) (4.6.40c)

mps () = mps,1 + x

Equation (4.6.40a) may be obtained from virtual work or DAlemberts principle (see Logan [160] for the derivation). An additional term is required to account for centrifugal stiening, essentially increased stiness, k cent . Two approaches are possible to account for this: body forces ( term) or pre-stress (U term). The latter is presented here, as it is more general (e.g. vibrating tensioned string). The starting point is to consider a general strained bre in the beam, as shown in Fig. 4.34. As the bre is strained from length dS to ds = ((dS + d)2 + d2 )1/2 , u v the strain developed is [161, 3.3]:
x

= =

ds dS ds = 1 dS dS d u 1+2 + dS d u dS
2

d v dS

1/2

(4.6.41) 1

In keeping with linear theory, dS d, and using the rst three terms of the x binomial expansion of the (. . .)1/2 term above yields:
x

d 1 u + d 2 x

d u d x

d v d x

(4.6.42)

A lumped-mass formulation was found inadequately accurate.

4.6 Structural Modelling

133

Figure 4.34 Strained bre

The origin of Eq. (4.6.35) is now relevant:


x

U=
V 0

E x d x dV

(4.6.43)

obtained from the work of internal forces on a dierential volume dU = x d x dV (see [160]), substituting x = E x , and integrating over the volume. Substituting the rst term of Eq. (4.6.42) into the above yields Eq. (4.6.35). The second term in Eq. (4.6.42) is ignored as higher order, leaving the third non-linear v displacement term. This latter term may be integrated by noting that it is a constant (= f ( x )) to yield: U= 1 2
V

d v d x

dV

(4.6.44)

Equation (4.6.44) is the strain energy for the beam bending problem, owing to prestress x . With an assumed uniform stress over the cross-sections, this reduces to:
L

1 U= 2
0

d v d x

d x

(4.6.45)

where F is the force on the cross-section. The equation may be vectorized as:
L

1 U= 2
0

x F d T N T N x d d x

(4.6.46)

where N x is the rst derivative with respect to x of the shape functions N (see Eq. (4.6.30a)). The centrifugal stiness matrix kcent is obtained by taking the derivative as before:
L

k cent =
0

F N T N x d x x

(4.6.47)

134

Chapter 4 Analytic Development

It only remains to dene the section force F that varies over the blade length. Assuming that external forces perfectly balance the centrifugal forces in the y direction, the centrifugal axial force is found with reference to Fig. 4.35 as:
L

F = F2 + cos
x

mps rdx

(4.6.48)

where F2 is the aggregate of the outboard forces and mps varies as in Eq. (4.6.40c). The radius r = Rh + (s1 + x) cos depends on the hub, as Rh = Rhinge for xed and outboard apping hinges, and Rh = Rhinge cos for centrally apping and teetered roots. The nal integrated matrix form of Eq. (4.6.47) is quite large and therefore omitted.

Figure 4.35 Element centrifugal force

4.6.4.2 Assembly and Master-Slave Method The overall stiness matrix ( K ) and mass matrix ( M ) are assembled by super position (direct stiness method [160]) of the the various element matrices ( k stif f , k cent and m ), by matching DOF for each blade section element and blade.1 A central hub node (with displacement and rotation) is also included. Note that all rotational DOF are the same direction, and the displacements are normal to the un-deformed blade axes (i.e. not using general rotated elements). The Master-Slave method [162] is used to couple the blade root nodes to the central nodes. The only exception is for a hub with known stiness, in which case the hub stiness matrix is simply assembled as above. The general method begins by partitioning the matrix in master nodes u m , slave nodes u s , and uncommitted
No symmetry is in general present to allow reduction in DOF (in-plane symmetric/antisymmetric modes, teetering.
1

4.6 Structural Modelling nodes u u : k uu kT um kT us k um k mm kT ms fu k us u u k ms u m = f m k ss us fs

135

(4.6.49)

The constraint equation prescriptively couples DOF with innite stiness: Amum + Asus = g Rearranging, and solving for u s denes the matrix T : u s = A 1 A m u m + A 1 g = T u m + g s s which nally yields a reduced system of equations:
T

(4.6.50)

(4.6.51)

k uu kT um

k um T k mm T

f u k us g uu = um f m k ms g

(4.6.52)

In this application, a mass matrix is also required, g is zero and no external forces are present, giving: M u + K u =0 (4.6.53)

The T matrix is augmented from the general formulation, to map the complete set of original DOF to a reduced set u containing u u and u m . u = T u M = T K = T
T T

(4.6.54a) (4.6.54b) (4.6.54c)

M T K T

Two couplings are used between the hub and blade root nodes to dene T : Rigid link of length L 1 0 v1 1 0 1 v1 = v2 1 L 1 2 0 1 Hinge at end of rigid link of length L 1 0 v1 v1 1 = 0 1 1 v2 1 L (4.6.55)

(4.6.56)

A global T matrix is assembled in the same manner as the mass and stiness matrices by superposition, starting with T = I .

136 4.6.4.3 Boundary Conditions

Chapter 4 Analytic Development

The boundary conditions and coupling are specic to the hub. In all cases the displacement of the hub node is zero. For OP modes: Rigid No rotation of hub and rigid coupling of blade roots to hub node. Central Hinge No rotation of hub and hinged coupling of blade roots to hub node. Hinge at Rhinge No rotation of hub, no translation of blade roots. Teetered Rigid connection between blade roots and hub node. For IP modes: Brake Applied (Fixed) No rotation of hub and rigid coupling of blade roots to hub node. Free-Hub Rigid connection between blade roots and hub node.

4.6.4.4 Modal Solution The mode shapes are obtained from the eigenvalues (n ) and vectors ( x n ) returned from Matlabs eig function. This solves Eq. (4.6.53), by assuming a solution form of the form u = x exp(in t) to yield the standard eigenvalue problem n M x n = K x n . Before solution, Eq. (4.6.53) is further reduced by removing zeroed DOF. After solution, the modes are sorted from lowest to highest frequency, the displacements normalized to a maximum of 1 for the largest tip deection. The free-body (zero frequency azimuthal rotation) in-plane mode for the free-hub is also discarded.

4.7 Control
This section describes the solution of the wind turbine control problem. Section 4.7.1 presents the derivation and implementation in ExcelBEM of an optimal steady-state control scheduling algorithm, complete with equilibrium coning of the coning rotor. Section 4.7.2 then provides a brief introduction to the synthesis of a dynamic controller, including extraction of a suitable dynamic model for the task. The detailed execution of the task is left for future work, as only constant speed rotors and xed applied hinge moments are considered in this thesis.

4.7 Control

137

4.7.1 Steady State Operation


BEM programs typically solve for steady state conditions with a xed geometry and

ow conditions. The solution iterations are conned to the aerodynamics (a and a ), possibly including yaw correction factors which require azimuthally averaged and iterated values. When searching for an operating point dened in the torque/speed plane, another level of iteration is required, since there is no closed-form solution to the operating condition problem. The wake iterations presented in 4.4.11 also require iteration in a loop outside the induction factors, as does free-hinging operation. BLADEDTM includes steady-state power curve calculations, but does not include provision for a ap hinge in this mode. The denition of the control is also dierent, requiring the same torquespeed relationships used as target schedules in dynamic simulations. These must be provided by the user as either look-up tables or constants dening a generic curve. In the present work, steady solutions are required for variable geometry (equilibrium coning), as well as optimal energy capture and power limiting control schedules. The latter are used to dene operational target control schedules over the complete wind speed range. Dened schedules with wind speed for applied actuator hinge moment, cone angle, and rotor speed may also be dened. The latter two may be used either as initial conditions for the optimum and limiting schemes, or as fullydened control schedules. The challenge of optimal operation for the coning rotor is discussed further in 6.4.2.

4.7.1.1 Cone Angle Equilibrium The aerodynamic loading on the blade will change with cone angle, as will the centrifugal stiening forces, creating a non-linear system. BLADEDTM does not have a capability within its steady calculations to model this eect. Only static pre-cone may be specied. Alternatively, dynamic simulations may be run to steady state. In ExcelBEM, a secant method [131] is run as an iterative loop outside the induction factor solution. The following equation is solved for the net moment about the hinge: Maero () + Mcentrif ugal () + Mactuator = 0 (4.7.1)

138

Chapter 4 Analytic Development

The algorithm is initialized with the value of in memory and a second cone angle 10% larger (or 2 for zero intial cone angle). A solution termination tolerance of 0.1 is used on between iterates, together with solution lagging after 25 iterations.

4.7.1.2 Optimal Energy Capture and Limiting Operation The operating point algorithm computes either deterministic performance over the wind speed range, or an optimal control schedule (Region II) with or without limiting (Region III). The overall ow of these calculations is shown in Fig. 4.36. Switches for limiting (Limiting?) and optimization (Optimizing?) specify the output quantity for iteration f (xi ) at iterate xi . The f (xi ) are either power or torque, produced aerodynamically or electrically (including parasitic losses). Optimizing is only done in xi = ; limiting may be with xi = or xi = pitch . The Run BEM steps include the iterations of 4.7.1.1, if a free hinge is dened. Each of the scheduling steps is only done if a schedule is dened; otherwise nominal values are used. The unconed is limited to less than 15 at max by limiting , in addition to user-input generator rotation speed bounds. This was done to avoid numerical problems associated with free hinging and negative power (propeller mode). The limiting and optimization iterations are converged to 0.5% change in the limit and nominal values respectively; the change in variables ( or pitch ) must be less than 0.001min and 0.1 respectively. Robust search algorithms are used to guarantee proper solutions in what is a highly non-linear problem. Optimal solutions for are found with a modied Golden Section search algorithm [163]. The algorithm is initialized with bounding values min and max . The function f (xi ) in general has multiple local minima, so at each step of the algorithm the f values on the bounds (min , max ) are compared to the f (x1 ) and f (x2 ) values on the interior of the standard algorithm. Contraction of the search domain then occurs towards the absolute minimum, which may occur on the bounds. For optimization, the pitch angle is assumed saturated at its upper or lower limit for PTS and PTF control respectively, over the Region II range below rated. Figure 4.37 illustrates the speed windows dening the operational envelope of the rotor, bounded by the limits on rotation speed, for two wind speeds. Point A represents the upper speed limit at a low wind speed, and B the optimal solution for that wind speed.

4.7 Control

139

Figure 4.36 Operating point solution ow chart

The limiting iterations are done with a bounded secant method, rather than the slower Regula Falsi method [131]. The algorithm is bounded for VSS rotors by min . . . max and by pitch,min . . . pitch,max for PTS and PTF. The starting point x1 is max for VSS, pitch,max for PTS and pitch,min for PTF. An initial line search is performed in the variable to locate x2 where f (x1 )f (x2 ) < 0, or until the other bound is reached, by stepping towards the opposite bound. This line search is necessary to accommodate all potential shapes of power/torque curves (e.g. negative slopes). If the other bound is reached, the algorithm is terminated and limiting is impossible. Otherwise, the modied secant algorithm is used with x1 and x2 as starting points.

140

Chapter 4 Analytic Development

Figure 4.37 Speed windows for optimal and limiting operation

At each step, the xi step is bounded to the variable limits. The feedback path from optimized solution to limiting operates only on . This step is required in some cases when the speed window is straddling the peak of the power curve, as shown in Fig. 4.37 by point C. The wind speed is large enough in these cases that the dimensional power (or torque) exceeds the limit value. In these cases, both D and E are valid solutions, as they both correspond to the CP for limiting. The correct choice depends on the operating strategy. For VSS, Point E is an appropriate choice, since the stall half of the curve is to be tracked past rated power. Tracking D would yield lower drive-train torque, but would make a controller very hard to design. Eventually, a drastic speed reduction would be required at higher wind speeds to move from the drag-limited right half of the CP curve to the left stalllimited range. For the pitch controlled rotor, the choice is taken to track point D, since the rotor can eventually use pitch when required to limit. This is only one possible choice; ideally point C would be located, but this would require a more costly multi-dimensional optimization. In general, these are pathological cases in any event, and will only be encountered if the upper rotor speed limit is unrealistically high. In reality, there is usually some constant speed mode near rated to smooth controller mode switching and/or limit tip-speed.

4.8 Generator Modelling

141

4.7.2 Dynamic Control


The purpose and time-frame of the current project has precluded any detailed work on a dynamic controller for the coning rotor. Connor et al. [164] detailed the controller design for the CONE-450 to eect the steady-state control reference targets developed along the lines of 4.7.1. A relatively standard PI-type was used, with estimators for aerodynamic torque and consideration of control-mode switching. The design of the controller was carried out by linearising the EOM for the 2
DOF system, in that case a single central actuator (i.e. apping) and rotational DOF. A similar approach may be taken on the decoupled apping concept. Equa-

tion (4.6.20) is rst fully linearised around an operating point, taking due care of the equation parameters that are functions of the linearising parameters (, , j and j ). By then taking the Laplace transform of the equations, the constant terms of the linearisation disappear, and a set of 1 + B coupled equations result. The equations are expressed in perturbation variables of the linearising parameters. They contain parameters dependent on the mass properties of the system, and the gradients of the aerodynamic driving ( direction) and coning ( direction) torques. The latter gradients are determined numerically from maps of the aerodynamic performance, computed at the linearisation point. Using the standard representation of the system thus developed, controller synthesis techniques may then be applied for a set of operating points from start-up through shut-down. In the original work [66], Bode plot transfer function shaping was used. An alternative approach would be to numerically linearise the system directly from a full dynamic simulation, including any exible body modes. This is a relatively new approach for wind turbines, and may be carried out in BLADEDTM . The output is a state-space model described by internal state variables and control outputs. If pitch control is to be considered in a revised coning rotor design, this approach may prove expedient, and will certainly be more accurate. Of course, a whole host of controller synthesis methods may then be applied (e.g. loop shaping, PID, state-space), as they have been for standard wind turbines.

4.8 Generator Modelling


Generator analysis, similar to the aerodynamic and structural analyses presented earlier, can be conducted at various levels. The focus here is on the magnetic circuit, lying at the heart of the generator and driving the cost. The most detailed

142

Chapter 4 Analytic Development

analysis required for nal design work includes FEM simulation of the magnetic ux and temperatures through the rotor and stator, over the operating conditions of the machine. Clearly, this level of analysis is well outside the requirements of the current work. Instead, what is required is a mix of a very simple model using basic parameters, and a more involved parametric analytic model. Both will be required in Part III to place bounds on the generator capacity, so that together with the costs given in 4.9.2, the generator requirement does not become unrealistic. To that end, the following two models are considered. Section 4.8.1 computes representative values for the simplest shear stress model, while 4.8.2 briey introduces a parametric model developed by Dubois [165].

4.8.1 Shear Stress Generator Model


For electrical machines with low rotation speed, the design is driven by the torque requirement, rather than the machine power rating [166]. This is because the magnetic elds must be quite strong to develop sucient torque according to: gen = Pgen gen (4.8.1)

The magnetic forces developed by the interaction of magnetic elds (created by the magnets and energised windings) manifest themselves as an eective shear stress ag in the air gap between rotor and stator. The shear stress is itself fundamentally limited by the magnetic ux density that can be accommodated in the iron core and magnets. Considerations of rotor and stator geometry and thermal management serve to further limit the maximum value that ag can attain. The thermal management in particular has a large eect. Concepts employing free convection are quite limited relative to actively cooled designs. Of course, there are trade-os in complexity and additional cost of the cooling system. Once the general concept of the generator is determined, ag can be used to estimate an overall size: gen = 2Rlag of air-gap radius R and length l. Backing out the value of ag for a freely cooled design, the Vensys 1.2 MW [81] direct drive machine, yields approximately 14.5 kPa. For comparison, an actively cooled 1.5 MW design uses a value of 42.1 kPa [167]. (4.8.2)

4.9 Cost Modelling

143

4.8.2 Analytic Magnetic Circuit Generator Model


The equations governing magnetic elds may be integrated over the cross-sectional area and average path lengths of a magnetic device, to yield a system of equations analogous to electric circuits [168, 169]. The basic elements are the magnet M M F and winding nI motive forces, reluctance R, and ux , equivalent to the voltage V , resistance R, and current I of an electric circuit. Dubois [165] has constructed this type of model for sizing of a radial-ux generator. That model has been implemented in a spreadsheet analysis by the current author, to further investigate parametric variation of the generator design parameters in Part III. The details of the model can be gleaned from the original thesis. The only modication to the model that has been made is to change the dimensioning variables to describe an outer rotor (magnets) and inner stator (windings). For optimization purposes, a diameter D and rotation speed are specied as parameters, and eciency as a constraint (95%). The DVs are then pole pitch p , current density J, no-load ux density in the air-gap Bg , tooth/slot length ratio in the stator bt /bs , an the ratio of stator length to diameter Krad . All parameters are given realistic limits. Constraints are also included for eld saturation and shortcircuit currents. An additional constraint for ag < 25 kPa as also imposed in the present work. Using Excels solver, it has been possible to optimize for a specied power or torque requirement.

4.9 Cost Modelling


Obviously an accurate measure of the cost of any new concept must be developed and compared to the existing state-of-the-art in order to denitively demonstrate an economic advantage. Cost models vary greatly for various scoping studies, ranging from simple functions of diameter and power, through aggregated mass dependent cost functions for individual components, and ultimately vendor tenders for fullydened components. Mass dependent models [16] typically incorporate a xed and variable component: c = cref (1 ) + v vref (4.9.1)

where cref is the cost of the reference machine/component and the percentage of the cost assumed to vary with metric v (e.g. rotor area R2 , mass m). The renement of the cost model should be commensurate with the level of engineering analysis and stage in the design process. The level of detail of the cost model

144

Chapter 4 Analytic Development

will also dictate the condence that can be placed in any optimization results. The CONE-450 report [66] presented a detailed accounting from manufacturers of the costs of major components, especially those specic to the coning rotor. This level of detailed cost is appropriate at a highly rened, engineering study stage. Given the resources and focus of the current work, this level of cost detail is inappropriate and beyond the current scope. Ultimately, any revised coning concept must be accounted for in a similar manner to the original study. However, this activity is better tackled by a commercial design study. The current approach is a proxy cost metric, overviewed in 4.9.1, and more quantiable estimates for the magnetic material costs (4.9.2).

4.9.1 Proxy Cost Metric


In Part III, the strictly technical merits of the concept are of primary focus. For the most part, cost is only accounted for by a proxy metric, material mass. This approach eectively sets = 1 and v = m in Eq. (4.9.1). Even cost estimates for bulk quantities of material can vary widely, depended on market conditions and international exchange rates, making a fundamental comparison dicult. The attempt here is therefore to examine the performance trade-os of various design choices, leaving to more detailed future studies the task of determining economic optimality, in current market conditions. This approach is viable, as the rotor blades are the primary component under consideration, so that conventional materials and techniques can be compared on a like-for-like basis. This should make mass a fairly consistent and reliable measure of cost-eectiveness when comparing the coning rotor to a standard rotor. As already mentioned, parts such as the coning actuators and hub are unique to the concept, and hence will require more detailed work in the future. Based on the original work, they should not drive the cost of the concept to any great extend, although reliability may be of concern.

4.9.2 Magnetic Materials Cost


One component that has been examined in more detail is the generator, for reasons that will become clear in Chapter 6. For a generator, the value of in Eq. (4.9.1) will be quite high for the active materials used. Some notion of the installed costs of these materials has therefore been collated in Table 4.2. The costs include both raw material and assembly into a generator. They clearly cover a large range, confused

4.9 Cost Modelling

145

further by currency conversion, temporal supply/demand variation aecting cost, and the details of the generator design (winding layout, fabrication techniques). The magnet costs are representative of rare-earth materials (NdFeB most common). Table 4.2 Active magnetic specic material costs Item Copper Iron Magnet
a

Dubois [170] (e/kg) 6 6 40

Polindera(e/kg) 1025 3 20

WindPACT [167] ($US/kg) 13 5 50

Data from personal communication with Dr. Polinder, Delft University of Technology, Laboratory of Electrical Power Processing, Delft, The Netherlands, September 2004

Chapter

Validation
In order to have condence in the analysis methods of Chapter 4 for use in the design work of Part III, they must be validated to the extent possible. The aerodynamic models are validated rst in 5.1, against externally available CFD and experimental results, the former for both idealized and real geometry. The acoustic model is checked next in 5.2, followed by the structural sectional and beam models in 5.3. In 5.4, the outputs of BLADEDTM are investigated to ascertain the modications required to properly analyse the coning rotor. Finally, a brief verication of the generator model is made in 5.5.

5.1 Aerodynamic Validation


Aerodynamic validation is done for three increasingly complex cases, for both coned and unconed rotors. First, a uniformly loaded theoretical rotor is examined numerically (5.1.15.1.4), to exclude blade shape and airfoil eects. Secondly, numerical simulations are compared for a real rotor in 5.1.5, to include airfoil eects. Wind tunnel test data for a large rotor is then compared to ExcelBEM predictions in 5.1.6. The aerodynamic behaviour in dynamic inow conditions is then compared to BLADEDTM in 5.1.7. Finally, numerical predictions are compared for a yawed rotor in 5.1.8. It should be noted that experimental results for wind turbines are quite sparse, and the NREL UAE in C.2 is virtually the de facto standard comparison dataset. The primary problem in turbine testing, as opposed to say aircraft, is the requirement for large models to simultaneously match Re, and geometric similarity. Atmospheric testing is troubled by the non-uniform inow that cannot be input exactly into simulations for comparison. Very large wind tunnels are therefore required, doubly so since the wake must freely expand, demanding tunnels much larger than the rotor. The results in this section are identied by the following labels and unless otherwise noted, indicate results from the modied BEM method of 4.4 (ExcelBEM): 147

148

Chapter 5 Validation

CT Bladed and CT Spera indicate the thrust model used in ExcelBEM (see 4.4.3). Unex and Ex indicates unexpanded and expanded wake geometries in ExcelBEM (see 4.4.11). Unex is implicit unless otherwise stated. Madsen and Mikkelsen indicate CFD study results. BLADED indicates BLADEDTM code results.

5.1.1 Uniformly Loaded Rotor


The classic uniformly loaded rotor (with no tangential loading f ) provides a valuable check on the modied BEM method. At low induction factors with an innite number of blades, the vortex model geometry should be an almost exact representation of the ow conditions. With constant loading, vorticity is only shed at the tips. No swirl is present for the purely axially loaded rotor (i.e. a irrelevant). No airfoil data is used, avoiding any ambiguity associated with stall delay, etc. Standard BEM formulations predict a constant induction a along the blade, regardless of cone angle or CT . The power and thrust coecients are also constant when referenced to the coned disc area. The numerical results from two full-eld CFD studies were used for comparison1 : Madsen and Rasmussen [106] used an axisymmetric Navier-Stokes model with volumetric body forces representing the blade loading; Mikkelsen et al. [88] and [89] adopted a vorticity-swirl-streamline model, in both axisymmetric and fully 3D actuator line formulations, but only the axisymmetric portion of the work is relevant here. Madsens study investigated only downwind coning on straight, winglet equipped and curved blades. Mikkelsen only looked at straight rotors, but covered upwind and downwind coning, as well as real rotor performance, to be considered in 5.1.5. The unit area axial loading fu,z of the disc is specied according to the desired thrust coecient CT . For the applied loading to be perpendicular to the coned rotor surface, the normal unit loading fu,n at the set coning angle is found from: 1 fu,z = fu,n = CT V02 2 Power output is calculated from:
R

(5.1.1)

P =
0
1

V1 fu,n 2rds

(5.1.2)

Note that the CFD studies dene cone angle in the opposite sense.

5.1 Aerodynamic Validation

149

Numerically this equation is evaluated between each blade section assuming linear variation in velocities (induction factors, , etc.).

5.1.2 Uniform Loading Results


The unconed rotor is examined rst in 5.1.2.1, followed by the coned rotor in 5.1.2.2. 5.1.2.1 Unconed Rotor The most basic case of the unconed rotor is presented rst in Fig. 5.1. The current method with no wake expansion predicts uniform induction, as per the standard theory. With the inclusion of wake expansion however, the correct trend in induction (increasing towards tip) is predicted. The thrust model is also observed to have a large eect on the results, with the Spera model giving lower a as expected from Fig. 4.5. Note that Mikkelsens CFD results deviate from Madsens, illustrating that even a full-eld method is subject to considerable uncertainties in this simplest of cases.
0.50 0.45 0.40 0.35 CT_Bladed, Unex CT_Spera, Unex CT_Bladed, Ex CT_Spera, Ex Madsen Mikkelsen

az
0.30 0.25 0.20 0.15 0.0 0.2 0.4 0.6 0.8 1.0

r/Rtip

Figure 5.1 Uniformly loaded axial induction ( = 0 , CT = 0.89)

At low induction factors, the CFD method, like BEM without expanded wake geometry (the light grey horizontal lines in Fig. 5.2), predicts a uniform induction. As CT rises from 0.4 to 1.0 in Fig. 5.2, the induction factor near the tip of the disc grows in both the CFD results and the corrected BEM method with wake expansion.

150

Chapter 5 Validation

Again, this is predicted qualitatively by the vortex model, as outboard sections see more inuence from the proximal shape of the sheet. Near the centre of the disc, the induced velocity is reduced as the wake moves outboard, away from the disc.
0.60 0.55 0.50 0.45 0.40 CT = 0.4, CT_Bladed 0.4 Mikkelsen 0.8 CT_Bladed 0.8 Mikkelsen 0.89 CT_Bladed 0.89 Mikkelsen 1.0 CT_Bladed 1.0 Mikkelsen

az

0.35 0.30 0.25 0.20 0.15 0.10 0.0 0.2 0.4 0.6 0.8 1.0

r/Rtip

Figure 5.2 Uniformly loaded axial induction as a function of loading CT ( = 0 )

5.1.2.2 Coned Rotor The present formulation is able to predict the correct induction factor trend for both upwind () and downwind coning (+), as shown in Figs. 5.3 and 5.4 respectively. The variation over a wider range of cone angles excluding wake expansion is shown in Fig. 5.5. The downwind results of primary interest are better predicted, for reasons elicited later. The somewhat unintuitive result that the CT Spera induction factors are higher at the centreline than the CT Bladed ones in this case owes to the adisc indexing of the the thrust models.1 Induction factors normal to the blade an follow similar trends to the axial ones a = az and are not discussed further. Returning to the ow eld picture in Fig. 4.8, for positive coning, the disc is sitting in a region of reduced inuence from the wake cylinder, while the reverse is true for negative coning angles. In order to balance the momentum equations in
Referring to Fig. 4.5, for the same CT , the Spera model has a lower adisc . This is still true in the downwind case, from the denition of adisc .
1

5.1 Aerodynamic Validation


0.50 0.45 0.40 0.35 CT_Bladed, Unex CT_Spera, Unex CT_Bladed, Ex CT_Spera, Ex Madsen Mikkelsen

151

az
0.30 0.25 0.20 0.15 0.0 0.2 0.4 0.6 0.8 1.0

r/Rtip

Figure 5.3 Uniformly loaded axial induction ( = 20 , CT = 0.89)

0.50 0.45 0.40 0.35

az
0.30 0.25 0.20 0.15 0.0 0.2 0.4 0.6 0.8 1.0

CT_Bladed, Unex CT_Spera, Unex CT_Bladed, Ex CT_Spera, Ex Mikkelsen

r/Rtip

Figure 5.4 Uniformly loaded axial induction ( = 20 , CT = 0.89)

152
0.65 0.60 0.55 0.50 0.45 0.40

Chapter 5 Validation
-40 -30 -20 -10 0 10 20 30 40

az

0.35 0.30 0.25 0.20 0.15 0.10 0.05 0.0 0.2 0.4 0.6 0.8 1.0

r/Rtip

Figure 5.5 Uniformly loaded axial induction as a function of (deg) for Bladed thrust model (CT = 0.89, no wake expansion)

light of greater than and less than unity be inversely modied.

respectively, the induction factors must

The aggregate power coecients (based on coned area) are compared next in Table 5.1. At CT = 0.89, classic BEM theory predicts a power coecient of 0.593, the Betz limit. The CFD models and modied BEM all predict reasonably close to this value for the straight and downwind coned rotor in all cases. The invariance of CP found here and in the CFD results is explained by the wake structure, which for the uniformly loaded rotor is determined uniquely by the thrust level. The attendant wake structure results in identical power outputs for equally sized rotors, when a control volume bounded by the wake and far-eld upstream/downstream boundaries is considered. Table 5.1 Uniformly loaded power coecients (CT = 0.89) CP Mikkelsen Madsen CT Bladed, Unex CT Spera, Unex CT Bladed, Ex CT Spera, Ex 0 0.601 0.573 0.593 0.619 0.586 0.616 (deg) 20 0.601 0.571 0.593 0.618 0.588 0.615 -20 0.604 0.592 0.620 0.586 0.616

5.1 Aerodynamic Validation

153

Within like thrust and wake expansion models, the variation of CP with is at the level of numerical accuracy for the modied BEM method. The Betz limit was of course derived from the momentum thrust equation, which from Fig. 4.5 may be slightly low at this induction level. This eect provides some explanation for Mikkelsens and the CT Spera cases exceedance of the Betz limit.

5.1.3 Fundamental Model Deficiencies


The results obtained demonstrate that it is too simplistic to dismiss BEM as invalid because of the stream-tube independence assumption or planar disc derivation. It is in fact the classic Bernoulli derived relation between disc induced velocity and fareld velocity that is presumptuous, once a radial ow component is included and the mass ow term correctly derived. In fact, Glauert discussed the induced velocities of the shed vortex wake as the key link between momentum and blade element theory [98]. For convenience, the Bernoulli rationalisation is usually presented, but this approach bypasses the core understanding of the physical situation required to account for coning. Clearly, there are still deciencies in the present method, even for the most basic uniformly loaded case. As stated, the method is strictly only valid for lightly loaded conditions, implying no expansion of the wake. Based on the results, inclusion of this geometrical eect is capable of resolving the bulk of the disparity between reality and the standard BEM method. Leaving aside problems with 2D airfoil predictions, such as stall delay, present for real rotors (and not present for the uniformly loaded rotor), the key model uncertainty is in the thrust model, as evidenced by the variation in the results with the chosen model. The deviation is much worse for rotors with large a, encountered more in the upwind case than the downwind one. It is instructive to consider the origin of the breakdown in the momentum thrust model. For a lightly loaded rotor, the wake is conned to a thin shear layer, well modelled by a thin sheet of vorticity. As the loading increases, the layer thickness will grow and follow the tip streamline (i.e. expand), but is still represented well as a vortex sheet. At still higher loadings, the wake thickness will grow even larger before nally lling the entire wake space as unsteady large-scale eddies. This is analogous to inviscid airfoil modelling, which predicts ever increasing lift, by neglecting the large separated region at high angles of attack. Even strongly coupling a viscous boundary larger to an inviscid outer ow is only valid up to a limit slightly beyond stall.

154

Chapter 5 Validation

The CFD results were obtained by inviscid steady-state formulations. As a eld method, it is capable of resolving the thickening of the wake shear layer, and so correctly predicts the induced velocity at the disc. Mikkelsen does report however, that as loading is increased past CT = 0.89 and for low wind speeds (both high a conditions), steady-state convergence is increasingly hard to obtain, eventually becoming impossible. The unsteady ow is also mentioned as an upper constraint on in Nygaards Euler ow solution [171]. The wake state of a uniformly loaded rotor has been studied specically using a time-marching full-eld scheme to track the development of the ow eld from initial uniform conditions to steady state [172]. It was found that the turbulent wake state is in fact an unstable transition between the windmill brake state and a vortex ring or propeller brake state ow, characterized by a recirculation region through and upstream of the rotor. Methods based solely on vortex laments, using the Biot-Savart equation to evaluate induced velocities, have been tried to correct for high tip speed conditions [128, 173175]. Numerical convergence problems appear and in any case they are steady formulations that are only capable of accounting for wake expansion, as is done here. The pitch of the wake laments is considered, but in one approach degenerates to the momentum thrust equation [174]. Yet another study assumes all turbulent mixing to occur in the very far wake [176], which must occur for the ow to return to free-stream, patched onto the conventional up-stream-downstream inviscid analysis. The thrust models are therefore an attempt to capture unsteady viscous behaviour within the connes of a steady inviscid method. Both the Spera and Bladed models are ts to the experimental data points in Fig. 4.5, which are most likely unsteady conditions as evidenced by their scatter and CFD results [172]. BEMs implicit stream-tube independence assumption is therefore hard pressed to deal with either the unsteady mixing or steady-state recirculation ow elds. The upwind coned case is less accurate because of geometry. As the wake sheet thickens, the meanline moves towards the centreline, and for the upwind case towards the blade. This in turn increases
z

which is below unity in this case, and

hence reduces the predicted induction. This eect can be seen by comparing the unexpanded and expanded wake cases in Fig. 5.4; in the latter the wake is further from the centre of the rotor and so
z

is larger and a smaller. The eect can also be

thought of as a result of the stream-tube independence assumption, since turbulent mixing will bring energized ow in from the tip towards the core of the ow. This

5.1 Aerodynamic Validation

155

will have a non-uniform eect on the inboard sections of an upwind coned rotor as the ow mixes in downstream of the rotor tip towards the root sections. The straight and downwind coned cases are relatively insensitive in a inboard because of the relative position of wake and blade. Further prediction of this eect is beyond the scope of this work, as the downwind case is of primary interest, but is clearly required for a better understanding and prediction of high a performance.

5.1.4 Radial and Far-Field Flow


The vortex model conrms the CFD results that the distribution of radial velocity along the blade is nearly independent of cone angle. Figure 5.6 shows the predicted radial velocity for the CT Bladed expanded wake. Although the magnitude is marginally and uniformly higher than the CFD results, the variation between cases is negligible.
0.5 -20 CT_Bladed, Ex -20 Mikkelsen 0 CT_Bladed, Ex 0 Mikkelsen 20 CT_Bladed, Ex 20 Mikkelsen

0.4

0.3

vr/V0
0.2 0.1 0.0 0.0 0.2

0.4

0.6

0.8

1.0

r/Rtip

Figure 5.6 Uniformly loaded radially induced velocity as a function of (deg) (CT = 0.89)

This result is best explained by reference to Fig. 4.8, noting the radially induced velocity isocurves. As they are centred about the tip, and a wake will be shed from that point in the coned and unconed congurations, the radial component is almost constant in a circle about the tip. The result is only slightly skewed by the fact that a constant tip radius requires a longer blade for a coned rotor. The isocurves will therefore cross the blade at modied points. Notice that Mikkelsens BEM

156

Chapter 5 Validation

method radially induced velocity vector (normal to blade) is in the wrong direction for upwind coning, since the radially induced velocity is in fact outboard in all cases. The CFD results predicted uniform induction across the far-wake, inside the limiting stream-tube. This is consistent with the vortex model which predicts a uniform induction downstream away from the leading edge of the sheet (analogous to the magnetic eld in an innite solenoid).

5.1.5 Tjaereborg Rotor


The modied BEM formulation is immediately applicable to real rotor geometries, unlike the more qualitative results presented by Chaney et al. [100] for an assumed unmodied induction factor. Mikkelsens CFD study included results for the Tjaereborg rotor, which is considered presently (see Appendix C.1 for machine details). The root of the blade is at 1.46 m radius; however, the CFD results were presented to the centreline. No 3D stall delay eects (see 4.4.8 and 4.4.9) were included in the CFD study, so spanwise ow eects were not included in the ExcelBEM runs either. The induction factors shown are the product of tip loss factor F and calculated a value, to show the average induction. All single operating point results are for 10 m/s free-stream with a rotor speed of 22.1 RPM. 5.1.5.1 Unconed Rotor The induction factor distribution for the unconed rotor is shown in Fig. 5.7. The eect of the thrust model is again evident, lowering the curve for the Spera model results. No results including wake expansion corrections are shown,1 as the inclusion of tip loss factor F mitigates the tip-eects seen in the uniformly loaded case, by forcing aF to zero at the tip. The Bladed thrust model results are almost identical to those from BLADEDTM , since radial ow does not aect either the axial momentum balance or ow relative to the section (i.e. aerodynamic force calculation) for an unconed rotor. 5.1.5.2 Coned Rotor The predicted induction factors are examined rst in 5.1.5.3, as these are good indicators of the ecacy of the aerodynamic model. Loading predictions are presented second in 5.1.5.4 and are of ultimate interest in performance prediction for design.
Almost identical results to the unexpanded case were obtained by including wake expansion.
1

5.1 Aerodynamic Validation


0.50 0.45 0.40 0.35 0.30

157

az

0.25 0.20 0.15 0.10 0.05 0.00 0.0 0.2 0.4 0.6 0.8 1.0 CT_Spera CT_Bladed Mikkelsen BLADED

r/Rtip

Figure 5.7 Tjaereborg axial induction ( = 0 )

5.1.5.3 Induction Results The downwind coning case is examined rst in Fig. 5.8. As the radial velocity component becomes important, the BLADEDTM results are much higher than all of the other results. The modied BEM results for both Spera and Bladed wake models track the magnitude and trend of Mikkelsens CFD results quite well. As ac is lower for the Spera thrust model, it predicts lower a values, compared to the Bladed thrust model that produces results closer to the CFD ones, especially over the outer portion of the blade. The alternate wake geometries (see 4.4.12) provide a marginally better prediction at mid-span, but overall degrade the accuracy of the results. Results for the single inboard model are not shown inboard of 0.2Rtip , as it predicts zero induction (
z

= 0)

for r < 0.2Rtip . Obviously the chosen vorticity distribution for the multiple vortex model is arbitrary; iteration on the BEM calculated loading back into the shed vorticity calculations would have doubtlessly improved the results. Unfortunately, that iteration would detract from the simplicity of the BEM equations. Figure 5.9 shows the upwind coning predictions for the various models. The Spera thrust model again predicts lower a than the Bladed thrust model, as the induction factors are above ac in both models. This is particularly evident near the tip where aF rises for the Bladed thrust model, since the values of a become quite large as

158
0.50 0.45 0.40 0.35 0.30

Chapter 5 Validation

az

0.25 0.20 0.15 0.10 0.05 0.00 0.0 0.2 0.4 0.6 0.8 1.0 CT_Spera CT_Bladed Mikkelsen BLADED CT_Bladed, Single CT_Bladed, Mult

r/Rtip

Figure 5.8 Tjaereborg axial induction ( = 20 )

F 0. The single vortex wake geometry moderately improves the results near to the attachment point at 20%Rtip . Otherwise, both alternate wake geometries again generally degrade the results. The modied BEM method does again follow the right trend, outperforming the BLADEDTM results in the sense that BLADEDTM is invariant with the coning direction. The radial component becomes more important in the real rotor case, as the disc loading must be predicted from the velocity relative to the blade, creating a double dependence on the radial velocity. The predictions for < 0 are generally poorer than those downwind. This may be explained by reference to Fig. 4.8, noting that a cylindrical vortex sheet does not produce much inuence upwind and at greater radii than the sheets own outer radius. For the downwind coned rotor, this means that most of the rotor is sitting in a region of reduced sensitivity to the sheets, due to the variable loading along the blade length. In the upwind coned case however, the entire blade is inboard and behind the forward lip of the sheets shed by the outer portions of the blade. The cumulative inuence is therefore much greater and felt increasingly towards the root, as evidenced by the increasing deviation between BEM and CFD results in Fig. 5.9 towards the root.

5.1 Aerodynamic Validation


0.50 0.45 0.40 0.35 0.30

159

az

0.25 0.20 0.15 0.10 0.05 0.00 0.0 0.2 0.4 0.6 0.8 1.0 CT_Spera CT_Bladed Mikkelsen BLADED CT_Bladed, Single CT_Bladed, Mult

r/Rtip

Figure 5.9 Tjaereborg axial induction ( = 20 )

5.1.5.4 Loading Results The aerodynamic loading on the blade is shown in Fig. 5.10. For the unconed case, the CFD and BEM results agree quite well. The BLADEDTM and CT Bladed results are almost identical for the unconed rotor, demonstrating recovery of the standard BEM results at = 0 . When coned downwind, the modied BEM method tracks the CFD results fairly well, under-predicting around r/Rtip = 0.8. For upwind coning the loading is uniformly under-predicted. BLADEDTM predicts the same distribution for , under predicting the modied BEM method and CFD in all cases for r/Rtip < 0.9. In terms of integrated loading at s = 1.46 m (the blade root location), for = 20 the shear force predicted by BLADEDTM is 6.6% below the modied BEM results and bending moment 4.9% too low. For = 20 , the loading under-predictions are 3.8% and 5.2% respectively. The other main aggregate performance metric is the power coecient CP (based on coned radius), shown in Fig. 5.11. The gure is fully non-dimensional, but obtained from the same operating points as used in the CFD study, which set the average pitch angle of the blade as a function of wind speed. With no coning, the CFD results are higher but in relative agreement. The shift with downwind coning is similar for both CFD and modied BEM method, given the initial discrepancy at = 0 . The upwind coning prediction is much less well predicted by the current

160
3000 2500 2000 1500 1000 500 0 3000 2500 2000 1500

Chapter 5 Validation

Fx, aero (N/m)

0.0

0.2

0.4

r/Rtip

0 Mikkelsen 1000 0 CT_Bladed 0 BLADED 20 Mikkelsen 500 20 CT_Bladed 20 Bladed 0 0.6 0.8 1.0 0.0

Fx, aero (N/m)

-20 Mikkelsen -20 CT_Bladed 0 Mikkelsen 0 CT_Bladed 20 Bladed 0.2 0.4

r/Rtip

0.6

0.8

1.0

Figure 5.10 Tjaereborg aerodynamic loading as a function of (deg)

method, due to the reasons previously discussed. BLADEDTM is of course only capable of a single prediction for an absolute coning angle. The shift in CP is much larger than either modied BEM or CFD in both cases. The discrepancy at = 0 for both BEM methods presumably relates to the thrust model, which is absent from the CFD predictions. This is evidenced by the increasing agreement between CFD and BEM towards lower , as BEM transitions from the empirical thrust curve to the momentum derived one. The modications considered in this section all relate to a more accurate treatment of induction at the rotor, and can therefore be expected to have reduced signicance at low (low induction). This is shown in Fig. 5.11, but note that the corrections are substantial down to = 5, corresponding to a V0 of 14 m/s for the Tjaereborg machine. This covers the critical Region II energy capture wind speed range when the rotor is tracking optimal CP up to rated power. A coning rotor derives its energy capture advantage in this region, so accurate prediction of power over this range is important. For the coning rotor, the accurate prediction of root bending moment is also important, for determining operational coning angle and loading.

5.1.6 NREL UAE


The UAE dataset provides a unique repository of data collected from a large-scale, controlled wind tunnel rotor test. Appendix C.2 gives the background and relevant

5.1 Aerodynamic Validation


0.55 0.50 0.45 0.40 0.55 0.50 0.45 0.40

161

CP

0.35 0.30 0.25 0.20 4 6

CP
0 Mikkelsen 0 CT_Bladed 0 BLADED 20 Mikkelsen 20 CT_Bladed 20 Bladed

0.35 0.30 0.25

10

0.20 12 4

-20 Mikkelsen -20 CT_Bladed 0 Mikkelsen 0 CT_Bladed 20 Bladed 6

10

12

Figure 5.11 Power coecient map for rotor operating points as a function of (deg)

technical details of the UAE dataset. Two test sequences are of immediate interest in relation to exploring coning rotor aerodynamics, from among the numerous data campaigns available. The rst is Sequence S, the upwind baseline case with no probes installed on the baseline blade. The blade tip pitch was 3 , teeter was locked out with rigid links, and yaw angle 0 . The second case, Sequence F, is for a downwind rigid rotor coned to 18 . Sequence S has data available from 525 m/s and F from 1020 m/s.1 Sequence C was the downwind baseline case, but was not examined here because it had a rigid cone angle of only 3.4 .

5.1.6.1 NREL UAE Baseline Before the results of the wind tunnel tests were made public, a competition was held to predict the results of the various test sequences. The predictions were made solely on the geometric and operational parameters of the turbine, and a number of airfoil datasets for the blade airfoils were provided. A variety of codes were used from BEM to full CFD. It turned out that even for the most simple, steady case, Sequence S,
Hand et al. [177, p. 17] states that excessive inertial loading prevented operation at lower wind speeds. Upon clarication with NREL, this comment refers to an excessive centrifugal force along the blade that is not counterbalanced by the thrust loads until higher wind speeds.
1

162

Chapter 5 Validation

the predictions varied greatly from around stall to 25 m/s [178]. Only one full CFD method was able to predict the aerodynamic behaviour with any accuracy.1 Since then, numerous authors have analysed the data [104, 109], focusing on the stall-delay phenomenon. Gerber et al. [179] have used a vortex lattice method to determine the ow conditions at the blade sections (principally AOA), to then back out the lift and drag properties from the pressure tap measurements. It appears that a complicated stall region is developed on the blade, with the mid-section isolated by strong vortex stall fences inboard and outboard. All this is to say that all but the most advanced CFD codes have signicant trouble predicting this set of data. 5.1.6.2 Sequence S Figure 5.12 presents predictions and measurements of the power curve and normal force coecients (cn ) for Sequence S. Note the error bars for a few points on the measured data. Even in the controlled conditions of the wind tunnel, there was signicant variation of power with azimuth angle. The results are fairly typical of other BEM predictions [104], matching very well before stall occurs. The BLADEDTM and ExcelBEM results without stall delay (Plain) are in exact agreement, and under-predict power drastically in the stalled region. Including stall delay in ExcelBEM (w SD) improves the prediction markedly, but still with signicant deviation from the measured results. The cn plots (obtained from pressure tap data) clearly illustrate the delay in stall of the sections. While the inboard sections under-predict cn with the stall delay model, the outboard ones over-predict the power, leading to the over-prediction and underpredictions seen respectively in lower and upper parts the power curve. 5.1.6.3 Sequence F Bearing in mind the diculties of predicting even the most simple case, the results for the coned Sequence F are given in Fig. 5.13. The rst point to note is the generally low induction factors in Fig. 5.13(b), even at low wind speeds. This means that unfortunately, the results are relatively insensitive to the aerodynamic coning corrections. The error bars on the measurement results are also much larger, presumably owing to the greater inuence of the coned rotor inertia on the torque strain gauge measurements.
The machine and blades were quite rigid, so structural response had only a second-order eect on the results.
1

5.1 Aerodynamic Validation


12 10 3.0 2.5 2.0

163

Aero power (kW)

8 6 4 2 0

Measured BLADED Plain w SD 5 10

Cn

1.5 1.0 0.5 0.0 Measured Excel BEM 5 10 15 20 25

Wind speed (m/s)

15

20

25

Wind Speed (m/s)

(a) Power curve


1.5 1.5

(b) 0.30r/R

1.0

1.0

Cn

0.5 Measured Excel BEM 0.0 5 10 15 20 25

Cn
0.5 Measured Excel BEM 0.0 5 10 15 20 25

Wind Speed (m/s)

Wind Speed (m/s)

(c) 0.63r/R

(d) 0.95r/R

Figure 5.12 NREL UAE Sequence S measurements and predictions (cn data include centrifugal stall delay)

The BLADEDTM and Plain results are quite similar, the dierence owing to the geometrics of the two solution algorithms. Addition of the centrifugal pumping stall delay eects (with SD) improves the predictions markedly, especially around rated. Adding the coning correction factors
z

and

(w C and w C & SD)

only slightly modies the results without and with stall delay. Finally, including sweep stall delay eects as well (w C & SD & S) improves agreement with the measurements. Ultimately, experimental results at higher induction factors, and hopefully less complicated stall behaviour, are required to better compare analytic

164
12 10

Chapter 5 Validation
0.16 0.14 0.12 0.10 0.08 0.06 0.04 0.02
20
Measured BLADED Plain wC w SD w C & SD w C & SD & S

Aero power (kW)

8 6 4 2 0

Wind speed (m/s)

10

15

0.00

5 7 9 0.1 0.3 0.5

6 8 10 0.7 0.9

r/R

(a) Power curve

(b) Induction factors for low wind speeds (m/s) with coning corrections and stall delay

Figure 5.13 NREL UAE Sequence F measurements and predictions

predictions.

5.1.7 Dynamic Inflow


ExcelBEM in unsteady ow has been validated against BLADEDTM , to the extend possible (i.e. without coning or yaw). First, a simple blade of uniform chord and twist and airfoils with 2 lift curve were run with dynamic inow solution method. No dynamic stall model was turned on. The dierence in results between BLADEDTM and ExcelBEM was extremely small, and related to the solution tolerance on a. Next, the same blade prole but with LS1-GH airfoils (from C.3) was run to obtain results for both equilibrium and dynamic wake solutions. Again no stall delay model was used, but wind shear was added. Again, the results were in agreement, but required a solution tolerance on a of 105 for exact agreement. Finally, the dynamic stall models were turned on in both BLADEDTM and ExcelBEM. As expected, the results were not exactly identical, as expected from the dierence in dynamic stall models. Both models did produce lift hysteresis loops. The BLADEDTM model exhibited this feature well below stall as well, since it is more advanced and includes unstalled unsteady aerodynamic eects.

5.1 Aerodynamic Validation

165

5.1.8 Yawed Flow


Mikkelsen [89] also compared his CFD results to actual test data for the real Tjaereborg machine C.1. The datasets available were azimuthally binned apwise bending moments (properly My ) at the blade root (r = 1.46 m). The CFD results, Mikkelsen in Fig. 5.14, are compared against the experimental Exp, BLADEDTM BLADED and predictions from ExcelBEM. The CFD results (and of course Exp) include a exible structural model, while the rest do not. The dynamic inow model of 4.4.14.6 is used by default, except for the steady (equilibrium wake) formulation for the Corr Stdy results. The Corr results include the terms from vortex theory, while the NoCorr results use the default values that transform the equations to the standard ones in BLADEDTM .1 The wind shear exponents and hub-height wind speed V used (based on experimental data) are noted in the gures. The nominally un-yawed case in Fig. 5.14(a) shows good agreement between the various BEM formulations. There is a relative oset of the tower eect in all cases past = 180 in the CFD and Exp results. This eect, and the lower peak moments for yaw = 3 , are most likely related to the structural exibility that is unmodelled in the BEM results. Noting this un-yawed discrepancy, the other yawed cases are generally in good agreement with the Exp and CFD results, when the correction factors are included. The reduction in the moment decrease through the tower wake of Fig. 5.14(b) is likely a structural modelling eect. The prediction improvements for the revised
BEM method are most noticeable in the peaks near = 90 and 270 where the

wake inuence is greatest. In general, it appears that the corrected steady formulation is a closer match to the comparison data, relative to the dynamic inow formulation. This is understandable, given that the ow is in fact steady. Even so, the dynamic formulation articially introduces a time-lag, with the artefact that the changing blade forces and induction factors with azimuth are accelerating the ow. In fact, they should be merely varying with local ow conditions created by a steady-state wake.

The formulations are not entirely equivalent, owing to the dierent treatments of induced and structural velocities in the iterative solution.

166
750 700 650 600

Chapter 5 Validation

700

My (kNm)

650

My (kNm)
0 90 180 270 360

550 500

600

550

450 400

500

90

180

270

360

Azimuth (deg)
(a) yaw = 3 , V =8.6 m/s, = 0.17
650 600 550 500 550 500 450

Azimuth (deg)
(b) yaw = 32 , V =8.5 m/s, = 0.31

My (kNm)

450 400 350 300 250 200 0 90 180 270 360

My (kNm)

400 350 300 250 200 150 0 90 180 270 360

Azimuth (deg)
(c) yaw = 51 , V =8.3 m/s, = 0.27
700 Exp 600 Mikkelsen BLADED NoCorr Corr

Azimuth (deg)
(d) yaw = 54 , V =7.8 m/s, = 0.30

Corr Stdy

Figure 5.14 Tjaereborg root bending moments in yawed ow

5.2 Acoustics Validation

167

5.2 Acoustics Validation


Validation of the acoustics model was hampered by a lack of experimental or computational comparison data. Only qualitative comparison has been possible, in terms of approximate magnitude and frequency response, as reference machine data was not available. Fortunately both sets of results are for large, downwind machines. Wagner et al. [132] provides gures for the WTS-4 turbine (2 MW in 12 m/s wind), both frequency response and directivity patterns, including detailed and simplied models and experimental measurements. Sounds levels were found to be up to 80 dB for an observer 91.5 m downwind at ground level, with a peak near 5 Hz experimentally and roll-o to 65 dB at 20 Hz. The predictions followed fairly closely, except below 2 Hz where instead of continuing to decrease, it spiked to 95 dB. Spatially, lobes were found up-wind and downwind for an observer 200 m away from the tower base. Barman et al. [141] examined the Maglarp 3 MW machine, and again presented experimental and predictive results, both for actual pressure sound level. The observer location was 0.12 km downwind at ground-level. The experimental results showed pressure spikes of 1 Pa about a mean, whereas predictions were 1.6 3 Pa. Both were associated with passage through the tower wake. The predicted
SPL peaks were at 75 dB at 10 Hz, decreasing to 35 dB at 50 Hz and 60 dB at 1.6 Hz

either side of the peak. As comparison, results for the REF-1500 (see C.3) were computed for the upwind conguration, using ExcelBEM. The observer in all of the following is 100 m downwind in a wind of 11 m/s (rated power, but below pitch action). The rotor speed is 20 RPM, conveniently giving a 3P frequency of 1 Hz. The SPLs given are for the summed octave bands to 50 Hz.1 Post-processing is done for re-sampled data at 1000 Hz with a 4096-point FFT. Shaft tilt, wind shear, and tower inuence are excluded to start. All forms of stall delay are turned o, and dynamic wake modelling is used. Figure 5.15 presents pressure and SPL for the isolated rotor. Even without tower inuence, it is clear that the moving forces produce a pressure signature at the observer location. The primary source is the near-eld, as expected from the original equations (see 4.5.3.2). As will be seen throughout, the thickness noise is negligible.
This excludes the zero, near-zero and high frequency components, the former of which is not noise and the latter inaccurately predicted by the current method.
1

168
100

Chapter 5 Validation

SPL (dB)

0 -0.1 -0.2 thk near far total

50 0 -50 -100 0 100 1 2 3 4 5 6 7 8 9 10

p (Pa)

-0.3 -0.4 -0.5 -0.6 -0.7 -0.8 0 2 4

f (Hz) SPL (dB)


50 0 -50 -100 1.6 3.1 6.3 12.5

tObs (s)
(a) Pressure

f (Hz)
(b) SPL

Figure 5.15 Simple isolated rotor acoustic prediction

Figure 5.16 shows the eect of including the tower inuence (without tilt). The tower cd is 0.3 (above Recrit ). The frequency content is clearly centred at the blade passing frequency. Three cases are shown for increasing renement of the aerodynamic simulation time-step, governed by the azimuthal step in degrees. It is clear from Fig. 5.16 that to capture 9P content and above, 0.5 steps are required (at least near the tower; this could be relaxed away from the tower). The same accuracy in not evidently required for the overall SPL, as shown in Table 5.2. A 0.5 step is used in the remainder of the results. Including the tilt (5 ) moves the blades away from the tower, reducing the SPL by 9 dB. The SPL still appears quite high, but bear in mind the observer is only 100 m away, and as shown in Fig. 5.17, the assumed pressure around the tower causes quite large velocity uctuations even for this upwind case. With wind shear, and then cd = 0 for the tower, the SPL is reduced further. Figure 5.18 illustrates the observer pressure signature, which typically exhibits peaks at the 3P frequency. It also shows the peak reduction by tilt. Adding wind shear has fairly minimal eect on either spectral content or SPL. With the rotor downwind and taking lw = 3.25, = .3 and ww = 2.5 as typical for a cylindrical tower [87], the SPL is 88.4 dB without tilt and 85.3 dB with tilt. Figure 5.19 shows the energy content is shifted to somewhat higher frequency (audibility) compared to the upwind case. These are only mildly higher than the data presented in the other quoted studies. It is likely that the current test rotor is

5.3 Structural Validation


100

169

SPL (dB)

75 50 25 0 0 2 4 6 8 10 12 14 16 18 2 1 0.5 20

f (Hz)
100

SPL (dB)

75 50 25 0 1.6 3.1 6.3 12.5 25

f (Hz)

Figure 5.16 SPL for upwind rotor with tower inuence for varying azimuthal step size (deg) Table 5.2 Upwind SPL predictions Case Isolated rotor Tower inuence 2 step 1 step 0.5 step Tilt Tilt and wind shear Tilt and wind shear and cd = 0 SPL (dB) 17.3 77.4 81.7 83.1 73.9 73.0 72.6

proportionally closer to the tower. The acoustic footprint of the rotor is given in Fig. 5.20. The shift in the crossstream direction is consistent with the comparison data set [132] and is caused by the rotor advancing/retreating from the observer. The lobes are caused by the dominant forces being aligned with the free-stream. No account is made of acoustic reectivity from either the ground or tower, which would further modify the footprint.

5.3 Structural Validation


The cross-sectional model is validated rst in 5.3.1, followed by the FEM beam model in 5.3.2.

170

Chapter 5 Validation

11

11 10.5

Vop (m/s)

10 9 8 7

120 100 40 80 60 40 -40 -20 20 0

Vop (m/s)

10 9.5

9 120 100 80 20 60 40 -40 0 -20 40

(a) No tilt

(b) With 5 tilt

Figure 5.17 Eect of tilt on section out-of-plane velocity for upwind rotor (global coordinate system)

1 thk near far total

0.4 0.2 0 thk near far total

0.5

p (Pa)

p (Pa)
0 2 4 6

-0.2 -0.4 -0.6

-0.5

-1 -0.8 -1.5 -1 0 2 4 6

tObs (s)
(a) No tilt

tObs (s)
(b) With 5 tilt

Figure 5.18 Eect of tilt on pressure signature for upwind rotor

5.3 Structural Validation


100 80

171

No tilt SPL (dB)


60 40 20 0

Tilt

10

12

14

16

18

20

f (Hz)

Figure 5.19 Eect of tilt on downwind rotor SPL


105 90 75

120 135 150 165 90 180 -165 -150

60 45 30 15

80

70

0 -15 -30

-135 -45 -120 -60 -105 -90 -75

Observer azimuth (0 downwind)

Figure 5.20 Acoustic footprint (dB) of downwind rotor with tilt and shear

5.3.1 Sectional Properties


The sectional computations were veried against analytic formulas for a thin ring, box section and I-beam, for all stiness and mass properties. The analytic formulae were obtained and derived from the standard analysis methods (parallel-axis theorem, etc.) presented by Ugural and Fenster [151]. For the circle, renement of the element size (arc-length) showed asymptotic convergence to the analytic value (below 0.2% at for 100 chordwise sections). The other shapes matched identically.

172

Chapter 5 Validation

5.3.2 Beam Model


The beam model has been checked against BLADEDTM s modal predictions, for all but the apping boundary condition, which is not part of BLADEDTM s capability. Four test cases are considered: uniform (unity) mass, stiness, and exible blade length L = 1 (5.3.2.1); linearly varying mass and stiness (both 1.10.1, root tip) with L = 1 (5.3.2.2); GH Demo blade (5.3.2.3); and the CONE-450 blade (5.3.2.4). Two and three bladed rotors were checked, as were clamped and freerotation for the in-plane modes. In the gures to follow, the FEM results are given by dots, those from BLADEDTM by crosses. The modal frequencies n are given as e.g. 4 Hz (5 Hz) for FEM (BLADEDTM ) results. Where it is stated that the two sets of results are same, what is meant is that the modal basis vectors are equivalent. BLADEDTM solves for modes that typically have motion of all blades simultaneously. The FEM method typically produced modes in which one blade is displaced and the others have no displacements (for clamped root conditions). When the frequencies are identical, the modal basis vectors describe the same eigenspace. 5.3.2.1 Uniform Blade For reference, the OP and IP modes for a rotor with three uniform blades, with freely rotating hub, are given in Fig. 5.21 and Fig. 5.22. The results have no ap hinge and lhub = 0,1 and the rotation speed is 100 RPM. As expected, the OP modes show the blades completely un-coupled. The mode shapes and frequencies are identical to those from BLADEDTM . Figure 5.23 shows the modes with a teetered hub.2 The rigid-body modes associated with teeter are clearly present. The static and rotating mode shapes are identical in all cases, and the rotating modes are higher frequency, as expected. With apping hinges, the FEM results agree with Eq. (4.6.24). The IP modes however showed some discrepancies. For two (three) bladed rotors with freely rotating hub, every second (third) mode diered in frequency and shape, as shown in Fig. 5.24. The BLADEDTM results for this test case deviate from the
FEM results for the collective IP mode shapes. Double curvature, as evident in
1 Note that lhub = Rhinge in the vernacular of 4.6.4 for an un-hinged blade, representing the oset of the blade root from the rotation axis. 2 The dierences between the two sub gures owe to the coordinate denitions; BLADEDTM uses a positive displacement for upwind motion of all blades, whereas the FEM formulation couples the rotation of the hub node in teeter, so that positive displacement is upwind for one blade and downwind for the other.

5.3 Structural Validation

173

#1: 1.86 Hz 1 0.5 0 -0.5 -1 0

0.2 0.4 0.6 0.8 1

#2: 1.86 Hz 1 0.5 0 -0.5 -1 0 0.2 0.4 0.6 0.8 1

#3: 1.86 Hz 1 0.5 0 -0.5 -1 0 0.2 0.4 0.6 0.8 1

r
#4: 5.50 Hz 1 0.5 0 -0.5 -1 0

r
#5: 5.50 Hz 1 0.5 0 -0.5 -1 0 0.2 0.4 0.6 0.8 1

r
#6: 5.50 Hz 1 0.5 0 -0.5 -1 0 0.2 0.4 0.6 0.8 1

0.2 0.4 0.6 0.8 1

r
#7: 12.06 Hz 1 0.5 0 -0.5 -1 0

r
#8: 12.06 Hz 1 0.5 0 -0.5 -1 0 0.2 0.4 0.6 0.8 1

r
#9: 12.06 Hz 1 0.5 0 -0.5 -1 0 0.2 0.4 0.6 0.8 1

0.2 0.4 0.6 0.8 1

Figure 5.21 Uniform blade OP rotating modes

the FEM results, must exist for these modes, owing to the free hub.1 As shown in Fig. 5.22, there should be sets of B 1 reactionless modes and a single collective mode. Increasing lhub to 0.5, the OP modes still match. The IP modes shown in Fig. 5.25 again exhibit the slope error just discussed. BLADEDTM gives a zero displacement at the root. As the hub is free to rotate, the oset should yield some displacement at the root (root = hub lhub for small angles). Further investigation, by increasing v lhub to 49, gave the results in Table 5.3. The mode shapes are not shown, however the BLADEDTM results were found to have some minimal displacement at the root for both OP and IP modes. The IP modes were also again dierent for the collective modes. The discrepancies in the BLADEDTM IP results seem to indicate some nite stiness is associated with the hub. The internal details of the BLADEDTM method were not available, so the cause could not be denitively determined. As will be
This may also be justied by drawing the BLADEDTM modes as a rotor. It is then obvious that a moment would have to be present at the hub for the deections predicted by BLADEDTM to occur.
1

174

Chapter 5 Validation

#1: 1.86 Hz 1 0.5 0 -0.5 -1 0

0.2 0.4 0.6 0.8 1

#2: 1.86 Hz 1 0.5 0 -0.5 -1 0 0.2 0.4 0.6 0.8 1

#3: 4.85 Hz 1 0.5 0 -0.5 -1 0 0.2 0.4 0.6 0.8 1

r
#4: 5.50 Hz 1 0.5 0 -0.5 -1 0

r
#5: 5.50 Hz 1 0.5 0 -0.5 -1 0 0.2 0.4 0.6 0.8 1

r
#6: 10.59 Hz 1 0.5 0 -0.5 -1 0 0.2 0.4 0.6 0.8 1

0.2 0.4 0.6 0.8 1

r
#7: 12.06 Hz 1 0.5 0 -0.5 -1 0

r
#8: 12.06 Hz 1 0.5 0 -0.5 -1 0 0.2 0.4 0.6 0.8 1

r
#9: 19.35 Hz 1 0.5 0 -0.5 -1 0 0.2 0.4 0.6 0.8 1

0.2 0.4 0.6 0.8 1

Figure 5.22 Uniform blade IP rotating modes

1 0.8

B1:1 Normalized displacement B1:3

0.8 0.6 0.4 0.2 0 -0.2 -0.4 -0.6 -0.8

Normalized displacement

0.6 0.4 0.2 0 -0.2 -0.4 -0.6 -0.8 -1

B1,2:1 B1:2

B1,2:2

B2:3 B2:1

B1,2:3
0.2 0.4 0.6

B2:2
0.8 1

0.2

0.4

0.6

0.8

-1

r/R
(a) BLADEDTM

r/R
(b) FEM

Figure 5.23 Uniform blade OP rotating modes for teetered hub (Blade number:mode number)

5.3 Structural Validation


-0.8 1 1 0.8

175

Normalized displacement

0.8 0.6 0.4 0.2 0 -0.2 -0.4 -0.6 0 0.2 0.4 0.6 0.8 1

Normalized displacement

0.6 0.4 0.2 0 -0.2 -0.4 -0.6 -0.8 0 0.2 0.4 0.6 0.8 1

r/R
(a) Mode 3, 4.85 Hz (1.86 Hz)

r/R
(b) Mode 6, 10.59 Hz (5.50 Hz)

Figure 5.24 Free-hub rotating uniform blade IP modes, FEM (BLADEDTM )

-0.6 1

-0.8 1

Normalized displacement

Normalized displacement
0.2 0.4 0.6 0.8 1 1.2 1.4

0.8 0.6 0.4 0.2 0 -0.2 -0.4 0

0.8 0.6 0.4 0.2 0 -0.2 -0.4 -0.6 0 0.2 0.4 0.6 0.8 1 1.2 1.4

r/R
(a) Mode 3, 4.01 Hz (2.35 Hz)

r/R
(b) Mode 6, 8.94 Hz (6.47 Hz)

Figure 5.25 lhub = 0.5 free-hub rotating uniform blade with IP modes, FEM (BLADEDTM )

176

Chapter 5 Validation Table 5.3 Uniform blade with long hub eigenfrequencies (lhub = 49) Mode #
OP

Static (Hz) BLADEDTM FEM 0.56 0.56 0.56 0.56 0.56 0.57 0.56 0.56 0.56 0.56 0.56 0.90

Rotating (Hz) BLADEDTM FEM 2.64 2.64 2.64 2.64 2.64 9.71 14.24 14.24 14.24 14.24 14.24 22.58

IP

1 2 3 1 2 3

shown for realistic blade properties in 5.3.2.3 and 5.3.2.4, the BLADEDTM results do not exhibit this error. The present test case is therefore a pathological example that with very small blade mass, stiness and length drastically alters the usual hubblade stiness relationship. The BLADEDTM method is a pragmatic one, and is valid for real blades. The OP frequencies do match in all cases, which is curious, considering the nite deection at the root for the BLADEDTM results. 5.3.2.2 TaperedBlade The tapered blade was compared for both lhub = 0 and 0.5. Overall, the results matched BLADEDTM , with the same discrepancies discussed for the uniform blade. 5.3.2.3 DEMO Blade The inclusion of generally varying stiness, mass and twist, as well as collective pitch and modest oset, tests all facets of the FEM method. Comparison with BLADEDTM results at 18 RPM, for both a rigid 3-bladed rotor and teetered 2bladed rotor, showed virtually identical results. There were small dierences in n for higher modes (>= 6 for 3 bladed rotor), presumably owing to the dierent numerical approaches. Figure 5.26 shows the third, sixth and ninth collective IP modes, and Table 5.4 list the associated n . In contrast to the uniform and tapered blade, for a real geometry the BLADEDTM and FEM results are in very close agreement for the collective modes. It is likely some stiness value is used for the hub in BLADEDTM , producing odd results only for unrealistic pathological test cases.

5.4 BLADEDTM Validation and Suitability


1 1

177

Normalized displacement

0.6 0.4 0.2 0 -0.2

Normalized displacement

0.8

0.8 0.6 0.4

6 9

6
0.2

9
0 -0.2

3
-0.4 0 10 20 30 40 -0.4 0 10

3
20 30 40

r (m)
(a) BLADEDTM

r (m)
(b) FEM

Figure 5.26 Collective IP mode shapes for Demo blade Table 5.4 Collective IP eigenfrequencies for Demo blade Mode # 3 6 9 Static (Hz) BLADEDTM FEM 3.10 8.32 16.52 3.11 8.37 16.67 Rotating (Hz) BLADEDTM FEM 3.17 8.38 16.57 3.17 8.43 16.72

5.3.2.4 CONE-450 Blade The IP modes again match as per the DEMO blade. No BLADEDTM solution is available for the OP modes, due to the ap hinge. The integrated blade properties for this blade enable analytic computation of the rst three ap modes (rigid body) from Eq. (4.6.24). Testing at = 25 RPM, and with the properties in C.5, Eq. (4.6.24) predicts 0.456 Hz, which is exactly equal to the rst three frequencies from the FEM code, corresponding the three independent rigid-body ap modes.

5.4 BLADEDTM Validation and Suitability


BLADEDTM is commercial software, with full Germanischer Lloyd certication and a development period of 20 years. To achieve that status, it has been validated against numerous machines and used successfully in many projects. Throughout this thesis, the code has been stretched to allow analysis of the coning rotor. It should be noted that for the original CONE-450 work, Jamieson [66] used a custom version of

178

Chapter 5 Validation

the code, diering from the current version (v3.65) primarily in parked aerodynamics and the structural model. This section deals with the modications of the current version to handle the coning rotor concept. The areas of discrepancy found were aerodynamics (5.4.1), steady-state solution including coning and VSS (5.4.2), and the dynamic structural model (5.4.3). To reiterate, the issues discussed in this section pertain to the applicability of BLADEDTM to the coning rotor, and are not meant to indicate unsuitability for conventional machines.

5.4.1 Aerodynamics
Eorts towards implementing the aerodynamic renements developed in 4.4 in BLADEDTM have been made. Unfortunately, to date the work is incomplete, despite the best eorts of all parties. In addition to implementing the induction-factor corrections, it was discovered well into the project that a number of other aerodynamic modelling features will be required: Smooth blending of tower shadow at tower top (see 4.4.14.3) Some spanwise ow model should be included (see 4.4.9) A condition-dependent stall-delay model is required (see 4.4.8) The issues in the following sections further added to this list of outstanding issues, and precluded the development of a full enhanced implementation within BLADEDTM in the available time-scale.

5.4.2 Steady State Operation


For the purposes of design optimization (Part III), steady-state operation points are required. Unfortunately, no direct coning equilibrium calculation is available within BLADEDTM , akin to that developed in 4.7.1.1. Likewise, optimal operation calculations, as presented in 4.7.1.2, are not a feature of BLADEDTM . It would of course be possible to bypass these by utilizing the dynamic calculations with a self-tuning controller, but at increased computational eort. It was also found that the steady-power curve calculation in BLADEDTM , while able to iterate on pitch angle or rotor speed for limiting power 4.7.1.2, in some cases produced non-smooth power curves.1 An example is shown in Fig. 5.27 for
PTS operation of the REF-1500. This behaviour would have proven problematic for

the optimization studies which require accurate estimates for changes in objective
Interestingly the rated electrical power is constant, however the shaft (aerodynamic) power is not. The latter is presumably the determinant of the pitch/speed schedule.
1

5.4 BLADEDTM Validation and Suitability


1.8 1.6 1.4 0 -1 -3 -4 -5 -6 0.0 5.0 Shaft power 10.0 15.0 20.0 Pitch angle -7 25.0 -2

179

Power (MW)

1.2 1.0 0.8 0.6 0.4 0.2 0.0

Windspeed (m/s)

Electrical power

Figure 5.27 BLADEDTM computed steady power curve for PTS

function. Based on the authors experience with ExcelBEM, solving accurately for power limiting with either PTS or VSS requires extremely tight termination criteria for the iteration algorithm. This owes to the steep derivatives of power/torque with control variable, when the rotor is in stall. Unfortunately, the solution tolerances for this calculation in BLADEDTM are not accessible to the user.

5.4.3 Flapping Hinge


BLADEDTM was modied to feed ap angles j and ap velocities j to an externally implemented controller. A switch was also added to turn on/o the built-in ap restraints pre-existing in the code. The restraints are modelled as individual hydraulic cylinders attached at the pin-end to a point oset from the rotor axis, and the rod-end to a pin on the blade outboard of the hinge axis.1 The physical geometry and hydraulic parameters (including end-stops and accumulator pressure) are used to compute an equivalent hinge-moment that is applied to the ap hinge in the simulation. For generality, a facility was added for the current project to allow the external controller to impose an arbitrary hinge moment. This allows modelling of a generalized actuator, either passive or with active control. To model a apping hinge in BLADEDTM , the mode shapes and natural frequencies must be provided from the FEM predictions developed in 4.6.4. To exclude
1

This modelling was added to analyse the WTC prototype.

Pitch angle (deg)

180

Chapter 5 Validation

any complicating structural eects, the uniform blade from 5.3.2.1 was used with Rhinge = 0. Only the rst ap mode was input to BLADEDTM , known to be rigid body rotation about the hinge. Testing of the structural simulation, in isolation from the aerodynamics, was possible by dening airfoils with zero lift and drag. Gravity was also turned o with g = 0. A simple external controller was dened to apply a constant moment plus some damping to each blade. Analytic solutions are available from Eq. (4.6.20b) (non-linear) and Eq. (4.6.21) (linear) by setting j = j = 0.1 Using this approach, the blade was modelling in ExcelBEM using 5 sections, in both Steady and Dynamic modes, the later with critical damping applied. The results were compared to BLADEDTM models with 5, 25 and 100 equally-spaced sections. The results are given in Table 5.5, which yield the follow insights: The analytic results show that assuming linearity articially stiens the rotor, becoming important around 10 . ExcelBEM solutions all match the non-linear analytic solution. The slight numerical discrepancy in the Steady results owes to the accuracy of the algorithm in 4.7.1.1.2 BLADEDTM results follow quite closely the analytical linear predictions. The BLADEDTM predictions do not converge to either the linear or non-linear analytic solutions with increasing numbers of sections. The last point suggests another eect, other than just a linearity assumption, is present in BLADEDTM . Based on the operational prole of the coning rotor (up to = 40 ), the virtual ap stiness present in the BLADEDTM model means that equilibrium may be in error by approximately 30%. The power, thrust and other predictions will also be aected by approximately the cosine of this error level. The source of the discrepancy is evident from examination of the BLADEDTM theory manual [125]. As discussed in 4.6.3, BLADEDTM simplies the EOM by assuming a known modal frequency, even though the modes used are not normal. In the case of gross-coning, the modal frequencies are altered (as the centrifugal stiening changes with ), and so the embedded EOM adhere to a linearising assumption that is increasingly in error with . The original work [66] appears to have used only mode shapes, not frequencies, so those predictions should be accurate in coning.

1 From the solution to these equations is also apparent that two solutions for j are possible for a given Hact , either side of 45 . 2 The solution is obtained much faster with the Steady method however, which is why it is preferred during optimization over running the dynamic simulation to steady-state.

5.4 BLADEDTM Validation and Suitability

181

Table 5.5 Steady ap angle variation with actuator moment for =50 RPM (deg) Hact (Nm) 200 300 350 13.75 14.34 14.30 14.34 13.33 13.74 13.48 20.63 23.03 23.01 23.03 20.00 20.61 20.22 24.06 28.57 28.57 28.57 23.34 24.04 23.60

Model Analytic linear Analytic non-linear ExcelBEM Steady ExcelBEM Dynamic BLADEDTM 5 BLADEDTM 25 BLADEDTM 100

100 6.88 6.94 6.88 6.94 6.67 6.87 6.74

400 27.50 36.87 36.85 36.87 26.67 27.48 26.97

5.4.4 Mode Shape Modification by Coning


The pre-stress present in the blade owing to centrifugal loading will vary with cone angle, approximately as cos2 . In general, the aerodynamic and inertial forces will not precisely align with the axis of the blade, but as assumed in the derivation of 4.6.4.1, some insight is gained by ignoring these eects. The analytic frequency of the free-body apping mode may be computed with the non-linear form of Eq. (4.6.24), linearized about 0 : n = cos(20 ) + Rh Mb sCG cos(0 ) I (5.4.1)

Here again, n is imaginary for 0 > 45 (for Rh = 0), indicating instability. The FEM results do not agree with Eq. (5.4.1) for the fundamental ap mode, predicting higher natural frequencies as 0 is increased. The higher OP and IP modes are almost identical in shape and frequency to the unconed case. The reason for the discrepancy is the dierence in boundary conditions between the analytic and FEM equations. Equation (5.4.1) assumes a hinge moment is applied to create the equilibrium condition, whereas the FEM method assumes forces applied along the blade exactly balance the transverse centrifugal force component. To harmonize the two approaches would require including the equilibrium bending moment stress in Eq. (4.6.45), rather than assuming plain stress. This extension is left as future work, as the mode shapes are relatively unaected by the assumption of 0 = 0. Moreover, as just discussed in 5.4.3, modal frequencies cannot be used directly in the dynamic simulation of the coned rotor.

182

Chapter 5 Validation

5.5 Generator Model Validation


Dubois [165] presented 14 optimized design variables (p , J, Bg , bt /bs and Krad ) for a range of parameters (diameters D and eciency ). The Solver built into Excel has been used to conduct multi-parameter optimization of the spreadsheet model with the same parameters and constraints as the original thesis. Comparison of the original results with the current ones indicate that the model is consistently implemented.

Part

III

Design

Chapter

Rotor Optimization
Part II focused on developing tools capable of analysing the coning rotor. In addition to analytical dierences, compared to a conventional machine the operational concept of the coning rotor demands a somewhat dierent and more integrated design approach. This chapter rst uses the results of the original CONE-450 study to predict performance relative to modern machines, in terms of load (6.1) and energy capture (6.2). The generic aerodynamic behaviour of the coning rotor is examined next in 6.3. Section 6.4 then elucidates the challenges inherent in optimizing the coning rotor. Finally, a parametric study of the DVs of the coning rotor is presented in 6.5.

6.1 Initial Comparison and Updating


The present section rst contextualizes the original coning rotor work (6.1.1), before properly scaling up the loading results from that study in 6.1.2. Justication is then given in 6.1.3 for leaving the revisiting of parked conditions to future work.

6.1.1 Shifting Benchmarks


In the mid-1990s, when the original coning rotor work was done, the benchmark comparison machine was a 450 kW, constant-speed, FSS regulated machine (REF450) [7]. Technical evolution means that an appropriate benchmark is now a 1.5 MW variable-speed pitch regulated machine.1 This evolution has been by incremental renements in design and analysis within the wind turbine community, spurred on by cost and performance enhancements of actuators, materials, and power electronics. The two main shifts relevant to the present comparison have been to substantial variable speed operation and increasingly fast, individual blade pitch control.
1.5 MW was chosen as generic machine data was available for this size and is around the average capacity actually installed on land, notwithstanding the industry drive to capacities exceeding 4 MW for oshore sites where economies of size are still unclear.
1

185

186

Chapter 6 Rotor Optimization

6.1.2 Up-Scaling CONE-450 Results


As an initial scoping exercise, the representative operating curves from the original 450 kW machine (CONE-450) and REF-450 dynamic simulations [36] were scaled up to a 1.5 MW machine (CONE-450 US and REF-450 US). The loads shown in Fig. 6.1 were obtained by power scaling based on the reference rotor diameter Dref , in turn scaled proportionally to the square root of the rated power. These are compared against the loads computed for a generic 1.5 MW pitch controlled machine (REF-1500, see C.3) using BLADEDTM .
250

200

Thrust (kN)

150

100

50

REF-1500 CONE-450 US REF-450 US 0 5 10 15 20 25

Windspeed (m/s)

Figure 6.1 Operational rotor thrust curves for upscaled machines relative to REF-1500 The results indicate that the overall loading for the coning concept at a larger scale is reasonable, and highlights the dierences between the pitch control and active stall/coning strategies. The design of the original CONE-450 (see 6.4.3) was based on maintaining the same maximum thrust level as the REF-450, occurring for the CONE-450 at rated and for the REF-450 at cut-out (25 m/s). The CONE450 US attains the same maximum thrust, but stays at a relatively high thrust after rated relative to the REF-1500, as stalling imparts more force than pitching to feather. It can be expected that the parked extreme loading case, discussed further in 6.1.3, remains a non design-driving condition for the upscaled coning rotor. Fatigue loading comparisons must be deferred to more detailed time-domain analysis in the future.

6.1 Initial Comparison and Updating

187

The original energy yield results included not only a larger rotor for the CONE450, but also an increase in tower height from 35 m to 50 m. The tower height increase was enabled primarily by equal tower bottom overturning moment when parked (i.e. coning rotor thrust less severe in parked extreme conditions). The energy yield advantages, relative to other conventional machines (250600 kW), varied from -1564%, depending on machine rating and mean wind speed. The CONE-450 was superior in low winds (5-6 m/s mean), but lost this advantage in high winds (910 m/s), paricularily for machines of equal tower height and higher rating. The net competitive advantage (in terms of COE) was positive in all cases, ranging from 0.262%. It was noted that the CONE-450 was designed for true Class I conditions, whereas the reference machines were not, and so the CONE-450 would in fact be more cost competitive than indicated by the gures. Against the reference machine in normal UK conditions, the CONE-450 had an 18% Net Present Value (NPV) advantage without taller tower.

6.1.3 Parked Conditions


A novel aspect of the coning rotor is the parking strategy (fully coned to 85 ). Jamieson [66] examined this condition in some detail, including fully orthotropic turbulence and various delta-wing stall delay models (see 4.4.9.2).1 The results indicated that parking did not drive the design of the CONE-450. Based on this, the present work does not explore in detail this aspect of the coning rotor. A primary concern might be the overhang weight associated with a larger scale machine. For the blades discussed later in 6.5 (CONE-1500), deterministic steady blade root bending moments are 500 kNm owing to gravity. The blade structure design loads2 are 1500 kNm, indicating a large static factor of safety. The REF1500 experiences a maximum of 2107 kNm overturning moment at the yaw bearing. The 3 500 = 1500 kNm equivalent moment for the CONE-1500 is therefore not unreasonable (the extreme moment from the CONE-450 US is 913 kNm when parked, versus 3512 kNm in extreme operation). Further validation of the 1.5 MW scale concept must revisit this issue to verify the assumption that parked loads do not dominate. The models must include full turbulence, and ideally wind-tunnel testing to verify aerodynamic loading predictions for the fully-coned blades experiencing extensive spanwise ow.
1 2

The IEC standard of the day only specied steady winds for the parked case. Obtained as the root moments to hold the blades open at rated conditions.

188

Chapter 6 Rotor Optimization

6.2 COE and CF Performance


The interrelation between COE and CF, and their use as metrics in wind turbine optimization were introduced in 3.2. Figure 6.2 shows the ideal performance of a variable area rotor, able to operate at CP = 0.48 with full diameter up to rated, and 1.5 MW above. Performance is shown over a range of diameters (70100 m) for a nominal site with k = 2.2 in Fig. 6.2(a). Three mean wind speeds are compared, based on a realistic cost exponent with diameter (see 3.5) of 2.4 [15], relative to the baseline 70 m rotor diameter. Based on this simple analysis, COE rises quite quickly with CF, but less so for low-wind sites.
1.8 1.7 1.6 1.5 V=6 V=8 V=10
0.50 0.45 0.40 k=1.7 k=2.2 k=2.5 E 6000

1.3 1.2 1.1 1.0 0.2 0.3 0.4 0.5 0.6 0.7 0.8

CF

1.4

0.35 0.30 0.25 0.20 70 80

4000 CF

3000

2000

CF

D (m)

90

100

(a) CF dependency (k = 2.2)

(b) k dependency (Vmean =6 m/s)

Figure 6.2 Betz limit variable area rotor performance The eect of k is shown in Fig. 6.2(b), across the range of diameters. As expected,
CF and energy E both rise with diameter, while the k eect is rather subtle. Low k

factors (broad f curve) reduce both dependencies, and visa versa. The design choice of rating and rotor size varies between designers. Indeed, multiple rotors of varying diameter may be oered for the same nominal machine (i.e. generator, tower, etc.), based on operation in dierent wind classes (see 2.1.3). To examine the impact of these design decisions, a number of power curves for real 1.22 MW machines were obtained from product literature. Performance is compared for k = 2.29 and Vmean = 8.24 m/s, based on a typical US site [15], in Fig. 6.3. Tower height is assumed constant. Figure 6.3(a) shows the
PDFs of energy capture for the machines relative to the wind speed prole, ideal

Energy Yield (MWh)

5000

COE

6.2 COE and CF Performance

189

Betz limit (over entire wind speed range), and a scaled power curve for the CONE450 US. The ideal capture is clearly weighted to higher V than both the wind and actual machine capture. The machine proles are clustered quite closely around 11 m/s, with the CONE-450 US attaining the highest peak (tightest distribution).
0.14 0.12 0.10 0.08 0.06 0.04 0.02 0.00 0 Wind Bonus 2MW 0.50 5 10 Ideal NM 64-1500 15 20 25 Vensys 62 CONE-450 US 0.32 0.30

fV, fEann (s/m)

U (m/s)

V80 NM 72-2000

(a) Wind speed prole


6.5 6.0 0.45

D2.4/E (m2.4/MWh)

5.5 5.0 4.5 4.0 0.30 VENSYS 0.35 0.40 D/E P/E

CF

0.26 0.24 0.22 0.20 0.50

0.40 stall control 0.35 0.25 0.30 0.35 0.40 0.45 0.50

P/A (kW/m2)

CF

0.45

(b) Power density

(c) Cost metrics

Figure 6.3 Industry machine performance

Figure 6.3(b) shows that CF is in fact a linear function of the ratio of generator power to rotor area, as expected. The one PTS machine deviates somewhat from the overall trend, with lower CF. List prices are unavailable, but Fig. 6.3(c) compares

P/E (kW/MWh)

0.28

190

Chapter 6 Rotor Optimization

two proxy COE metrics: a cost function D2.4 /E assuming cost to vary with diameter (see 3.5), and the ratio of installed power to energy captured P/E. The former metric results show costs rising with CF, mirroring earlier results in 3.2. The Vensys machine evidently achieves a superior power curve relative to the conventional machines, while the CONE-450 US would have a high cost if the D2.4 relationship is not altered for the coning rotor. The P/E ratio shows that installed generator power relative to rotor size is an almost linear function with CF. This mirrors the
CFP/A relationship and suggests capture area A is a dominant factor.

6.3 Aerodynamic Behaviour


It is instructive to examine the non-dimensionalized aerodynamic behaviour of a coning rotor, before considering its optimization. Figure 6.4 shows the performance maps, non-dimensionalized by real rotor area, of the REF-1500 over a range of and . The optimal CP location moves in with , and in this case CP,max rises with . These trends are generally present in other rotors examined, however the direction of shift varies. Thrust behaviour is also seen to change with .

0.5 0.4
0.5

1 0.8
0.8

0.3

CP

0.4

0.6

CT

0.2

0.3
0.1 0 0 20 40 0 60 2
0.2

0.6
0.4 0.2 0

0.4

0.1
4 6 8 10

0.2
0 10 20 30 6 8 10

(deg)

(deg) 40 50

(a) CP

(b) CT

Figure 6.4 Performance maps for actual tip radius Figure 6.5 shows a set of curves for the REF-1500 non-dimensionalized by the reference tip radius, for a range of cone and pitch angles. The intent is to illustrate the extent of control possible based solely on coning versus a pitch-controlled rotor (ne or stalling pitch). Power control is clearly more direct with pitch control, both

6.4 Rotor Optimization Challenge

191

in terms of CP reduction with angle and the ability to change pitch directly instead of via an indirect speed-stall-cone relationship. The extremely fast decrease in CP for PTS is also evident.
0.5 0.5

0
0 0.4 0.4 5 -5 10 0.2 15 -10

0.3

0.3

40 45
0.2

CP

50

CP
0.1 0

0.1

0 2 4 6 8 10

10

(a) Coning angle ( = 0 )

(b) Pitch angle ( = 0 )

Figure 6.5 Performance maps for reference tip radius and control variable ( or pitch in 5 increments)

6.4 Rotor Optimization Challenge


Attention now shifts to rening the coning rotor design by considering a viable optimization process. The procedure for a conventional rotors is outlined rst in 6.4.1, followed by the challenges introduced by the coning rotor in 6.4.2 and the approach to the problem taken by previous investigators in 6.4.3.

6.4.1 Conventional Rotors


The modern variable-speed rotor is typically rst optimized aerodynamically for a specic tip speed ratio opt [51]. The optimal operational speed prole is unambiguously dened in both the rotor speedwind speed (U ) and torquerotor speed planes, and can be relatively easily computed from a knowledge of the nondimensional power curve and the equations in Fig. 3.7(b). Treating airfoil choice (both shape and percent thickness) as parameters in the optimization, the equations in 4.4.13 prescribe the chord and twist DV distributions for the optimal blade. The overall scale of the machine will modify the results

192

Chapter 6 Rotor Optimization

somewhat, owing to Re variation, but otherwise the design is non-dimensional. The blade length is guided by experience and an estimate of CP in Eq. (3.2.2) to yield the rated power desired for the machine (inuenced by the results of 6.2). After this rst-pass, the blade must be further rened in a number of ways. The most apparent is the requirement to smooth the chord and twist proles. Airfoil data is not smoothly varying between thicknesses, so Eq. (4.4.36) will not yield smooth proles. Additionally, constraints are imposed on the geometry by functional (a circular root for the pitch bearing) and manufacturing (maximum twist angle) requirements. These changes may be incorporated by simply smoothing the results of the analytic equations, or by numerically optimizing the proles directly. FSS rotors, with constantly varying , must be optimized numerically across the wind speed range. Dynamic load simulations are required to certify a design, and also to optimize an internal structure. Paradoxically, those loads are unavailable without a dened structure, as the mass and stiness properties are required for the simulations. This is usually handled by designing for some simplied, steady load cases. Malcolm and Hansen [51], like others [180], uses extreme bending loads to size the sparcap thickness at 4 spanwise locations, with interpolation for intermediate sections. Grin [55] does likewise for the ap bending loads (yielding spar cap thickness), with combined maximum torque and gravity fatigue loading for edgewise loads (to size a
TE spline at 0.95c). Tip deection in a 50 year extreme gust is also checked to limit

minimum thickness. Some type of numerical optimization may be easily used at this stage, either as interpolation functions for spar cap thickness as in the referenced studies, or direct numerical optimization. The dynamic simulations are then run, the real limit and fatigue loads computed, and the resulting strains checked. Some manual renement of thicknesses, structural concept, and/or materials is then made to satisfy the material limitations. Aerodynamic optimality is usually the highest priority for the blade design, given that the rotor is only 1015% of the total cost, while improvment of energy capture more directly impacts COE. However, the interplay between aerodynamic and structural considerations in fact denes a pareto front, albeit heavily weighted towards the aerodynamics (CP ). The main competing factors are: High performance airfoils (high lift and lift/drag ratio) yield small chords, adversely reducing section thickness and aecting structural properties [181]. FSS rotors require limited lift in the tip region to control power and ensure

6.4 Rotor Optimization Challenge

193

smooth stall behaviour to alleviate edgewise vibrations (see 7.1.2.2) [65]. Gigu`re et al. [65] concluded in an optimization study that COE was best e for higher-lift airfoils, as material use was reduced by limiting the 50 year gust load with smaller chords. These are obviously conicting requirements. Small airfoil percent thicknesses limit both the structural area and height in a section. Higher strength materials (e.g. carbon bre) can accommodate these constraints, with associated cost. Lower modulus materials (e.g. wood or glass/polypropylene) struggle to t enough material into the section, with reduced structural eciency as the thickness grows towards the elastic axis. The airfoil shape itself inuences the section stiness. Boxier sections place more material away from the elastic axis providing ecient use of material, but possibly to the detriment of aerodynamic properties. The aerodynamic optimum typically waists the chord around mid-span, a critical area for fatigue requiring good structural properties (i.e. larger chords and thicknesses). These eects may be incorporated into a coupled structural-aerodynamic optimization, but more typically are manually iterated.

6.4.2 Coning Rotor Optimization


Unfortunately, the inherent complexity of the coning rotor is manifest when optimization is attempted. The coning rotor inherently ties to via the rotor inertia and aerodynamic hinge moment, so that below rated the optimal operation is not simply dened by opt . The aerodynamic and structural designs are more tightly coupled for the coning rotor, requiring more integration of the optimization process. In addition to the blade prole DVs discussed in 6.4.1, some inclusion of operational behaviour must be included either explicitly or implicitly. One approach is to track the locus of CP,max over , following the ridge of the 3D CP surface in Fig. 6.4 (of course the surface must be computed for each design iteration). This can be used to dene a map in the below rated condition, with the same speed limits and above rated operation as for the non-coning rotor. The performance of the rotor can then be determined over a range of wind speeds in an iterative manner. This approach has a number of drawbacks, including: diculty modifying the curve directly to satisfy constraints (e.g. tip-speed for noise) since the control curve is implicitly dened via the 3D performance map; optimality is not assured, as coned diameter also aects power, not just a max CP condition; and many iterations are

194

Chapter 6 Rotor Optimization

required for both denition of the control curve and solution for operating conditions. Additionally, although a simple non-dimensional inertia is readily derivable for a centrally-hinged rotor, the inclusion of the hinge oset complicates the denition (requiring some relationship of mass and rst/second moments of inertia). Above rated, another optimization exercise is required in the non-dimensional space to dene a constant power curve. For a PTS rotor the space grows to four dimensions with the inclusion of pitch. Returning to a dimensional approach, real control strategies are posed in the plane, as wind speed cannot be directly measured. For the purposes of steadystate optimization, control may be dened in the -U plane, to avoid the required iterations to nd steady-state torque and speed.1 A naive approach is therefore to treat the entire curve as a function of a set of design variables, from cut-in to cut-out wind speed. This overly complicates the problem, which is better dened by limits on rotor (and/or tip) speed, some DVs dening the nominally rated portion of the power curve, and iteration on rotor speed (or pitch angle) to limit power to rated (see 4.7.1.2). The optimizer can then control the below-rated schedule, with implicit power control above rated. For a non-coning rotor, only two DVs, at cut-in and d/dU , are required to fully dene the constant below rated optimal operation. Of course, using only two variables will ignore any drive train ineciencies that may, in general, be non-linear and coupled with via the transmission, and Re eects. With variation as well, the optimal track in the -U plane will certainly no longer be linear. Therefore, the most generalized approach is to remove the denition of from the list of DVs. Speed optimization is then handled as a subproblem for each iterate, using the methods in 4.7.1.2. Analogous to using splines to describe the blade prole (see 6.5.1.1), constraints and optimality conditions are implicitly included in the formulation rather than the optimization process itself.

6.4.3 CONE-450 Approach


The CONE-450 concept was developed roughly by the following steps [66]. No stall delay or coning corrections were used in the original study. Vortex-lift corrections (based on delta-wing theory) were used in the coned-parked cases, with full 3D wind loading.
The map can be easily constructed once the operating prole over wind speed are known.
1

6.5 Parametric Study

195

1. Rotor size was determined by rst picking a rated wind speed (Vr =12 m/s) and rated cone angle (25 ). The CT of the REF-450 at this speed was then assumed equal to the CT of the coned rotor at rated, and used to compute the rotor diameter for equal maximum thrust as the REF-450. 2. Chord and twist proles were optimized on a section-wise basis for = 8.5 and = 0 . 3. Rotor speed schedules were computed for maximal energy capture, given a prescribed schedule. Peaks near rated were removed. 4. The chord proles were linearised and twist adjusted to accommodate hingeline and avoid excessive energy capture degradation. 5. Steady blade loading was computed at 11 m/s with 200 kNm hinge moment applied. Blade structure is then computed to survive the static forces, and mass-tuned to achieve 25 cone angle at Vr . 6. Final speed schedules were computed with free-coning and 60 kNm hinge moment between 59 m/s. 7. Detailed design of components and controller. 8. Iteration on blade structure and component design, based on survival of computed dynamic loads. The blade structure was designed with steady loads, eventually based on a steady 560 kNm applied hinge moment, to survive the dynamic fatigue and limit loading.

6.5 Parametric Study


The analytic methods developed in Chapter 4 (BEM, power curve solution, dynamic simulations), in their integrated form, are unfortunately not readily amenable to gradient-based optimization methods (e.g. Sequential Quadratic Programming (SQP)). Unlike a coupled FEM-CFD code, complex-step [182], auto-dierentiation, and adjoint methods are unable to provide easy and accurate derivative information. Finite-dierencing techniques are expensive, and were found to suer from code solution granularities. For example, it was found that the BEM code must be run to < 106 precision (see 6.5.2.2), while the power curve must be computed to less than 0.2 m/s step-size resolution. Each optimizer step requires N calculations to compute the gradient (for N DVs), plus a line-search step for SQP, making this approach quite time-consuming when the gradients are computed accurately for the complete coupled system.

196

Chapter 6 Rotor Optimization

Of course, heuristic algorithms (e.g. genetic algorithms [65, 183], simulated annealing,1 particle swarm, random walk) avoid gradient computations, but substitute many more direct objective and constraint calculations for each iterate. Tuning of the algorithms is also dicult, often requiring close to a full-factorial search after tuning! They are of course more able to avoid local minima than gradient methods, but at great expense. Simplex methods [72] can be more ecient, and also avoid the need for derivatives, but are not adept in incorporating general constraints. Another approach is to simplify the design space. Reduced order modelling [184], including basis functions, response surfaces, and Kriging, approximate the true design space. The optimizer then operates in this new space, which is much less expensive to evaluate and usually smoother. The main diculty in these methods is obtaining a sucient delity and accuracy of the reduced design space. Computational expense typically limits the number of DV to 5. Similarly, variable delity methods [185] switch between high and low-order codes (if available), using the former to update corrections to the latter in which the optimizer operates. The coning rotor design space is described by the following DVs: chord, twist, airfoil (thickness & shape), material structure, and operating parameters. Given the size and complexity of the design-space described by these DVs, a parametric study approach (Design of Experiments (DOE)) is adopted here, in common with other authors [186], rather than pure optimization of a nal solution. Gradient-based (efcient) optimization is used where appropriate, but exploration of the tradespace is considered more relevant at this stage than simply applying the optimization methods just outlined. The function used is Matlabs fmincon, a rened SQP algorithm. The specic DOE approach adopted is essentially to perform full-factorial studies on sub-sets of the problem. The author has tried other DOE methods in the past, such as Latin Hypercube and orthogonal arrays, and based on this experience has found the current approach more elucidatory. The primary areas explored in the following sections are: the inuence of coning on optimum aerodynamic prole (6.5.4); best-case energy advantage of the coning rotor over conventional designs (i.e. blade length increase, 6.5.5); control strategy ramications (6.5.6); and the inuence of aerodynamic uncertainty (6.5.7). A common element to all these areas is the impact of airfoil selection. Before this presentation, the specic choice of DV and parameters is given in 6.5.1, numerical optimization issues in 6.5.2, and the optimization approach adopted here to deal with the sub-problems in 6.5.3.
1

Adaptive simulated annealing (ASA) by Lester Ingber from http://www.ingber.com/

6.5 Parametric Study

197

6.5.1 Design Variables and Parameters


The coning rotor is assumed to mount onto the same generator/tower as the REF1500. No tower height advantage is included in the present results, as the parked loading of modern rotors is more benign than the comparison rotor used by Jamieson [66], and is less likely to aord such large increases. The results are therefore only reective of changes in the rotor itself. Hinge radius Rhinge has been chosen as 2 m to achieve a realistic hub and blade-root structure and to accommodate actuators capable of maximum coning deection, positioned downwind of the rotor. The control schedule is treated implicitly as outlined in 6.4.2. The control method (VSS, PTS, PTF) is treated as a parameter. 6.5.1.1 Shape Profiles The blade shape proles are controlled to provide the realistic geometry shown in Fig. 6.6, thereby avoiding the smoothing steps mentioned in 6.4.1. Bezier splines are chosen over cubic splines to dene chord and twist proles, as they posses the convenient convex-hull property [187], thereby avoiding oscillations in the prole. The deCasteljau algorithm is used for computation of the polynomial coecients. The spanwise locations of all control points are xed, except for the rst twist point, while the DVs prescribe the chord and twist at those points. The actual section locations move with the prole itself (not control polygon), to place a section at maximum chord and with cosine spacing towards the tip.

Figure 6.6 Chord and twist prole control A 7th order Bezier curve is used to dene the chord denition. The outer 4 points at s = [0.15 0.4 0.7 0.95] of blade span S are active DVs for chord. The

198

Chapter 6 Rotor Optimization

blade shapes incorporate a 2.2 m circular root (xing the rst spline point), rather than the blended root possible with a long ap hinge, in case a pitch actuator is required. The second and third points chords are equal to the the rst and fourth points respectively. The outer 5% of the blade has a parabolic tip shape, for noise mitigation. This also produceds a low optimization sensitivity, desireable as Prantls F factor reduces the solution sensitivity in this area. The chord formulation allows the cu (maximum chord cmax at sc,max ) location to move spanwise implicitly, without requiring another DV. It was found that chord is best bounded by DV limits (0. . . 1.5 3.8/stip m)1 and monotonic linear constraint on the DV to ensure reducing chord with s. An alternate to the linear constraint is to dene the 5th and 6th control points relative to the 4th and 7th points. This adversely aected the optimizer sentivities to these DVs. Twist is controlled by a 4th order Bezier curve and 4 active DVs. The rst control point is placed at sc,max (not 0.15s), and constant twist is dened inboard of sc,max . Twist is constrained by DV bounds of -10 25 . The airfoil thickness is tapered linearly from 100% at the root to 30% thickness at sc,max , and then linearly to 15% at the tip. The pitch axis tapers from 50% at the root to 30% at sc,max , and is uniform outboard. The nal blade shape thus implicitly incorporates smooth geometry and manufacturing constraints. Optimization over the entire power curve is facilitated by this formulation, as sectional optimization is only eective for a single prescribed operating condition. Global shape control also aids the optimizer by equalizing the sensitivity between DVs. Simultaneously, a limited number of DVs are required, rather than DVs to control each section. 6.5.1.2 Airfoils Airfoil choice is treated as a parameter, rather than a discrete DV, avoiding mixed discrete-continuous optimization issues. As will become apparent in 6.5.6 and 6.5.7, the proximity of the optimum operation point ( for (cl /cd )max ) to stall , and the behaviour after stalling, is critically important for blades incorporating particular airfoils. As a parametric study, the actual airfoil data used here is incidental; three sources have been used to dene three airfoil families for this study (sets I, II, III). Each family denes a complete range of thickness, and covers a range of characteristics,
The 1.5 factor is required to achieve a cmax of 3.8 m, since the DV are spline control points of a bounding polygon.
1

6.5 Parametric Study

199

to highlight the inuence of airfoil choice. In a real-world blade design, the operational Re and blade soiling levels must be carefully matched to the test data used in an aerodynamic design, to achieve the desired performance characteristics. Sections 3.3.3 and 5.1.6 have both highlighted the widely varying data obtainable from dierent sources, for ostensibly the same airfoil, at dierent Re. The root airfoil for all three sets is a circle with a cd of 0.3. Set I uses LS1 data from GH (same as REF-1500 in C.3), dened for 30%, 21%, 17%, and 13% thicknesses. Set II uses the NACA 63-4XX data given in C.6 to dene a second airfoil set over 21%30%. Finally, set III uses the FX66-17AII-182 data from C.5 for 17% thickness, combined with the 21%100% data of the NACA 63-4XX data. For thicknesses less than the covered range, the thinnest airfoil in the set is used. These airfoils were chosen to be representative of increasingly sharply stalling and high-performance (high cl /cd ) airfoils. The source data covers a range of Re that do not completely cover the operational range of Re, but is sucient for its purposes here. To reiterate, the data is illustrative, as each airfoil designation will produce diering performance with operating conditions. Centrifugal stall delay (4.4.8) is included in the following studies. Spanwise ow (4.4.9) is inactive until 6.5.7. 6.5.1.3 Internal Structural Layup A conventional glass-epoxy layup is used for this study. Future renements will likely include wood/bamboo-carbon-epoxy structural concepts. The layup and material values (see Table E.1) of the section are the same as that used by Grin [55]. Figure 6.7 shows a typical section. The layups for the outer skins and webs are given in Table 6.1 and Table 6.2 respectively. The webs join the ends of the spar caps. The relative placement of elastic (shear) and mass centres discussed in 4.6.1.3 was not considered here. A static load safety factor of 1.5 and cumulative material safety factor of 2.94 (representing the standard combination of factors [55]) are used to check all computed points in the cross-section against strain allowables.

6.5.2 Optimizer Tuning


Optimization is not simply a black-box solution. It requires a number of nuances in its application, including proper DV scaling (6.5.2.1) and set-up (6.5.2.2).

0.3 0.2 200 0.1 0 -0.1

Chapter 6 Rotor Optimization

-1

-0.9

-0.8

-0.7

-0.6

-0.5

-0.4

-0.3

-0.2

-0.1

Figure 6.7 Typical section layup Table 6.1 Outer skin layup schedule (Adapted from [55]) Layer 1 2 3 4 Chordwise Location 0.01.0 0.01.0 0.01.0 0.00.15 0.150.5 0.50.85 0.01.0 Material Gel coat Random mat CDB340 Balsa Spar cap mixture Balsa CDB340 Thickness 0.51 mm 0.38 mm 0.89 mm 0.005c Variable 0.01c 0.89 mm

Table 6.2 Web layup schedule (Adapted from [55]) Layer 1 2 3 Material CDB340 Balsa CDB340 Thickness 1 mm 0.01c 1 mm

6.5.2.1 DV Scaling

Student Version of MATLAB

Linear scaling of all DVs as per Eq. (6.5.1) is used throughout the present results. This yields gradients of the same order for each design variable and well-conditioned Hessian matrices [184]. The diagonal D matrix and vector c are calculated so that the scaled design vector y [1 . . . 1] corresponding to the limits on the dimensional design vector x. Scaling based on the condition number of the Hessian matrix (by scaling to unity the diagonal entries of the Hessian [188]) was experimented with, but found to have minimal eect on the results. The Hessian was already well conditioned with the existing scaling, proving to be a red herring, once the issues of solution accuracy in the following section were addressed.

x = Dy +c

(6.5.1)

6.5 Parametric Study 6.5.2.2 Finite Differencing

201

Anderson [72] stated that the BEM equations are not amenable to gradient-based optimization methods, due to the inaccuracy of the xed-point iteration solution. Selig and Tangler [189] also make reference to diculties in selecting appropriate nite dierencing step-sizes. A closer examination reveals that the nite dierencing step is quite sensitive, but tractable, by judicious selection of the solution tolerance tol and dierencing step-size h. A simple rst-order forward dierence formula is used throughout: f (x + h) f (x) df (x) = + O(h) dx h (6.5.2)

Figure 6.8 shows the eect of the BEM solution tolerance for the axial and tangential induction factors on the computed Finite Dierence (FD) approximation for chord and twist DVs. A tolerance of 1e-10 has been used to avoid solution tolerance errors. Above and below step sizes of 1e-6 and 1e-12, truncation and round-o error occur respectively. A manual step-size of 1e-6 has been used for the FD computations, in deference to the adaptive algorithm in fmincon.
1.4 1.2 1 0.8 1.4 1.2 1 0.8

dCP/dx

0.6 0.4 0.2 0 -0.2 -14 -12 -10 -8 10 10 10 10


-6 -4 -2

dCP/dx

0.6 0.4 0.2 0 -0.2 -14 -12 -10 -8 10 10 10 10


-6 -4 -2

10

10

10

10

10

10

h
(a) tol = 1e-4

h
(b) tol = 1e-10

Figure 6.8 Finite dierencing prediction for CP as a function of solution tolerance tol and step size h for chord and twist control points

6.5.3 Stepwise Optimization Approach


A sequential approach is followed for determining the optimum blade prole, structure, and operating schedule, given the xed parameters examined in the following

202

Chapter 6 Rotor Optimization

sections. The chord and twist optimization steps are done at a prescribed cone angle design,aero and (real) tip speed ratio design,areo : 1. The equations of 4.4.13 are used to compute initial chord and twist values for each blade section. 2. A fast non-linear least-squares t optimization determines the chord and twist
DVs of 6.5.1.1 that best match the analytic values over 0.15S0.95S (i.e. ac-

tively controlled section of the blade). This is done using Matlabs lsqnonlin function. 3. The SQP algorithm is used to nalize the blade prole with the bounds in given in 6.5.1.1, with CP as the objective at = max . With the resulting nalized aerodynamic design, apwise and edgewise bending loads are computed for a prescribed balance . These loads are used to numerically optimize (with SQP) the symmetric spar cap thicknesses for each section on the blade. Additional structural design variables were experimented with, including independent spar cap thickness, variable chordwise extent of the spar caps (both symmetric and with slanted webs allowing alignment of loading with section twist), and materials (wood, carbon). Weight reductions were certainly possible with more
DVs, but the static design loads would have to be supplemented by operational

fatigue simulations to verify the feasibility of the results in practice. These results did not aect the following remarks on the coning rotor, and are therefore excluded for brevity. The structural implications of the airfoils (in terms of structural shape) are of course implicitly included. Recent eorts to develop atback airfoils to maximize structural eciency have taken the airfoil shape DV to an extreme, with encouraging results [180, 190]. Finally, the blade is mass balanced, with a point mass at the tip, so that freehinging to balance is achieved at Vrated =11 m/s and maximum rotation speed max . The optimal operational schedule is computed, with a 400 kNm hinge moment applied from 58 m/s to an otherwise free hinge. This is based on scaling the CONE450 60 kNm low-wind applied moment. The directly applied moment was found to be increasingly ineective at higher wind speeds, requiring control either via pitch (aerodynamic moment) or (aerodynamic and centrifugal moments). A 7.5 m/s hub-height wind (Vmean ) is used for steady energy yield calculations, with shape factor k = 2.2.

6.5 Parametric Study

203

6.5.4 Aerodynamic Optimum


Some initial design point is required for the coning rotor. Rather than picking one design condition, as in 6.4.3, a range of blade designs is considered using airfoil set I. Section 4.4.13 predicts that the optimal blade shape should vary with the design,aero of the aerodynamic design point. As with conventional rotors, the design (real) tip speed ratio design,aero also has a large eect on performance. For this study, a 40.15 m blade is used (same as REF-1500), balance = 25 and is bounded to 1018 RPM. Figure 6.9 shows the results of the optimization. Figure 6.9(a) shows the familiar trend of increasing CP with design,aero , owing primarily to reduced rotational momentum losses. Evidently increasing design,aero also increases the eciency of the rotor, at the design point. Figure 6.9(b) also shows higher design,aero rotors to be more ecient in total energy capture. As the blades are all mass balanced to equal balance , the variation with design,aero is explained with reference to Fig. 6.10. The higher design,aero are more slender, mirroring the eect of increasing design,aero . The aggregate result is an increase in design,aero with design,aero for the isoperformance lines of Fig. 6.9(b). More subtle eects are also present, such as the o-design performance dierences ( = design,aero ) and dierent thrust behaviour aecting equilibrium coning angle over the power curves, away from balance . Figure 6.9(c) illustrates the blade mass variation with design variables. The increasingly slender proles with increasing design,aero lead to lighter blades. The variation with design,aero is somewhat more complex. It appears that for larger design,aero , blade mass grows more rapidly at lower design,aero , with the converse true at higher design,aero , relative to design,aero = 0 .

6.5.5 Blade Length


At the heart of the coning rotor energy advantage is an increase in blade length. This increase cannot be unbounded, and is grossly limited by the tolerable increase in rotor thrust at rated. To explore the eects of varying blade length, this section varies blade length Stip and balance . The variation of balance eectively varies the rated capture area, thereby altering the maximum thrust. Based on the results in 6.5.4, design,aero is chosen as 0 and design,aero as 8.0. The bounds on are varied with Stip to achieve unconed tip speeds of 4580 m/s as proxy noise constraints and to introduce peak shaving.

204
8
0.4 98

Chapter 6 Rotor Optimization


8

101.5

10

0.5

0.514

design, aero (real)

design, aero (real)

7.5

0.4
7 6.5 6 0

8 0.50 6 0.50 04 0.5

0. 49 4 0.4 0. 0. 0.4 88 49 49 86 2 0.4 0 82 .48 4

(a) CP (real) at aerodynamic design point


8

94

design, aero (real)

7.5
98

96

6.5
106 108

102 104

104
0 10

110
30

20

design, aero (deg)


(c) Blade mass (relative to ensemble mean, %)

Figure 6.9 Power coecient CP , annual energy yield Eann and total blade mass variation with aerodynamic design tip speed design,aero and cone angle design,aero

96

102

98

7 100

98

96

2 0.51

7.5

101
100.5

1 0.5

100.5 7
100

50 0. 2
0. 49

100
99.5 99 98.5 10
98

99.5 6.5 99

0.
6 20

98.5
30 6 0 20

5 0. 8 49

10

30

design, aero (deg)

design, aero (deg)


(b) Eann (relative to ensemble mean, %)

92

100

6.5 Parametric Study


4 3

205

c (m)

Increasing 2 1 0

design,aero

10

15

20

25

30

35

40

s (m)

Figure 6.10 Blade chord proles for design,aero = 68

Figure 6.11 shows the overlapped contours of energy yield and maximum thrust; both are relative to the values for the REF-1500. Note that as Prated is constant,
CF will follow the same trends (see 3.2 and 6.2). For all airfoils, the limit on Eann

increases with balance ; the blade length must also increase to maintain Eann . The blades weights obtained using the stated design condition are comparable to the REF-1500. The unbounded energy capture performance of all airfoils and control strategies is fairly similar with variations in Stip and balance . The achievable energy increase is, however, markedly dierent. The PTS rotors are self-limited by higher thrust levels, relative to the VSS rotors. The dierence in stall behaviour of the airfoils is also manifest by an increase in tolerable Eann with the sharpness of stall. Modifying the thrust bounds either side of 100% produces relatively similar increments in Eann between the various rotors. The energy yields of the rotors relative to the REF-1500 range from 030%, which at the upper-end is similar to the original 29.9% for the CONE-450/REF-450 comparison at 7.5 m/s (with tower height advantage). Figure 6.12 compares the REF1500 power curve to those of the set II VSS and PTS rotors considered next in 6.5.6 (chosen for their equal maximum thrusts). A shape factor of k = 2.2 was used, and Vmean is referenced to a nominal hhub =84 m. Figure 6.12(a) illustrates the generic non-linear trends of decreasing Eann with Vmean . The coning rotors maintain an energy advantage across the range, and all rotors benet from hhub increase relative to the nominal 84 m. The superior below rated power curve of Fig. 6.13 for the VSS rotor is evident in the Eann results. Figure 6.12(b) shows the coning rotor performance relative to the REF-1500 at equal Vmean and constant hhub =84 m. Again, increasing hhub is clearly benecial,

206
50 48 46 44 (m) 42 40 38 36 34 32 20 (m)

Chapter 6 Rotor Optimization


150

46 44 42 40 38 36 34 40 32 20 25

130 125 05 1211 0 115 1000 1 95


25 30

5 110 1 1 5 90 5 10 00 8 1 5 9 0 5 9 8
35

140 135 130 125 120 115 0 10 5 1 19 0 0 105 9 85 00 5 1 9 9085


30 35

tip

tip

10 11 11 5 0 5
40 40
115 0 11 135

145 0 14 5 13

50 48

150
145

balance (deg)
(a) I PTS
50 48 46 44 (m) 42 40 38 36 34 32 20
tip

balance (deg)
(b) I VSS
145
50 48 46 44 (m) 42 40 38 36 34 32 20
150

150

140 5 135 11 130 125 120 115 95 0 10 9 5 1 8 105


95 30

145
140 135 130 125 120

10 105 100 1

100
25

90 35

115 10 1 105 100 15 95 1 90 110 1 85 05 100


25

tip

95
30 35

90

40

balance (deg)
(c) II PTS
50 48 46 44 (m) 42 40 38 36 34 32 20 25 30
tip

balance (deg)
145

(d) II VSS
50 48
145
140

150

150

140 135 10 130 105 15 1 1 25 0 1 10 5 120 9


90 5 8

46 44 (m) 42 40 38 36 34 40 32 20

115 110 105 100 95


90
35
85

105 100

95 90 5 8105 100
25 30

130 125 120 115 110

tip

95

90

85

35

40

balance (deg)
(e) III PTS

balance (deg)
(f) III VSS

Figure 6.11 Relative energy yield and maximum thrust variation with blade length Stip and balance angle balance ( energy yield; thrust)

6.5 Parametric Study


200

207

Eann/Eann,ref (%)

150

100

50

10

11

12

Vmean
(a) Eann relative to REF-1500 at 7.5 m/s and hhub =84 m
200

180

Relative Eann (%)

160

140

120

100

10

11

12

Vmean
(b) Eann relative to REF-1500 at equal Vmean and hhub =84 m

Figure 6.12 Energy yield relative to REF-1500 varying with mean wind speed Vmean and hub height hhub ( 1.0hhub , 1.2hhub , 1.4hhub ; REF-1500, VSS, PTS)

208

Chapter 6 Rotor Optimization

more so for lower wind speeds. The relative impacts of the coning rotor and tower height eects are illustrated by the curves for the REF-1500, which only benet from hhub increase. The low wind advantage of the coning rotors also mirror the results of Jamieson [66]. Without denitive cost metrics, no precise answer can be gleaned for overall COE of the coning rotor. As all rotors have equal installed generator capacity, the CF will clearly benet for a coning rotor, as discussed in 6.2. The slope dEann /dVmean is largest around 7.5 m/s, indicating the cost trade-os are most acute in this range.

6.5.6 Control Strategy Impacts


The dynamic implications of the control strategy have been outlined in 3.6.2. Continuing to concentrate on the steady-state operational prole, Fig. 6.13 shows curves for rotors with similar thrust and equal balance , selected from Fig. 6.11. The blade lengths are 38 m, with the exception of 40 m for the set II and III VSS rotors. Although the PTS rotors in 6.5.5 cannot employ rotors as large as the VSS ones, additional constraints are imposed by the generator. A problem synonymous with
VSS is the excess torque requirement to slow the rotor into stall [58, 59]. The large

torque increment after the speed limit has been reached is undesirable from the generator perspective. Unlike a constant-speed stall controlled rotor, which must accept power loss above-rated from passive stalling as the CP curve is traversed, the active stall-controlled rotor can move over the map to initiate stall. The added expense of this strategy is a generator to provide the excess torque. Using the PMG model of 4.8, Fig. 6.14 illustrates the cost variation of active material required to achieve the extreme and rated torque/speeds of Fig. 6.13(f), optimized for minimum cost/torque. The reduction in cost with generator diameter Dgen is quite apparent for a PMG design. For the set III rotor (8001025 kNm), the torque increase leads to a 40% cost increase, based on equal Dgen . The set II rotor (8001120 kNm) leads to a 60% increase. This clearly represents an important design consideration. Recall that dynamic overloads are tolerable by both generator and power electronics; at discussion here are the steady torques that must be reacted to maintain control steady-state control. The excess torque requirement and high-wind behaviour are both directly attributable to the choice of airfoil. More specically, the proximity of the angles of attack () for maximum lift-to-drag ratio cl/cd,max and stall stall are key, as is the slope of the lift curve after stall. Experimentation with rotors designed with three

6.5 Parametric Study


1600 1400 1200 200 I PTS I VSS II PTS II VSS III PTS III VSS 5 10 15 20 25 Thrust (kN) 150 100 50 0 250

209

Power (kW)

1000 800 600 400 200 0 0

10

15

20

25

V (m/s)
(a) Power
40 35 1200 1000

V (m/s)
(b) Thrust

Torque (kNm)
0 5 10 15 20 25

30

800 600 400 200 0

(deg)

25 20 15 10 5

10

15

20

25

V (m/s)
(c)
14 12 Torque (kNm) 10 8 1200 1000 800 600 400 200 0 10

V (m/s)
(d) Torque

6 4 2 0 0 5 10 15 20 25

15

20

V (m/s)
(e)

(RPM)
(f)

Figure 6.13 Operational curves for optimal rotors at balance = 30

210
95,000 85,000 75,000 65,000 55,000 45,000 35,000

Chapter 6 Rotor Optimization


800kNm@18 RPM 1120kNm@13.6RPM 1025kNm@14.8RPM

Cost (Euro)

5.0

5.2

Outer diameter (m)

5.4

5.6

5.8

6.0

Figure 6.14 Generator cost variation with diameter and torque/speed levels

airfoil sets (I, II, III) has highlighted the common experience [36, 58, 59] that airfoil curve characteristics are key to the steady rotor performance curves. Figure 6.15 shows the CP map for the VSS rotors just examined. Note that the data includes stall-delay, heterogeneous blade lengths and blended airfoils, somewhat complicating a simple comparison between the 2D coecients in Appendix C. The aggregate eects of the airfoil behaviour are manifest in the peakiness of the CP,max point of the curves, and the slope variation dCP /d after stall. The set II airfoils have the greatest dierence between cl/cd,max and stall (10 ) and the 17% set III the least (1.8 ). This characteristic aects the stall half of the curve to the left of CP max , where the set III rotor moves quickly into stall as the airfoils reach stall, with minimal angle of attack change. The 17% set III airfoil is also the most peaky in terms of lift-to-drag ratio change with AOA. This means that for the optimized rotors operating at cl/cd,max the performance is rapidly degraded odesign. The set I rotor requires large movement in near rated, but limits extreme torque by the sharp dCP /d around = 3. The stall of the rotor is in fact too pronounced at low , leading to the power drop-o observed in Fig. 6.13(a) above V =23 m/s. It has been possible to optimize the twist to reduce the excess torque requirement for the VSS rotors, but at the expense of below rated performance. There are two other main dierences between VSS and PTS rotors: the coning angle above rated is less for the PTS rotors and the thrust levels higher. Both

6.5 Parametric Study


0.6 0.5 0.4

211

CP

0.3 0.2 0.1 0.0 I II III

12

Figure 6.15 CP curves for = 0 and varying airfoil set

eects arise from maintaining a constant above rated. This rotationally stiens the rotor, keeping it more open. In turn, the capture area is larger, increasing the thrust somewhat. More dominating, the faster rotation speed for the PTS rotors means that the inow angle is reduced relative to the VSS rotors and the relative velocity at the blade higher. By examining the local ow and resulting forces, it is found that even though the lift/drag coecients are similar, the ow properties result in a higher stream wise force component and smaller torque-inducing component for the
PTS rotors. Evidently the centrifugal stiening force, proportional to 2 , dominates

the proportional thrust force in determining the equilibrium cone angle.

6.5.7 Aerodynamic Uncertainty


When interpreting design and optimization predications, it behoves the designer to appreciate the limitations and potential pitfalls of the modelling tools employed. In the case of the coning rotor, the primary area of modelling uncertainty is in the aerodynamics. Many of the issues are not unique to this concept, but extend to any wind turbine, especially those operating in stall and at high yaw angles. The issues have been discussed in various sections of Part II, but are unied here to inform future designers. The issues may usefully be divided into an outer problem of induction and inner problem of 2D section aerodynamics. In the BEM model, these two aspects

212

Chapter 6 Rotor Optimization

are coupled by the velocity components resolved parallel to each blade section. Fulleld CFD methods fully couple the two problems by resolving the ow with ne scale grids near the blade surfaces, linked to coarser meshes extending well downstream into the wake. Due to this tighter coupling and more fundamental approach, CFD methods are theoretically better placed to avoid the issues discussed below. Practically speaking however, they remain too expensive for application to wind turbine design work, necessitating the use of BEM theory with certain caveats. Issues with the outer problem are typically manifest at lower wind speeds, and are important for energy capture (Region II). Inner problem issues appear especially during stall, when power limiting (Region III). The outer problem is essential that of determining the proper induced velocities which modify the incident ow as it approaches each blade section and in the wake. The origin of the induced velocities is the wake shed by the blades. It includes the following aspects: Flow at the Disc The induced ow at the disc contains axial, radial and azimuthal components. The proper prediction of all three components is important (see 4.4.2). Far Downstream For BEM, the velocities seen at each section on the disc must be properly related to those in the far wake (see 4.4.6) to balance the momentum equations. The modied BEM method in this thesis includes both of these considerations, producing improved aerodynamic and loading predictions relative to standard theory (see 5.1.2). However, within the connes of a computationally appropriate BEM method, only the dominant tip vortex tube may be considered (i.e. without further iterations on more complicated wake structures). This ignores the variation in spanwise loading on a real rotor which creates multiple vortex tubes, further modifying the spanwise variation of induction. A number of other approximations are also necessary: 3D Tip/Hub Losses The loading at the hub and tips of the blades must trend to zero. Physically, this is an artefact of a nite number of nite

blades creating distinct vortex sheets. BEM theory models this eect in the momentum equations as averaged induction factors via the Prandtl F factor (see 4.4.3.3).

6.5 Parametric Study Wake Expansion The trailing wake vorticity must strictly freely follow the streamlines. As induction increases, the wake must expand to satisfy continuity. In turn, the induced velocity at the disc is modied by the modied strength and location of the vortex sheets (see 4.4.11). High Induction The momentum equations are fundamentally in error, even with wake expansion, at very high induction factors, as they violate continuity. The thrust models of 4.4.3.4 attempt to overcome this deciency, but owing to the fundamental nature of the problem are unable to capture the recirculating stream tubes. Yawed Flow Akin to a coned rotor, a yawed rotor experiences varying induction with azimuth owing to the relative position of the wake.

213

Wake expansion is better handled by the new BEM method (see 5.1.2.1), however very high induction factor conditions remain quite challenging to predict for both coned and unconed rotors. Yaw prediction has been improved with current method (see 5.1.8), but again includes some error because a further iteration of the solution for each azimuthal blade position is not possible in the general dynamic case. The inner problem consists of predicting lift (cl ), drag (cd ) and moment (cm ) 2D coecients based on the velocity seen at the blade (as shown in Figs. 4.4 and Fig. 4.11). Even with perfect resolution of the outer problem, the 2D sectional problem is complicated by the following aspects: Large AOA Measurements and numerical predictions are fairly accurate in attached-ow regions (small AOA below stall). Design critical points occur when the airfoils are stalled, even out to 180 , which is far outside the standard test range for an airfoil. Fortunately, airfoils behave similarly past a critical AOA of approximately 50 enabling extension of available data (see Appendix C). Even below stall, drag prediction in particular can lead to multiple datasets, for ostensibly the same airfoil [177]. Operational Changes Surface modications during operation (e.g. soiling, ice formation) will modify the 2D characteristics. Re Eects Wind turbines operate in the Re range over which airfoils undergo large performance changes, requiring data in the proper regime. Spanwise Flow The independence principle (see 4.4.9) is generally valid below stall. For stalling rotors, large modications to the 2D characteristics may occur owing to boundary layer transport. The models used here are only a rst approximation to this condition, which should receive more

214

Chapter 6 Rotor Optimization detailed verication. In particular, the root and tip regions will tend to undergo opposite changes (higher and lower maximum lift respectively) as the boundary layer migrates outboard.

Centrifugal Pumping In addition to the spanwise ow stall delay (primarily important for coned rotors), centrifugal pumping (see 4.4.8) will modify lift/drag properties. The fundamental mechanisms remain misunderstood and only partially modelled. Dynamic Flow The ow around a 2D section behaves dierently in steadystate than dynamically. The ow responds with some lag, both in stalled and unstalled conditions. The general case, including spanwise ow, may only be treated with approximate models, which must be tuned to the particular airfoil section employed. These individual modelling diculties ultimately manifest themselves in the aggregated performance predictions. It is well recognized that stall prediction is very dicult for any rotor (i.e. even unconed) [91]. In fact, the verication of above rated performance for a real machine revealed a 17% under-prediction of power [58]. Surface roughness eects will drastically change stall behavior, although that should be favorable from a safety perspective, as power and torque can only decrease; from an energy capture perspective, it is of course detrimental. Any form of stall delay to higher AOA (see 4.4.8 and 4.4.9) will act to decrease safety margins. This is true for thrust, as well as the generator torque critical for the VSS rotors. The modelling diculties evident in 5.1.6 are a pathological worstcase above stall, given the simplied geometry used in the study relative to most commercial blades, and represent and upper bound on aerodynamic uncertainty. Based on these results, power predictions may be erroneous by up to 3060%, varying with the stall delay model used. To provide some quantitative indication of the aggregate modelling eects on the coning rotor predictions, the operational curves for the set II rotors of 6.5.6 are compared in Fig. 6.16. For the VSS rotor is increased by adding the stall delay models. The extrema value is approximately 4% too low without any model. The
PTS rotor predictions are in error in , in this case by 4% in extreme value. In both

cases, the spanwise ow eect is small around rated, but approximately equal to the centrifugal pumping eect at Vco . Adoption of a PTS strategy would provide a measure of design safety on the dicult stalled-ow predictions. Ample control would be available from the pitch

6.5 Parametric Study


1200 1000 800 40 35 30 25

215

(kNm)

600 400 200 0 No Corr CP CP & SF

(deg)

20 15 10 5

No Corr CP CP & SF

10

V (m/s)

15

20

25

10

V (m/s)

15

20

25

(a) VSS

(b) PTS

Figure 6.16 Variation in operational curves with stall delay models for set II rotors (CP = Centrifugal pumping, SF = spanwise ow)

system to compensate for any real world shortcomings of the model of the system. The dynamic controller design should take this possibility into account by adopting a robust approach. It is dicult to place quantitative bounds on the aerodynamic uncertainty present in aerodynamic predictions for wind turbines. Comparison of the modied BEM method to CFD results in 5.1 did yield bounds relative to a more accurate method, providing some indication of the level of error to be expected. This section has highlighted the various error sources, their domains of inuence, and their conservative or non-conservative eects on design predictions. Ultimately however, further experimental results are required against which to valid the predictive models, in order to obtain more quantitative bounds. Until such time as these results become available, the designer is forced to heed past experience and the insight oered in this section to synthesize robust designs that are relatively insensitive to the modelling errors which may be present.

Chapter

Design Considerations
Using the initial rotor designs presented in Chapter 6, future work must extend the analysis into unsteady simulations, to further rene the design. This chapter highlights a number of design considerations that become critical to the time-domain performance of the coning rotor. The rst set concern the dynamics of the system (7.1). The second area of consideration is the low-frequency acoustic prole of the coning rotor (7.2) that must be accounted for and mitigated in the layout and operation of any potential machine.

7.1 Dynamic Simulation


This section presents an exploration of the fundamental dynamic aspects of the coning rotor. In particular, the non-linear apping response to the various forcing functions is discussed in 7.1.1, followed by aerodynamic damping of the ap and edgewise modes in 7.1.2.1 and 7.1.2.2 respectively. To proceed further will require full dynamic simulations of the rotor, including exible blades and tower, drivetrain, yaw mechanism, and a viable controller. The synergistic modal interactions of the various components will need to be checked on a Campbell diagram and/or by examination of a linearised state-space model. The results of fully non-linear time-domain simulations must then be used to further rene and verify the design. Hopefully the basic insight oered in this section will clarify the execution of these future tasks.

7.1.1 Fundamental Response


The intent of this section is to explore the fundamental non-linear eects that will inuence the loading of the coning rotor, without complicating the picture with controllers or stochastic aerodynamic forces. Manwell et al. [17] presented the work of Eggleston and Stoddard [14] for a linearised model about 0 = 0, incorporating very simplied (linear) aerodynamics, steady yaw motion, and same type of hinge 217

218

Chapter 7 Design Considerations

oset, similar to the presentation in 4.6.3.1. Instead of applied damping or an active moment, a linear spring was used to tune the response to simulate the rst ap mode of a exible blade. The linearised equations were analytically examined, using a solution expansion of constants, cosines and sines of the rotor azimuth angle with xed rotation speed. The linearised equations of motion developed in 4.6.3.2 yielded basic insight into the fundamental frequencies and critical damping (4.6.3.3) of the coning rotor. In this section, the non-linear EOM of 4.6.3.1 are examined numerically, using data from the REF-1500, with the inclusion of tilt angle, unsteady aerodynamics and non-linear cone angles. Constant hinge moment is used to bias the steady cone angle, and a constant rotor speed is used for a downwind rotor. Figure 7.1 is useful for visualizing the motions as constant, cosine and sine responses.1

(a) Constant +0

(b) Cosine +c at = 0

(c) Sine +s at = /2

Figure 7.1 Flap angle solution expansion terms To start with, the aerodynamic contribution is omitted. The gravity term in Eq. (4.6.20b) imposes a cosine varying loading, and a constant term, both biased by the tilt angle t . With t = 0, gravity is purely a cosine input. The analytic solution yields a cyclic sharing term, dependent on the stiness of the hinge spring, determining the relative amount of cosine and sine response of the ap angle to the input. It is found that in general, including the other forcing functions discussed next, that sti blades respond with little phase-lag to the input, while freely apping blades exhibit a /2 phase lag. Therefore, gravity causes a cosine response in a sti rotor, and a sine response for a freely hinged rotor. A critical damping coecient HactD for = 1 may be computed from Eq. (4.6.27), and Hact for a steady 10 ap angle from Eq. (4.6.20b). Releasing from = 0, the
0 azimuth is dened here for the 1st blade vertically upwards, whereas Manwell et al. [17] dene it vertically downwards.
1

7.1 Dynamic Simulation

219

system quickly reaches the example cyclic responses shown in Fig. 7.2. For t = 0 in Fig. 7.2(a), the blade angle phasing is as expected 120 , and blade 1 has maximal at 1/4 rev showing a pure sine response to gravity. Changing to t = 5 , as shown in Fig. 7.2(b), the mean cone angle and apping magnitude is reduced, but the phasing remains the same.
10.5 1 2 3 10 1 2 3 9.8

(deg)

(deg)
2 2.5 3

10

9.6

9.4

9.2 9.5 2 2.5 3

Rotor revolutions
(a) t = 0

Rotor revolutions
(b) t = 5

Figure 7.2 Blade apping without aerodynamic contribution The aerodynamic loading is now added to the equations. In the absence of gravity and wind shear, a trivial result of both the analytic and numerical approaches is that the rotor operates at a steady cone angle. Hinge oset and positive resisting moments/springs reduce the angle. As expected, the aerodynamic moment balances the inertial centrifugal force and imposed root moments. As noted in 5.4.3, the
EOM admit two steady-state solutions, symmetric about = 45 . Dynamically, for

> 45 , the system is unstable; as increases, the moment arm of the centrifugal force increases, but this is overcome by the decreasing centrifugal force magnitude, so continues to decrease. Wind shear is a sine input, and therefore the response for the coning rotor is primarily a cosine one.1 Combined with gravity and tilt, both of these cyclic forcing functions therefore conspire to develop a real or eective2 yaw angle for the rotor. Manwell et al. [17] have shown the eect to be worse for faster-spinning and softer
1 An active hinge moment with some component proportional to will eectively stien the rotor and produce some sine response. 2 I.e. if held in xed in yaw, the rotor tips will not sweep out a circle centred on the rotor axis.

220

Chapter 7 Design Considerations

(no hinge spring) rotors. This has implications for a free-yaw design in which the developed cross-ow on the rotor creates a sine input, owing to the eects of advancing/retreating blades. The sine and cosine inputs will balance at some steady yaw angle. Operation in yaw, either eective or real, will also reduce the power capture somewhat. Removing the aerodynamic contribution again, with t = 0 and yaw angle held at 0 , the hub loads are shown in Fig. 7.3(a), in xed coordinates (i.e. x y z of Fig. 4.1(b), not rotating with blades, ), with the same hinge actuator moment and critical damping as earlier. The main force produced is a steady gravity force in the z direction, from total blade weight. There is also a very small mean y component, and the x component is zero (no aerodynamic thrust present). Both y and z components have a small 3P component (B = 3), /2 out of phase (not visible at the gure scale).
2 0 -2 Fx F F
y z

x 10

3 2.5

x 10

M M M

x y z

Force (N)

-4 -6 -8 -10 -12 -14 -16 0 1 2 3

Moment (Nm)
4

2 1.5 1 0.5 0 -0.5 0 1 2 3

Rotor revolutions
(a) Forces

Rotor revolutions
(b) Moments

Figure 7.3 Hub loads for blade apping without aerodynamic contribution (non-rotating CS)

This behaviour may be explained with reference to Fig. 7.4. Cyclic apping is superimposed on a mean cone angle (from constant Hact ), creating blade paths (equivalently tip paths or blade centres of mass) that are eectively yawed from the mean path by y as the blades pass through the x y plane. The net centre of mass of the rotor (averaging all blades) therefore tracks a circle with frequency BP , oset in the +y direction from the x axis. This eectively creates a rotating imbalance force Fy (), only present for a rotor operating at mean non-zero .

7.1 Dynamic Simulation

221

Figure 7.4 Dynamic situation with apping blades and rotating imbalance Lagging by /2, the maximum ap velocity develops as the blades pass through the x z plane. Operating about a non-zero mean , the sin (radial) component of the ap velocity is non-negligible, and so a Coriolis force is created.1 The net eect of both responses is the y and z cyclic forcing in the hub xed CS, the former with a negative mean. Both were observed to increase in amplitude with fewer blades and larger mean cone angle. Helicopter rotors are somewhat similar, in that the tip-path plane is tilted relative to the shaft axis, to direct thrust forward and sideways for ight control. The eects are smaller though, as the blades are much lighter. In addition, as the mean cone angle is small, both the sin component of ap velocity and magnitude of centre of mass movement are smaller. Two-bladed teetered rotors can obviate the rotor imbalance force with an under-slung rotor, by placing the pivot axis on the rotation axis, at the pre-coned rotor centre of mass. The transmission of Coriolis forces to the hub are eliminated with lead-lag hinges in fully-articulated rotors. The hub moments in Fig. 7.3(b) show Mx increasing then decreasing as the blades ap towards equilibrium. This illustrates the rotor torque moment developed with collective coning motion, to conserve rotational momentum. In this case is held
Note that the Coriolis force Fc = 2m vr is simply a manifestation of conservation of angular momentum and vice versa.
1

222

Chapter 7 Design Considerations

xed, but with variable speed may be used to reduce the moment magnitude. All moments have a small BP alternating component, owing to the eects just discussed. Again, as with helicopters, the hinge oset Rhinge imparts additional moments to the hub from the moment of the forces at the hinge location, relative to the hub centre.

7.1.2 Aerodynamic Damping


The motion of a body through a uid produces forces proportional to the velocity of movement. It is the work of this aerodynamic force component(F (V ) V ) that gives rise to either positive or negative damping. In the latter case, vibrations are reinforced, as the force component is in the same direction as the velocity. If unchecked, either fatigue life is reduced, or catastrophic divergent oscillations may develop. Aerodynamic damping is relevant to both discrete DOF (i.e. the rigid blades ap hinge) as well as the exible modes of the blades [113]. 7.1.2.1 Hinge Flapping Damping The gross motion of the blade is apping about its hinge axis. Jamieson [66] found that tower-top loading with rigid coning links was excessive (natural ap frequency and rotation speed too close), so damped links were envisaged instead. These links, and the collective actuator both also provided damping to compensate for the loss of aerodynamic damping on the stalled rotor. If inadequately damped, this motion can lead to destructive fatigue loading on various components. The ap motion is particularly prone to diculty, as the natural ap frequency is always close to the fundamental rotor speed (see Eq. (4.6.24)). The CONE-450 experience of resonant excitement is therefore an intrinsic issue to the coning rotor. Provision for adequate passive damping in a hydraulic coning actuator mechanism may be readily engineered. A safety factor over the critical damping factor should avoid excessive response. Alternatively, active feedback control may be used to better tune the system response over the wind speed range. The level of aerodynamic damping loss will vary with the rotor characteristics. Figure 7.5 shows the logarithmic decrement d (see 4.6.3.3) for the rotors considered in 6.5.6, obtained from the aerodynamically damped oscillation after release from = 0 . The results include uniform wind, no gravity, dynamic and centrifugal stall-delay models, and are computed at the optimal design operating points varying with V . For V below those shown, the system was over-damped.

7.1 Dynamic Simulation


6 5 4 3 I II III 1.4 1.2 1 0.8

223

d
2 1 0 -1 10 15 20 25

0.6 0.4 0.2 0 -0.2 10 15 20 25

V (m/s)
(a) VSS

V (m/s)
(b) PTS

Figure 7.5 Logarithmic decrement d for ap hinge damping

All rotors experience their lowest levels of damping near rated. Airfoils with sharp stall (set I) experience lower levels of damping. It appears that the PTS rotors are better damped at higher V , whereas the reverse is true near rated.

7.1.2.2 Flexible Body Damping The problems of aerodynamic damping are present in all rotor concepts employing some type of stall control (FSS, VSS or PTS). The industry shift to PTF as machines grew past 1 MW was at least partially motivated by the inherent technical diculty associated with stall, both from a prediction point of view and inherent physical challenges. During the period of transition from FSS to PTS, Petersen et al. [159]1 presented an excellent analysis of the fundamental issues associated with the eects of aerodynamic stall on the modal damping of exible blades. Concentrating on FSS machines, equations were derived and results presented on a sectional and whole blade-mode basis. The damping coecients in apwise, edgewise and coupled directions (from integrated exibility over twisted structural and geometric axes) were examined in terms of aerofoil lift/drag characteristics.2 The inuence in terms of overall power and thrust curves of the section and blade
Later repeated by Burton et al. [16, 7.1.9] Edgewise and apwise are true in this context, parallel and normal to the section chord respectively.
2 1

224

Chapter 7 Design Considerations

were also considered. The main results with applicability to the current concept may be summarized as: A negative power/wind speed curve creates negative damping. For an actively controlled rotor, this may be avoided. Sharply negative cl slopes after stall are detrimental. Dynamic stall plays a large role in modifying the behaviour. Low-lift (smoothly stalling) airfoils are preferred outboard, to minimise the contribution of negative lift after stall. The mode shapes weight the contribution to damping as radius squared, making outboard sections more critical aerodynamically. Structural-pitch towards feather is favourable for edgewise vibration, while minimally detrimental to apwise vibration. The mechanism is a modication of the sectional direction of motion to increase coupling of the edgewise mode to the well-damped unstalled apwise mode. Highly cambered airfoils, with 0 near 0 are preferable to minimise geometric and structural pitch towards stall, thereby aligning the edgewise direction favourably towards feather. Changing the structural or geometric collective pitch angle have essentially the same eect. Therefore, with PTS, edgewise damping can be expected to decrease. Increased chord increases negative aerodynamic damping, as well as increases stiness and structural damping. A compromise is therefore required, with relatively large chord inboard where aerodynamic damping is less critical but edgewise stiness is greatly enhanced. Edgewise mode coupling to yaw and nod modes (i.e. close to blade edgewise frequencies) should be avoided, by making the shaft/nacelle structure as sti as possible. Petersen et al. [159] also derived isolated section damping coecients, by linearising the variation in aerodynamic force F R about operating point Vn and r. The equations were derived assuming zero induction factor, and with linearised aerody namics including structural motion (i.e. relative aerodynamic velocities) in the and n directions vs, and vs,n . The forces in these two directions then have steady and damping (proportional to vs ) components: R F F R R R (r) V F F0 v FR = F R F n s, R R = FR n n Fn vs,n n0
(r) Vn

= FR 0

cR cR n v s cR cR nn n

(7.1.1)

7.1 Dynamic Simulation

225

The resulting equations yield a full damping matrix with these two DOF, with the signs accounting for direction of motion. Without repeating the derivation here, but expressing the damping constants in the coordinate system notation of Fig. 4.4, the damping coecients for an isolated section are: 1 r cR (r, Vn ) = c 2 Vrel 1 r cR (r, Vn ) = c n 2 Vrel 1 r cR (r, Vn ) = c n 2 Vrel 1 r cR (r, Vn ) = c nn 2 Vrel
2 2r2 2 + Vn cd V 2 cl cd Vn V cl + r r 2 2 + 2V 2 cd r cl n Vn cd r + cl + Vn r 2 c 2 2 + V 2 V 2r cl d n Vn cd n cl + Vn r r 2 + r 2 2 2Vn cd cl cd + Vn + Vn cl + r r

(7.1.2a) (7.1.2b) (7.1.2c) (7.1.2d)

A real section will move in a coupled direction, so the damping constants are transformed here to an xB axis with: cB = cos2 (s )cR cos(s ) sin(s ) cR cR + sin2 (s )cR nn x n n where s is the angle measured clockwise from , the same as AOA .1 The sectional power and thrust per unit span are computed with: 1 P = crVrel (Vn cl rcd ) 2 1 F = cVrel (rcl + Vn cd ) 2 (7.1.4a) (7.1.4b) (7.1.3)

Using these equations, a comparison between airfoils and control strategies can be made at the fundamental sectional level. A section radius r of 30 m is used, and local speed ratio r = r/Vn of 6 below rated. A nominal chord of 1 m is used for the set I airfoils, and the other airfoils chords are scaled to achieve the same P at 11 m/s. The airfoils are operated at their respective maximum cl /cd points below rated. The airfoils have static twist tw to achieve this, and are actively pitched by op above rated. All twist angles are positive towards feather. The section damping cB angle s is expressed with str (structural twist relative to chordline) as: x s = str + op + tw (7.1.5)

Edgewise and apwise vibration directions are nominally 90 apart, so varying str = 90 . . . 90 covers both. With the integrated inuence of inboard sections, edgewise vibration is approximately in the range of str = 0 . . . 20 .
1

This is opposite to Petersen et al. [159].

226

Chapter 7 Design Considerations

Figure 7.6 shows the damping achieved for the set I airfoil. All control strategies achieve the same damping below rated, as expected. Above rated, the controls utilizing stall clearly suer from a loss of damping, particularly in the ap direction. The PTS rotor is particularly aected, as a large pitch well into stall is required. In addition, active pitch re-aligns the vibration direction in a detrimental way.
400 300 200 0

cB (Ns/m2) x
25 20 15 10 5 V (m/s)

200

-200 -400 -600 -800

cB (Ns/m2) x

100

-1000 -1200 -90 -60 -30 0 30 (deg) 6090


str

-100 90 60 30

str (deg)

0 -30 -60 -90

25

20

15

10

V (m/s)

(a) PTF
25 400 200 20

(b) PTS

op (deg) or (RPM)

15 10 5 0 -5 -10 -15 5 10 15 20 25 PTF PTS VSS

cB (Ns/m2) x

0 -200 -400 -600 -800 -90-60 -30 5 10 0 30 60 90 15 20 V (m/s) 25

str (deg)

V (m/s)
(d) Control variables

(c) VSS

Figure 7.6 Damping constant variation with control strategy for 15% thick set I airfoil (Ns/m2 )

Additional airfoils from set II and III are compared in Fig. 7.7, for PTS control. The 5 m/s case is quite similar and uniformly well damped in all directions. Moving into stall at 1525 m/s, the behaviour is a somewhat complicated mix of factors.

7.1 Dynamic Simulation

227

Firstly, the set III has the smallest chord (and set II the largest), increasing damping (see earlier list [159]). Secondly, the pitch action shown in Fig. 7.8(a) is largest for the set II airfoil. The vibration direction is therefore more negative, an overall detriment to damping. Finally, and most importantly, note that the stall drop-o in cl is roughly comparable for the set II and III airfoils, and both are much less severe than set I, over the range of AOA shown in Fig. 7.8(b). This is the dominant factor in the highly negative levels of damping for the set I airfoils. The set II and III airfoils experience similar levels of damping, owing to their similar stall behaviour, shifted in str from the overall pitch angle op + tw .
200 150 100 5 15 25

cB (Ns/m2) x

50 0 -50 -100 -150 -200 -80 -60 -40 -20 0 20 40 60 80

II I III

str (deg)

Figure 7.7 Damping constant variation with airfoil for 15% thick airfoils and PTS ( 5 m/s, 15 m/s, 25 m/s; I, II, III) Figure 7.9 shows damping data for a thicker set of airfoils.1 The set I damping is clearly improved across the entire range of V and str , owing to the smoother stall. The set II/III damping is reduced somewhat relative to the 15% airfoils. This is explained by the second stall peak in the thin airfoil data, relative to the more negatively sloping cl curve for the 21% thick data. Apparently the operation at relatively constant op + tw = 8 (above rated) for the set II/III airfoils has little positive damping eect, relative to the -18 angle of the 15% thick airfoil. Overall, it can be seen that there are multiple competing eects that inuence damping, making airfoil choice and blade design all the more complicated. In the original CONE-450 study [66], no edgewise mode shapes were included, only rigid apping and a exible ap mode. Any potential edgewise damping issues would
1

Set III airfoil data is the same as set II data at this thickness.

228
0 -2 -4

Chapter 7 Design Considerations


35 30 25

tw + op (deg)

-6

(deg)
25

-8 -10 -12 -14 -16 -18 5 10 15 20 I II III

20 15 10 5 0

10

15

20

25

V (m/s)
(a) Total twist

V (m/s)
(b) AOA

Figure 7.8 Variation of AOA and total twist angles over wind speed range with airfoil choice (15% thick) for PTS

150 100 50 5 15 25

cB (Ns/m2) x

II & III
-50 -100 -150 -200 -80 -60 -40 -20 0 20 40 60 80

str (deg)

Figure 7.9 Damping constant variation with airfoil for 21% thick airfoils and PTS ( 5 m/s, 15 m/s, 25 m/s; I, II, III)

7.2 Tower Thump

229

therefore not have been manifest in the results. The unavailability of a suitably modied version of BLADEDTM prohibited more detailed examination of these effects for a complete blade, which should be revisited in future work.

7.2 Tower Thump


Lowson and Lowson [139] concluded that the CONE-450 would produce only a minimal noise increment over an equally-rated VSS machine. Within the connes of the analysis method, increased trailing-edge-length was the primary culprit. Tip-noise can be avoided by proper design, and is now relatively well understood and fundamentally limited by tip-speed (<70 m/s). As discussed in 4.5.3.1, an important eect not originally considered is tower-thump, present for any downwind rotor passing through a tower wake. This noise has been problematic in the past for downwind machines [191]. The model of 4.5.3 is used in this section to parametrically examine the eects of wind speed (7.2.1), wake prole (7.2.2), oset from the tower (7.2.3), control strategy (7.2.4), and rigid-body ap motion (7.2.5). Section 7.2.6 gives a brief outline of future directions for acoustic investigation. It is not possible to draw absolute conclusions from the current investigations. Rather, a quantitative indication of the eect of design parameters is sought. A complete design would be required, as would a strict measure of acceptability, for objective quantication in terms of maximum acceptable SPL. Some indications were given in 5.2; Barman et al. [141], who measured real data for a downwind turbine that caused LFN disturbance to neighbours, found pressure peaks of 1.4 1.6 Pa, corresponding to 75 dB peaks around 10 Hz (25 RPM rotor speed). ExcelBEM predictions presented in 5.2 for an up-wind rotor (for which LFN is not a practical consideration), indicate that 72 dB is an acceptable upper-limit on LFN (at 11 m/s and 100 m observer distance).1 Attenuation with distance is minimal for low-frequency noise, so observer distance is not a inuencing factor in this analysis. This will be true even for nearby oshore machines. The REF-1500 rotor is used as a baseline, with the following conguration unless otherwise stated: The observer is 100 m downwind of the tower (0 azimuth), at ground level. The nominal wind speed is 11 m/s, below the rated power of the machine but at maximum rotor speed.
Noise is measured on a logarithmic scale, so the nominal 1215 dB dierence in 5.2 between upwind and downwind is substantial.
1

230 Wind shear is included ( = 0.15).

Chapter 7 Design Considerations

The tower wake prole has wake decit 0.3 and width w 2.5D, both dened 3.25D downwind of the tower, typical values for the wake from a cylinder Wang and Coton [87]. No tilt angle is included. No dynamic motion, only xed cone angles (0 nominal) and rotor speeds (20 RPM nominal). Dynamic stall delay (referred to below as simply stall delay) is turned o.1 Centrifugal pumping and spanwise ow corrections are turned o throughout. The dynamic inow model is used for induction calculation.

7.2.1 Variation with Wind Speed


Barman et al. [141] presents non-dimensional wake prole data over a range of wind speeds for the velocity decit behind a tower, which are fairly invariant with respect to free stream velocity, given the error-bounds. Based on this, the SPL can be expected to grow with wind speed. To test this, predictions were made over the operational prole of the REF-1500, including variable speed and pitch angle (for the nominal control schedule, without stall delay). The results are shown in Figs. 7.10 and 7.11.
95 Stall delay No stall delay

90

SPL (dB)

85

80

75

10

15

20

25

Windspeed (m/s)

Figure 7.10 SPL variation with wind speed These results conrm that the noise grows, at least until rated power. With pitch action, the forces remain relatively constant, and therefore the SPL only continues
1

Eects are only important above rated.

7.2 Tower Thump


80 70 60

231

SPL (dB)

50 40 30 20 10 0 0 5 5 8 15 20 25

10

15

20

f (Hz)

Figure 7.11 Acoustic spectrum variation with wind speed (m/s), including stall delay

to grow with increasing absolute velocity decit. This is counterbalanced against reduced overall force with pitch-to-ne, resulting in fairly constant LFN above rated. Including stall delay eects is shown to have an important eect, since the force magnitudes are larger, as are the transients in negative stall (for 2025 m/s).

7.2.2 Wake Profile Influence


The tower design will inuence the noise signicantly. A thin, circular tower produces a deep, sharp wake, while a multi-element (truss) tower would create a lesssevere but broader wake. Another, possibly costly, option would be a streamlined tower, able to rotate with the rotor in yaw either as a complete unit or by means of a fairing around a circular tower. The inuence of both are examined below. The rotor is unconed in both cases operating at 20 RPM. The wake parameters are dened 3.25D downwind. 7.2.2.1 Wake Width The SPL trend with wake width w (expressed as multiples of tower diameter D), at constant wake decit = 0.3 is shown in Fig. 7.12. Even though the total decit remains constant, the larger time-period means that lift variations are more sedate. Hence, less acoustical noise is produced. Figure 7.13 shows that the acoustic energy is essentially constant for low-order harmonics, but falls o at higher (audible) frequencies with increasing wake width.

232

Chapter 7 Design Considerations

88.6 88.4

SPL (dB)

88.2 88 87.8 87.6 87.4 87.2 2.5 3 3.5 4

Figure 7.12 SPL variation with wake width w


100

80

SPL (dB)

60

40

20

2.5 3 3.5 4 0 5 10 15 20

f (Hz)

Figure 7.13 Acoustic spectrum variation with wake width w

It is worth noting that the wake is assumed uniform. The smoothed, broader wake from a small-element truss tower would mimic this. More substantial members of a tripod-type structure would not, but would create more sharp individual wakes.1 7.2.2.2 Wake Deficit Holding the wake width w xed at 2.5D, Fig. 7.14 shows that as expected, SPL grows with wake decit . It appears that the growth is non-linear, decreasing in
1

This was a problem on the US MOD-1 machine [15].

7.2 Tower Thump

233

additive eect as is increased. Examining the detailed spectrum in Fig. 7.15, it appears that the higher harmonics grow more quickly that the lower ones. The steady pressure remains virtually constant, as the average forces are virtually unaltered with . The quick growth with , especially of the more audible higherfrequency components makes minimization of this parameter (at the blade location) quite important.
96 94 92

SPL (dB)

90 88 86 84 0.2

0.4

0.6

0.8

Figure 7.14 SPL variation with wake decit

90 80 70

SPL (dB)

60 50 40 30 20 0.2 0.4 0.6 0.8 0 5 10 15 20

f (Hz)

Figure 7.15 Acoustic spectrum variation with wake decit

234

Chapter 7 Design Considerations

7.2.3 Mitigation by Offset


Structurally, the rotor hub is most eciently supported near that tower centre. Acoustically, the rotor should be placed as far downwind as possible. One option is an elongated hub, which has been the common solution: Carter [75], MS-4 [9, 192], Proven [193], and WTC [70]. The generator in all cases is overhung on the upwind end of the nacelle, and a shaft transmits torque from the rotor. All but the small Proven machines use gearboxes with induction generators. Klinger et al. [81] and Versteegh [82] present the Vensys and Zephyros direct-drive machines respectively, both of which have conventional upwind rotors. The generators are located on the same side as the rotor, and use ecient compact-hub construction. In order to avoid a long shaft for a PMG coning rotor, the acoustic eects must be quantied. The CONE-450 with space-frame hub located the hinge axes 3 m downwind of the yaw axis, while the direct-drive concept proposed in that report had almost zero oset to the yaw axis [66]. Obviously tower diameter will directly aect any strategy, in turn inuenced by static and fatigue loading.

7.2.3.1 Offset from Yaw Axis Figure 7.16 shows the SPL as the hub oset o is varied by its ratio with tower-top diameter. It is clear this is an eective parameter in mitigating tower interaction noise. Even at 7D however, the SPL is above the 72 dB that is the nominal reduction target.
90 88 86

SPL (dB)

84 82 80 78

Ratio of offset to tower top diameter

Figure 7.16 SPL variation with oset from tower o/Dtowertop

7.2 Tower Thump

235

Examining the spectrum in Fig. 7.17 reveals that increasing the oset has a nonlinear eect on SPL. The 3P SPL is hardly eected, while the mid-range around 10 Hz is most eected. At higher harmonics, above osets of 3D the SPL changes minimally, but is strongly dependent below this oset.
80 70 60

SPL (dB)

50 40 30 20 10 0 0 2 4

1 2 3 4 5 6 7 6 8 10 12 14 16 18 20

f (Hz)

Figure 7.17 Acoustic spectrum variation with oset from tower o/Dtowertop

7.2.3.2 Cone Angle In order to obtain a fair comparison of coned congurations, the aerodynamic power of the rotor is kept constant. This is achieved by scaling the blade sections and chords by a constant multiple. The true tip-speed-ratio is kept constant, so that tip noise would remain relatively unaltered, even if the TE noise would grow somewhat with blade length. The coning hinge is at the outboard edge of the hub (1.75 m) and the nominal 3.3 m oset from the yaw axis is used. With this set of scaled blades, the acoustic variation in SPL, shown in Fig. 7.18, is found to be even more eective than linear oset in reducing SPL, at least for the 0 observer azimuth (directly downwind). In fact, 10 of cone is equivalent to 3D (8.5 m) of oset in this case. The outer portion of the blades, responsible for the bulk of noise, moves fairly quickly away from the tower with . The spectrum of the noise in Fig. 7.19 shows that noise is reduced quite strongly across the entire range. Both lower (infra sound) and higher audible sounds are mitigated more eectively than with simple oset.

236
90 85 80

Chapter 7 Design Considerations

SPL (dB)

75 70 65 60 0 90 180 270

10

20

30

40

Cone angle (deg)

Figure 7.18 SPL variation with cone angle (deg) and observer azimuth angle (deg)

80 70 60

SPL (dB)

50 40 30 20 10 0 0 0 10 20 30 40 5 10 15 20

f (Hz)

Figure 7.19 Acoustic spectrum variation with cone angle (deg)

7.2 Tower Thump

237

With coning, the location and orientation of the acoustic sources changes more than with simple oset. The acoustic footprint for the rest of the variations considered in this section is relatively invariant in prole. Figure 7.20 makes it obvious however that the presence of coning alters not only the magnitude of the SPL, but also the spatial lobes. With increasing cone angle, the upwind lobe is much larger than the downwind one. This means that although the nominal downwind observer considered here benets, more investigation is warranted on the directionality of the
LFN. The trends for other observer locations in Fig. 7.18 show that the eects are

very non-linear, and should be considered in an acoustic evaluation of the coning rotor.
120 135 150 165 90 80 70 60 50 180 -165 -150 105 90 75 60 45 30 15 0 -15 -30

-135 -45 -120 -60 -105 -90 -75

Observer azimuth (0 downwind)

Figure 7.20 Footprint SPL changes with variation in cone angle (: 40 )

20

7.2.4 Control Strategy Effects


Above rated power, the control strategy is found to have a strong eect on the SPL, as shown in Table 7.1. Both stalling strategies (PTS and VSS) exhibit reduced noise relative to the PTF machine (as mentioned in 3.8.3, PTF is un-viable in any case for the coning rotor). Although the steady forces are higher (see 6.5.6), Fig. 7.21 shows that as each blade passes through the tower wake, the change in force is actually much larger for the PTF concept. Acoustic noise is generated by changes in pressures, so the larger variations increase the overall SPL. It is also evident that stall delay eects are important in modifying the response, altering both the magnitude and azimuthal prole.

238

Chapter 7 Design Considerations Table 7.1 SPL variation with control strategy above rated Without stall delay Windspeed (m/s) 15 20 25
20

With stall delay


PTF PTS VSS

PTF

PTS

VSS

90.8 92.0 91.0

84.3 80.7 82.7

83.2 79.6 82.3

90.5 93.2 94.6

86.1 84.0 84.7

84.8 82.9 84.4

T - Tmean (kN)

-20 PTF PTF s PTS PTS s VSS VSS s 0.5 0.55

-40

-60

-80 0.45

Rotor revolutions

Figure 7.21 Rotor thrust variation over azimuth for control strategy choice (20 m/s; s indicates stall delay model included)

7.2.5 Dynamic Motion


Allowing the blades to freely ap has two potentially opposite eects. The rst is that cyclic inputs (e.g. wind shear, gravity), by creating cyclic variation, increase the noise by the motion of the acoustic forces. On the other hand, dynamic motion of the blades, if fast enough, will reduce the transient aerodynamic forces and hence noise. Using the same lengthened blades and rotation speeds as in 7.2.3.2, and retaining the nominal section weights in C.3, steady simulations were then run at each cone angle and the root bending moment noted. This moment was then applied in the dynamic simulations to attain roughly the same cone angle.1 The results given next include dynamic stall delay, wind shear and gravity. Given the observer azimuth dependency found earlier in 7.2.3.2, Table 7.2 gives the SPL results for the same range of angles. There is little variation found by
1

There is a slight oset owing to the cyclic inputs.

7.2 Tower Thump

239

including ap motion, especially for larger cone angles. Examination of the ap angle histories revealed no discernible structural response to the tower wake. The aerodynamic forces were therefore relatively un-altered, although extremely exible blades (i.e. include modal response) may change this result. Table 7.2 SPL variation with dynamic motion and cone angle Cone angle (deg) Without ap motion 0 10 40 With ap motion 0 10 40 Observer azimuth (deg) 0 90 180 270 88.4 83.2 68.5 87.9 83.2 68.7 74.8 64.8 71.6 75.8 66.2 71.5 88.2 85.0 80.0 87.6 85.0 80.2 74.7 75.6 76.5 75.1 75.9 76.5

Apparently, the cyclic ap motion of the blades has fairly minimal eect on the acoustics. The SPL is slightly reduced upwind/downwind of the rotor, and slightly increased SPL to the sides. These trends were observed across the frequency spectrum.

7.2.6 Remaining Issues


The primary requirement is for more experimental data to validate the acoustic model. Although it is physics-based with minimal assumptions, the input tower wake prole is of critical importance. Reynolds number has a large eect on the prole, as does tower diameter and conguration (e.g. truss, multi-element or streamlined towers). Additionally, no account is made of the modication of the wake by the passing blades. In combination with experiment, it has been postulated that this eect widens the wake [87]. Terrain modication of the LFN should also be investigated. In particular, mountainous terrain may channel the noise. Propagation over water should also be looked at for oshore locations. Conventional noise mechanisms should also be quantied. Modern PTF machines operate away from stall. A comparison gross stall noise should be carried out, to compare a stalling strategy to PTF in terms of noise production. The original study [139] did not include this comparison, as both the REF-450 and CONE-450 operated in stall.

240

Chapter 7 Design Considerations

Finally, any nal design must take mitigation of tower noise into account and strive to mitigate the eects. The specic variable rotor speed schedule and turbulent winds should be checked for their eect on the noise prole of the machine.

Chapter

Conclusions
Proposing and advancing an alternate wind turbine concept is a multi-step process, with extremely broad scope, for which this thesis can only begin to tackle fundamental considerations. Part I began in Chapter 2 with an exposition of the contextual placement and requirements of a wind turbine. Chapter 3 then approached the problem from a functional design perspective. The basic trade-os in design choices were used to justify the coning rotor concept and describe the unique features of the concept in topological and operational terms. Its relationship to conventional and past unconventional designs was examined, with a view to exposing the critical elements requiring analysis. Throughout the remainder of the thesis, reference was made to a standard code, BLADEDTM , to identify the areas in conventional models requiring alteration to analyse the coning rotor concept.

8.1 Analytical Contributions


Part II focused rst in Chapter 4 on developing the analytical tools necessary to properly design a coning rotor. Section 4.4 presented a novel BEM method developed by returning to the underlying assumptions and derivation of the theory. By considering the wake structure with vortex laments, a new BEM method was formulated that more properly handles coned and yawed rotors in steady and dynamic ow. Importantly, the low-cost computational structure of the BEM method was retained. It is readily implementable in a fully dynamic, industry-level code incorporating aero-elastic coupling, at negligible computational cost. In 5.1, it was demonstrated that the root of errors in the BEM method lies primarily not in the planar disc assumption, but rather in a decient accounting for the relative position of the wake. The validity of the results has been demonstrated for uniformly loaded and real rotors, with coning and yaw, by comparison with more accurate CFD models, wind tunnel and eld test results. Very high induction factor and stalled conditions, however, remain dicult to predict with great certainty. 241

242

Chapter 8 Conclusions

Building on the aerodynamic analysis, the equations governing low-frequency acoustics were presented in 4.5. Although the equations have been developed by other authors, a number of important nuances in their implementation are presented. No quantitative comparative results were available, however parametric variation produced reasonable predictions from what is an accurate physics-based model. Structural modelling tools were developed in 4.6, beginning with a highly exible cross-sectional property analysis. The equations governing distributed loads in the blades were then developed. Next, the dynamic equations of motion for the rigidbody coning rotor were shown to lead to analytic expressions for ap frequency and damping. Finally, an FEM approach was used to derive a centrifugally stiened beam element with varying mass and stiness properties, assembled to predict the modes of a complete rotor with varying boundary conditions. Analytic and numerical results from BLADEDTM were used for comparison. Combining these elements, the computation of an optimal steady-state control schedule was described, taking into account the equilibrium cone angle and applied hinge moment bias. Based on these integrated analyses, the primary assumptions requiring modication in BLADEDTM were identied. The BEM model was originally thought to be the primary diculty, however it became evident that there are additional linearity assumptions in the structural model of BLADEDTM that are violated by the coning rotor. Simplied PMG models were introduced, based on the driving magnetic materials required. Together with cost estimates for those components, an estimation of the cost-torque sensitivity of the generator was enabled.

8.2 Design Refinements


Part III used the analytic tools of Part II to aect a study of the DVs and parameters determining the topology and operation of the coning rotor. In Chapter 6, results from the original study on the coning rotor were up-scaled to a modern equivalent machine in 6.1. A favourable comparison was made on a loading and energy capture basis. The COE/CF and non-dimensional aerodynamic performance of a coning rotor was then examined in 6.2 and 6.3. The challenge of optimizing a coning rotor, with its inherent aerodynamic-structural integration, was then contrasted to that of a conventional rotor in 6.4. Based on this complexity, and as a precursor step to a full engineering design including dynamic simulations, a parametric study was

8.3 The Next Steps

243

made of blade shapes, airfoil families, and control strategies in 6.5. The results illustrate the bounds placed on enhanced energy capture of the coning rotor by steady loading and generator requirements. The importance of airfoil selection was also underlined. The coning rotor range of energy capture advantage relative to a conventional reference was found to be 10-30% at equivalent tower height. Chapter 7 surveyed the areas of critical importance and uniqueness to the coning rotor. The rst was the dynamic response of a ap-hinged rotor (7.1), in the presence of aerodynamic, inertial and gravitational loading. The non-linear eects (aerodynamic and structural) were found to be important for the coning rotor in terms of modelling and cyclic loading, operating at non-zero cone angle, away from the standard linearised conditions. The aerodynamic damping inherent to the system was presented next, starting with the rigid-body ap mode which may enter regions of negative damping. In the absence of a exible-body simulation code, sectional equations were used to examine the impacts of airfoil choice and control method on the edgewise damping mode. For a stall-controlled rotor, this mode tends to be unstable due to a lack of aerodynamic damping, and may require compensation. Section 7.2 parametrically then explored the low-frequency acoustics of the coning rotor. The eect of tower wake prole, operation and relative position were quantied, to determine the boundaries placed on the coning rotor to avoid this noise source, critical to downwind machines.

8.3 The Next Steps


As is clear from the length of this thesis, numerous technical areas must be properly integrated to advance the coning rotor concept. The work presented here is only a rst step to improve the underlying theory and provide a starting point for further design. The next immediate step is to modify BLADEDTM to incorporate the BEM method and any structural modelling changes required. Additional experimental test data for coned rotor congurations should also be obtained, to further rene the aerodynamic modelling. To further advance the actual design of the concept will require a full set of dynamic simulations (to an established standard, e.g. IEC-14000) using that code. This will of course require a controller, based initially on a rotor design from this thesis. The controller and simulations will also have to revisit parked conditions and mode-switching, over the entire operational prole. Working with the dynamic simulations, the design can be further rened, to denitively size and cost a concept

244

Chapter 8 Conclusions

machine. Jamieson [66] also concluded that the dynamic response of the system was the critical element of the design requiring further investigation. In particular, if a PMG is to be employed, a more detailed design of the generator will be required. The detailed design of the coning actuators and hub will also require careful attention, heeding the lessons of diculties encountered by previous appingblade concepts. Assuming the nal paper concept design demonstrates merit over conventional machines, the subsequent step should be a prototype machine of modest scale, to test the predictive accuracy of the analysis suite and design methods. Based on any required renements, the concept may then be scaled up and implemented on a commercial scale.

Bibliography
The numbers after each reference are the page numbers where the reference is used. [1] International Energy Agency. Key World Energy Statistics 2005. International Energy Agency (IEA), 2005 1 [2] Global Wind Energy Council. Global Statistics 2005. Global Wind Energy Council (GWEC), 2005 1 [3] Tyndall Centre. Decarbonising the UK: Energy for a Climate Conscious Future. Tyndall Centre for Climate Change Research, Norwich, 2005 1 [4] Department of Trade and Industry. The Energy Challenge: Energy Review Report 2006. Department of Trade and Industry, UK Government, July 2006 1 [5] Elsam Engineering A/S. Life Cycle Assessment of Oshore and Onshore Sited Wind Farms. Elsam Engineering A/S, October 2004 2 [6] White, S. W. and Kulcinski, G. L. Net Energy Payback and CO2 Emissions from Wind-Generated Electricity in the Midwest. Fusion Technology Institute, Department of Engineering Physics, University of Wisconsin-Madison, Madison, WI, December 1998 2 [7] Jamieson, P. and Jarey, A. Advanced Wind Turbine Design. ASME, Solar Energy Division, Wind Energy, 16: pp. 2330, 1995 2, 30, 185 [8] Putnam, P. C. Power from the Wind. Van Nostrand Reinhold Company, New York, 1948 4, 30 [9] Armstrong, J. R. C. A Lightweight 3-Bladed 600kW Wind Turbine. In European Union Wind Energy Conference, Goteborg, Sweden, 1996 31, 234, D-1 [10] Cochran, J., Orrell, A., Pappas, C., and Scheer, G. The Future of Wind Power. In 2003 Symposium on the Global Commercialization of Environmental Technologies, University of Washington, Seattle, USA, June 6 2003 4 245

246

Bibliography

[11] Crawford, C. Re-Examining the Precepts of the Blade Element Momentum Theory for Coning Rotors. Wind Energy, 9(5): pp. 457478, 2006 6 [12] Crawford, C. and Platts, J. Updating and Optimization of a Coning Rotor Concept. In 25th ASME Wind Energy Symposium/44th AIAA Aerospace Sciences Meeting and Exhibit, Reno, Nevada, January 9-12 2006 6 [13] Crawford, C. and Platts, J. Updating and Optimization of a Coning Rotor Concept. ASME Journal of Solar Energy Engineering, 2006 (In press) 6 [14] Eggleston, D. M. and Stoddard, F. S. Wind Turbine Engineering Design. Van Nostrand Reinhold, New York, 1987 9, 217 [15] Spera, D. A. Wind Turbine Technology: Fundamental Concepts of Wind Turbine Engineering. American Society of Mechanical Engineers, New York, 1995 30, 52, 103, 188, 232, D-1 [16] Burton, T., Sharpe, D., Jenkins, N., and Bossanyi, E. Wind Energy Handbook. John Wiley and Sons, Inc., New York, 2001 24, 27, 29, 49, 52, 59, 73, 76, 77, 78, 88, 90, 91, 93, 94, 143, 223 [17] Manwell, J. F., McGowan, J. G., and Rogers, A. L. Wind Energy Explained: Theory, Design and Application. John Wiley & Sons, Ltd., Chichester, 2002 10, 52, 217, 218, 219 [18] Harrison, R., Hau, E., and Snel, H. Large Wind Turbines: Design and Economics. Wiley, Chichester, 2000 24, D-1 [19] Hansen, M. O. L. Aerodynamics of Wind Turbines. James and James Science Publishing, London, 2000 9 [20] International Electrotechnical Commission. IEC 61400 Wind Turbine Standards. International Electrotechnical Commission (IEC), Geneva, Switzerland, 1999 10, 11 [21] Green, S. Gotland: The HVDC Pioneer. Power Engineering International, July 2004 11 [22] van Hulle, F. Large Scale Integration of Wind Energy in the European Power Supply: Analysis, Issues and Recommendations. EWEA, Brussels, 2005 11 [23] Heier, S. Grid Integration of Wind Energy Conversion Systems. John Wiley & Sons, New York, 1998 12 [24] Gull, F. Distributed generation versus centralised supply: a social cost-benet

247 analysis. Cambridge Working Papers in Economics (CWPE) 0336. CMI Electricity Project Department of Applied Economics, Cambridge, 2003 12 [25] Cotrell, J. and Pratt, W. Modelling the Feasbility of Using Fuel Cells and Hydrogen Internal Combustion Engines in Remote Renewable Energy Systems. In American Wind Energy Association Windpower Conference, Austin, TX, May 18-21 2003 12 [26] Gonzalez, A., McKeogh, E., and Gallachoir, B. The Role of Hydrogen in High Wind Energy Penetration Electricity Systems: The Irish Case. Renewable Energy, 29: pp. 471489, 2004 12 [27] Surugiu, L. and Pparaschivoiu, I. Acceptability, Enironmental and Social Aspects of Wind Energy. In European Wind Energy Conference, Nice, France, March 1-5 1999 13 [28] Hammarlund, K. and Martensson, A. Planning for Acceptance - Windpower in a Social Landscape. In European Wind Energy Conference, Nice, France, March 1-5 1999 13 [29] Gipe, P. Design as if People Matter: Aesthetic Guidelines for the Wind Industry. In American Wind Energy Association Conference, Washington, DC, March 30 1995 13 [30] Hodos, W. Minimization of Motion Smear: Reducing Avian Collisions with Wind Turbines. NREL/SR-500-33249, National Renewable Energy Laboratory, Colorado, USA, August 2003 13 [31] Dooling, R. Avian Hearing and the Avoidancee of Wind Turbines.

NREL/SR-500-30844, National Renewable Energy Laboratory, Colorado, USA, June 2002 13 [32] Kirby, A. MoD Threatening UK Energy Plans. BBC News Online, March 1 2004 13 [33] Beck, A. Turning to Stealth Technology. Power Engineering International, May 2004 13 [34] Love, M. Land Area and Storage Requirements for Wind and Solar Generation to Meet the US Hourly Electrical Demand. MSc, University of Victoria, Victoria, Canada, 2003 13 [35] Landscape Design Associates. Cumulative Eects of Wind Turbines: A Guide to Assessing the Cumulative Eects of Wind Energy Development. ETSU

248

Bibliography W/14/00538/REP, UK Department of Trade and Industry, 2000 14

[36] Jamieson, P. M. The Prospects and Cost Benets of Advanced Horizontal Axis Wind Turbines. ETSU-W23/00355/REP, Harwell Laboratory, UK Energy Technology Support Unit, 1995 15, 24, 35, 186, 210 [37] Scott, N. W. J. A Flexible Blade Wind Turbine for Electricity Generation. PhD, Department of Electrical and Electronic Engineering, Imperial College, London, 1997 15 [38] Pahl, G., Beitz, W., Wallace, K., Blessing, L., and Bauert, F. Engineering Design: A Systematic Approach. Springer Ltd., London, 1995 15, 16, 18 [39] Chakrabarti, A. and Bligh, T. A Scheme for Functional Reasoning in Conceptual Design. Design Studies, 22(6): pp. 493517, 2001 16 [40] Bracewell, R. and Wallace, K. A Tool for Capturing Design Rationale. In International Conference on Engineering Design, Stockholm, August 19-21 2003 16 [41] Suh, N. The Principles of Design. Oxford University Press, New York, 1990 16 [42] Paynter, J. Analysis and Design of Engineering Systems. MIT Press, Cambridge, USA, 1961 16 [43] Liu, Y., Bligh, T., and Chakrabarti, A. Towards an Ideal Approach for Concept Generation. Design Studies, 24(5), 2003 16 [44] Yao, Z. Constraint Management for Engineering Design. PhD, Cambridge University, 1996 16 [45] Suh, N. P. Axiomatic Design: Advances and Applications. Oxford University Press, New York, 2001 16 [46] French, M. Conceptual Design for Engineers. Springer, London, 3rd edition, 1999. ISBN 1852330279 17 [47] Jensen, T. Function Integration Explained by Allocation and Activation of Wirk Elements. In ASME Design Engineering Technical Conferences, Baltimore, Maryland, September 10-13 2000 17, 19 [48] Taguchi, G., Elsayed, E. A., and Hsiang, T. C. Quality Engineering in Production Systems. McGraw-Hill, New York, 1989 17

249 [49] Savansky, S. Engineering of Creativity: Introduction to TRIZ Methodology of Inventive Problem Solving. CRC Press, Boca Raton Florida, 2000 17, 31 [50] Yang, K. and Zhang, H. A Comparison of TRIZ and Axiomatic Design. Department of Industrial and Manufacturing Engineering, Wayne State University, 2000 19 [51] Malcolm, D. and Hansen, A. WindPACT Turbine Rotor Design Study. NREL/SR-500-32495, National Renewable Energy Laboratory, 2002 20, 30, 35, 191, 192 [52] TPI Composites Inc. Parametric Study for Large Wind Turbine Blades: WidnPACT Blade System Design Studies. SAND2002-2519, Sandia National Laboratories, New Mexico, 2002 24, 116 [53] Shafer, D., Strawmyer, K., Conley, R., Guidinger, J., Wilkie, D., Zellman, T., and Bernadett, D. WindPACT Turbine Design Scaling Studies: Technical Area 4: Balance of Station Cost. NREL/SR-500-29950, National Renewable Energy Laboratory, Colorado, USA, July 2001 24 [54] Coulomb, L. and Neuho, K. Learning Curves and Changing Product Attributes: The Case of Wind Turbines. Faculty of Economics, Cambridge University, December 2005 24 [55] Grin, D. WindPACT Turbine Design Scaling Studies: Technical Area 1: Composite Blades for 80-120 Meter Rotor. NREL/SR-500-29492, National Renewable Energy Laboratory, Colorado, USA, April 2001 24, 192, 199, 200, E-1 [56] Leithead, W. and Connor, B. Control of Variable Speed Wind Turbines: Design Task. International Journal of Control, 73(13): pp. 11891212, 2000 26 [57] Leithead, W. and Connor, B. Control of Variable Speed Wind Turbines: Dynamic Models. International Journal of Control, 73(13): pp. 11731188, 2000 26 [58] Mercer, A. S. 214, C-12 [59] Homann, R. A Comparison of Control Concepts for Wind Turbines in Terms of Energy Capture. PhD, Technischen Universitat Darmstadt, Germany, 2002 Stall Regulation of Variable Speed HAWTs. ETSU

W/42/00293/REP, Garrad Hassan for ETSU, Bristol, 1996 26, 29, 208, 210,

250 26, 29, 208, 210 [60] Grin, D.

Bibliography

Evaluation of Design Concepts for Adaptive Wind Turbine

Blades. SAND2002-2424, Sandia National Laboratories, 2002 27, 31 [61] Lobitz, D. and Veers, P. Aeroelastic Behaviour of Twist-Coupled HAWT Blades. In ASME Wind Energy Symposium, 36th AIAA Aerospace Sciences Meeting and Exhibition, Reno, Nevada, January 12-15 1998 27, 117 [62] Veers, P., Bir, G., and Lobitz, D. Aeroelastic Tailoring in Wind-Turbine Blade Applications. In Windpower 98, American Wind Energy Association Meeting and Exhibition, Bakerseld, California, April 28 - May 1 1998 27 [63] Muljadi, E., Pierce, K., and Migliore, P. A Conservative Control Strategy for Variable-Speed Stall-Regulated Wind Turbines. In 19th American Society of Mechanical Engineers (ASME) Wind Energy Symposium, Reno, Nevada, January 10-13 2000 28 [64] Fuglsang, P., Sangill, O., and Hansen, P. Design of a 21 m Blade with Ris-A1 Airfoils for Active Stall Controlled Wind Turbines. Riso-R-1374(EN), Ris National Laboratory, Roskilde, Denmark, 2002 28 [65] Gigu`re, P., Selig, M., and Tangler, J. Blade Design Trade-Os Using Lowe Lift Airfoils for Stall-Regulated HAWTs. In ASME/AIAA Wind Energy Symposium, Reno, Nevada, January 1999 28, 193, 196 [66] Jamieson, P. Evaluation of the Coning Rotor Concept. Garrad Hassan, Bristol, May 1996 28, 30, 34, 69, 125, 128, 129, 141, 144, 177, 180, 187, 194, 197, 208, 222, 227, 234, 244, C-12, D-1 [67] Raben, N., Jensen, F., Oye, S., Petersen, S., and Antoniou, I. Experiences and Results from Elkraft 1MW Wind Turbine. In European Wind Energy Conference, Nice, France, March 1-5 1999 30 [68] Rasmussen, F. and Kretz, A. Dynamics and Potentials for the Two-Bladed Teetering Rotor Concept. Ris National Laboratory, 1992 30 [69] Johnson, W. Helicopter Theory. Princeton University Press, New Jersey, 1980 30, 124, 125 [70] California Energy Commission. The Next Generation Wind Turbine Development Project. P500-02-031F, California Energy Commission, March 2002 30, 31, 234 [71] Eggers, A. J., Chaney, K., and Digurmarthi, R. An Exploratory Study of

251 Motion and Loads on Large Flap-Hinged Rotor Blades. In 43rd AIAA Aerospace Sciences Meeting and Exhibit, Reno, Nevada, January 2005 31, 35, 36 [72] Anderson, M. An Experimental and Theoretical Study of Horizontal-Axis Wind Turbines. PhD, Cambridge University, Cambridge, 1981 196, 201 [73] Drost, L. F. and Follings, F. J. Design and Construction of Innovative Flexible Rotor Systems. In European Wind Energy Conference, (pp. 212215). Hamburg, October 22-26 1984 30 [74] Rasmussen, F. and Petersen, J. A Soft Rotor Concept - Design, Verication and Potentials. In 1999 European Wind Energy Conference, Nice, France, March 1999 30, D-1 [75] Quarton, D. C. Monitoring and Analysis of a Carter 200/300 Wind Turbine: Final Report. Garrad Hassan, Bristol, UK, February 1997 30, 31, 234 [76] Kelley, N., Wright, A., and Osgood, R. Validation of a Model for a TwoBladed Flexible Rotor System: Progress to Date. In ASME/AIAA Wind Energy Symposium, Reno, Nevada, January 11-14 1999 31 [77] Pierce, K. Investigation of Load Reduction for a Variable Speed, Variable Pitch, and Variable Coning Wind Turbine. In WINDPOWER: American Wind Energy Association, (pp. 399406). 1997 31, 35, 125 [78] Larsen, T. J., Madsen, H. A., and Thomsen, K. Active load reduction using individual pitch, based on local blade ow measurements. In The Science of Making Torque from Wind, TU Delft, The Netherlands, 2004 32 [79] Bossanyi, E. A. Developments in Individual Blade Pitch Control. In The Science of Making Torque from Wind, TU Delft, The Netherlands, 2004 32 [80] Olsen, T., Lang, E., Hansen, A. C., Cheney, M. C., Quandt, G., VandenBosche, J., and Meyer, T. Low Wind Speed Turbine Project Conceptual Design Study: Advanced Independent Pitch Control. NREL/SR-500-36755, National Renewable Energy Laboratory, Golden, Colorado, December 2004 32, 117 [81] Klinger, F., Rinck, J., and Balzert, S. The Next Generation of Gearless Wind Turbines Goes Into Production. In European Wind Energy Conference, Madrid, Spain, June 16-19 2003 35, 142, 234 [82] Versteegh, C. J. A. Design of the Zephyros Z72 Wind Turbine with Emphasis

252

Bibliography on the Direct Drive PM Generator. In NORPIE 2004, NTNU Trondheim, Norway, June 14-16 2004 35, 234

[83] Engstrm, S., Hernns, B., Parkegren, C., and Waernulf, S. Development of NewGen - a New Type of Direct-Drive Generator. In European Wind Energy Conference, 2004 35 [84] Deering, K. US Patent 5,584,655: Rotor Device and Control for Wind Turbine. December 1994 35 [85] Deering, K. J. US Patent 5,660,527: Wind Turbine Rotor Blade Root End. August 26, 1997 1995 36 [86] Wang, T. and Coton, F. An Unsteady Aerodynamic Model for HAWT Performance Including Tower Shadow Eects. Wind Engineering, 23(5): pp. 255268, 1999 45 [87] Wang, T. and Coton, F. A High Resolution Tower Shadow Model for Downwind Wind Turbines. Journal of Wind Engineering and Industrial Aerodynamics, 89(10): pp. 873892, 2001 45, 168, 230, 239 [88] Mikkelsen, R., Sorensen, J. N., and Shen, W. Z. Modelling and Analyis of the Flow Field Around a Coned Rotor. Wind Energy, 4(3): pp. 121135, 2001 45, 54, 148 [89] Mikkelsen, R. Actuator Disk Methods Applied to Wind Turbines. PhD, Mechanical Engineering, Technical University of Denmark, Lyngby, 2003 45, 47, 48, 49, 89, 93, 148, 165, C-1 [90] Bertagnolio, F., Sorensen, N., Hansen, M., and Gaunaa, M. Aeroelastic Simulation of a Wind Turbine Airfoil by Coupling CFD and a Beam Element Method. In European Wind Energy Conference, Madrid, Spain, June 16-19 2003 45 [91] Coton, F. N., Wang, T., and Galbraith, R. An Examination of Key Aerodynamic Modelling Issues Raised by the NREL Blind Comparison. Wind Energy, 5(2-3): pp. 199212, 2002 46, 214 [92] van Kuik, G. A. M. An Inconsistency in the Actuator Disc Momentum Theory. Wind Energy, 7(1): pp. 919, 2004 46 [93] Rankine, W. J. M. On the Mechanical Principles of the Action of Propellers. Transactions of the Institution of Naval Architects, 6: pp. 1330, 1865 47 [94] Froude, R. E. On the Part Played in Propulsion by Dierences of Fluid

253 Pressure. Transactions of the Institution of Naval Architects, 30: pp. 390 405, 1889 47 [95] Betz, A. Das Maximum der Theoretisch Mglichen Ausntzung des Windes o durch Windmotoren. Zeitschrift Fr das Gesamte Turbinewesen, (pp. 307 309), 1920 47 [96] Froude, W. On the Elementary Relation Between Pitch, Slip and Propulsive Eciency. Transactions of the Institution of Naval Architects, 19: p. 47, 1878 47 [97] Drzewiecki, S. Mthode Pour la Dtermination des Elments Mcaniques des e e e e Propulseurs Hlicoidaux. Bulletin del Association Technique Maritime, 1892 e 47 [98] Glauert, H. Division L: Airplane Propellers. In Durand, W. F. (ed.), Aerodynamic Theory, volume 4. Springer, Berlin, 1935 47, 49, 51, 53, 153 [99] Glauert, H. The Elements of Aerofoil and Airscrew Theory. Cambridge University Press, Cambridge, UK, 2nd edition, 1947 47 [100] Chaney, K., Eggers, A. J. J., Moriarty, P. J., and Holley, W. E. Skewed Wake Induction Eects on Thrust Distribution on Small Wind Turbine Rotors. Journal of Solar Energy Engineering, 123(4): pp. 290295, 2001 47, 54, 75, 77, 78, 156 [101] Castles, W. and de Leeuw, J. H. The Normal Component of the Induced Velocity in the Vicinity of a Lifting Rotor and Some Examples of its Application. NACA-TR-1184, NACA, 1953 47 [102] Coleman, R. P., Feingold, A. M., and Stempin, C. W. Evaluation of the Induced-Velocity Field of an Idealized Helicopter Rotor. NACA-WR-L-126, NACA, Langley Field, Va, June 1945 54 [103] Heyson, H. H. and Katzo, S. Induced Velocities Near a Lifting Rotor with Nonuniform Disk Loading. NACA-TR-1319, NACA, 1956 47 [104] Lindenburg, C. Investigation into Rotor Blade Aerodynamics: Analysis of the Stationary Measurements on the UAE Phase-VI Rotor in the NASA-Ames Wind Tunnel. ECN-C03-025, ECN, 2003 51, 53, 60, 62, 63, 162 [105] Corten, G. Flow Separation on Wind Turbine Blades. PhD, Utrecht University, 2001 53, 62 [106] Madsen, P. and Rasmussen, F. The Inuence of Energy Conversion and

254

Bibliography Induction from Large Blade Deections. In European Wind Energy Conference, Nice, France, March 1-5 1999 54, 148

[107] Oye, S. A Simple Vortex Model. In Third IEA Symposium of Aerodynamics of Wind Turbines, 1990 54 [108] Szymendera, C. Computational Free Wake Analysis of a Helicopter Rotor. MSc, Pennsylvania State University, Department of Aerospace Engineering, 2002 56 [109] Tangler, J. and Kocurek, J. D. Wind Turbine Post-Stall Airfoil Performance Characteristics Guidelines for Blade-Element Momentum Methods. In 43rd AIAA Aerospace Sciences Meeting and Exhibit, Reno, Nevada, January 10-13 2005 61, 162, C-1 [110] Harris, F. D. Preliminary Study of Radial Flow Eects on Rotor Blades. The Journal of American Helicopter Society, 11(3): p. 121, 1966 62, 66, 67, 68, 69, 70 [111] Snel, H. Scaling Laws for the Boundary Layer Flow on Rotating Wind Turbine Blades. In Fourth IEA Symposion on the Aerodynamics of Wind Turbines, Rome, November 20-21 1990 62 [112] Snel, H., Houwink, R., and Piers, W. J. Sectional Prediction of 3D Eects for Separated Flow on rotating blades. In 18th European Rotorcraft Forum, Avignon, France, September 15-18 1992 62 [113] Lindenburg, C. Aeroelastic Analysis of the LMH64-5 Blade Concept. ECNC03-020, Energy Center of the Netherlands, 2003 63, 222, C-4 [114] Jones, R. T. Wing Theory. Princeton University Press, Oxford, 1990 65 [115] Hunton, L. W. Eects of Finite Span on the Section Characteristics of Two 45 degree Swept-Back Wings of Aspect Ratio 6. NACA, 1952 66, 68 [116] Furlong, G. C. and McHugh, J. G. A Summary and Analysis of the LowSpeed Longitudinal Characteristics of Swept Wings at High Reynolds Number. 1339, NACA, 1952 68, 69 [117] Leishman, J. G. Principles of Helicopter Aerodynamics. Cambridge Aerospace Series. Cambridge University Press, Cambridge, 2000 68, 71, 72 [118] McCormick, B. W. Aerodynamics, Aeronautics, and Flight Mechanics. John Wiley & Sons, Inc., New York, 1995 69

255 [119] Leishman, J. G. Modeling Sweep Eects on Dynamic Stall. Journal of the American Helicopter Society, 34: pp. 1829, 1989 71 [120] Barthelmie, R. J. Proceedings of the ENDOW Workshop Oshore Wakes: Measurements and Modelling. Riso-R-1326(EN), Ris National Laboratory, Roskilde, Denmark, March 2002 75 [121] Veers, P. S. Three-Dimensional Wind Simulation. Sandia National Laboratories, Albuquerque, New Mexico, March 1988 83 [122] Moriarty, P. J. and Hansen, A. C. AeroDyn Theory Manual. National Renewable Energy Laboratory, Golden, Colorado, January 2005 83, 85, 93 [123] Powles, S. J. R. The Eects of Tower Shadow on the Dynamics of Horizontal Axis Wind Turbines. Wind Engineering, 7(1): pp. pp. 2642, 1983 83 [124] Powles, S. J. R. Horizontal Axis Wind Turbines. PhD, University of Cambridge, 1985 [125] Bossanyi, E. A. GH Bladed Theory Manual. Garrad Hassan, Bristol, 2003 83, 93, 94, 95, 129, 180 [126] Leishman, J. G. Challenges in Modeling the Unsteady Aerodynamics of Wind Turbines. In 21st ASME Wind Energy Symposium and the 40th AIAA Aerospace Sciences Meeting, Reno, Nevada, 2002 88, 89 [127] Larsen, J. W. Nonlinear Dynamics of Wind Turbine Wings. PhD, Department of Civil Engineering, Faculty of Engineering and Science, Aalborg University, Denmark, 2005 89 [128] Koh, S. G. and Wood, D. H. Formulation of a Vortex Wake Model for Horizontal-Axis Wind Turbines. Wind Engineering, 15(4): pp. 196210, 1991 93, 154 [129] Tangler, J. and Bir, G. Evaluation of RCAS Inow Models for Wind Turbine Analysis. NREL/TP-500-35109, National Renewable Energy Laboratory, Golden, Colorado, February 2004 93 [130] Suzuki, A. Application of Dynamic Inow Theory to Wind Turbine Rotors. PhD, Department of Mechanical Engineering, University of Utah, Salt Lake City, 2000 93 [131] Burden, R. L. and Faires, J. D. Numerical Analysis. Brooks/Cole Publishing Company, New York, 6th edition, 1997 94, 137, 139

256

Bibliography

[132] Wagner, S., Bareiss, R., and Guidati, G. Wind Turbine Noise. Springer-Verlag Berlin and Heidelberg GmbH & Co. K, London, 1996 96, 98, 100, 102, 104, 106, 167, 169 [133] Moriarty, P. and Migliore, P. Semi-Empirical Aeroacoustic Noise Prediction Code for Wind Turbines. NREL/TP-500-34478, National Renewable Energy Laboratory, Golden, Colorado, December 2003 99 [134] Jonkman, J. M. and Buhl, M. L. New Developments for the NWTCs FAST Aeroelastic HAWT Simulator. In 42nd Aerospace Sciences Meeting and Exhibit Conference, Reno, Nevada, January 2004 100 [135] Migliore, P. and Oerlemans, S. Wind Tunnel Aeroacoustic Tests of Six Airfoils for Use on Small Wind Turbines. In AIAA Wind Energy Symposium, Reno, Nevada, January 5-8 2004 100 [136] Morris, P. J., Long, L. N., and Brentner, K. S. An Aeroacoustic Analysis of Wind Turbines. In 42nd AIAA Aerospace Sciences Meeting and Exhibit, Reno, Nevada, January 5-8 2004 101, 102, 103 [137] Arakawa, C., Fleig, O., Iida, M., and Shimooka, M. Numerical Approach for Noise Reduction of Wind Turbine Blade Tip with Earth Simulator. Journal of the Earth Simulator, 2, 2005 101 [138] Lighthill, M. J. On Sound Generated Aerodynamically. Proceedings of the Royal Societry London Series A, 211: pp. 564587, 1952 101 [139] Lowson, M. V. and Lowson, J. V. Noise Evaluation of Coning Rotor. 94/07, Flow Solutions Ltd., Bristol, September 1994 103, 229, 239 [140] Viterna, L. A. The NASA-LeRC Wind Turbine Sound Prediction Code. In Second DOE/NASA Wind Turbine Dynamics Workshop, Cleveland, Ohio, February 24-26 1981 103 [141] Barman, K., Dahlberg, J.-A., and Meijer, S. Measurement of the Tower Wake of the Swedish Prototype WECS Maglarp and Calculations of its Eect on Noise and Blade Loading. In European Wind Energy Conference, (pp. 5663). Hamburg, October 22-26 1984 103, 107, 167, 229, 230 [142] Beyer, W. H. CRC Standard Mathematical Tables. CRC Press, Boca Raton, FL, 28 edition, 1987 107 [143] DATAQ Instruments. A Closer Look At The WinDaq Derivative Algorithm. DATAQ Instruments, 2003 108

257 [144] Press, W. H., Flannery, B. P., Teukolsky, S. A., and Vetterling, W. T. Numerical Recipes in FORTRAN 77: The Art of Scientic Computing. Cambridge University Press, Cambridge, UK, 1992 108, 109 [145] Savitzky, A. and Golay, M. J. E. Smoothing and Dierentiation of Data by Simplied Least Squares Procedures. Analytical Chemistry, 36: pp. 1627 1639, 1964 108 [146] Gorry, P. A. General Least-Squares Smoothing and Dierentiation by the Convolution (Savitzky-Golay) Method. Analytical Chemistry, 62: pp. 570 573, 1990 109 [147] Randall, R. B. and Tech, B. Frequency Analysis. Brul and Kjr, Denmark, 3rd edition, 1979 109 [148] Dally, J. W., Riley, W. F., and McConnell, K. G. Instrumentation for Engineering Measurements. John Wiley & Sons, Inc., Toronto, 2nd edition, 1993 109, 111 [149] National Instruments. The Fundamentals of FFT-Based Signal Analysis and Measurement in LabVIEW and LabWindows/CVI. 2006. Retrieved from National Instruments http://www.ni.com/, May 2006 109 [150] Harris, F. J. On the Use of Windows for Harmonic Analysis with the Discrete Fourier Transform. Proceedings of the IEEE, 66(1), 1978 112 [151] Ugural, A. and Fenster, S. Advanced Strength and Applied Elasticity. Prentice Hall, New Jersey, 1995 114, 117, 171 [152] Crawford, C. An Integrated CAD Methodology Applied to Wind Turbine Optimization. MSc, MIT, 2003 114, B-1 [153] Bir, G. and Migliore, P. Preliminary Structural Design of Composite Blades for Two- and Three-Blade Rotors. NREL/TP-500-31486, National Renewable Energy Laboratory, Golden, Colorado, September 2004 117, 119 [154] Politis, E. Aeroelastic Stability Investigation of a Wind Turbine Blade by Coupling a 2D Navier-Sokes Solver and a Beam Element Method. In 2006 European Wind Energy Conference, Athens, Greece, 2006 117 [155] Malcolm, D. J. and Laird, D. L. Modeling of Blades as Equivalent Beams for Aeroelastic Analysis. In ASME/AIAA Wind Energy Symposium, Reno Nevada, 2003 117

258

Bibliography

[156] Bir, G. S. Computerized Method for Preliminary Structural Design of Composite Wind Turbine Blades. ASME Journal of Solar Energy Engineering, 123, 2001 119 [157] Rao, S. S. Mechanical Vibrations. Addison-Wesley Publishing Company, Reading, Massachusetts, third edition, 1995 129, 130 [158] Smith, C. B. and Wereley, N. M. Transient Analysis for Damping Identication in Rotating Composite Beams with Integral Damping Layers. Smart Materials and Structures, 5: pp. 540550, 1996 129 [159] Petersen, J. T., Madsen, H. A., Bjrck, A., Enevoldsen, P., ye, S., Ganander, o H., and Winkelaar, D. Prediction of Dynamic Loads and Induced Vibrations in Stall. Ris National Laboratory, May 1998 129, 223, 224, 225, 227 [160] Logan, D. L. A First Course in the Finite Element Method. PWS Publishing Company, Boston, 1992 130, 131, 132, 133, 134 [161] ANSYS Inc. ANSYS Inc. Theory Reference 6.1. ANSYS, Inc., Southpointe, PA, USA, 2002 132 [162] Zienkiewicz, O. C. and Taylor, R. E. The Finite Element Method. McGrawHill, London, 4th edition, 1993 134 [163] Pike, R. W. Optimization for Engineering Systems. Van Nostrand Reinhold, New York, 1986 138 [164] Connor, B., Leithead, W., and Jamieson, P. Control Design of the CONE 450 Wind Turbine. In Wind Energy Conversion, Proceedings of the 1996 18th British Wind Energy Association Conference, Exeter, UK, September 25-27 1996 141 [165] Dubois, M. R. Optimized Permanent Magnet Generator Topologies for DirectDrive Wind Turbines. PhD, Delft University, 2004 142, 143, 182 [166] Grauers, A. Design of Direct-Driven Permanent-magnet Generators for Wind Turbines. PhD, Chalmers University of Technology, School of Electrical and Computer Engineering, Gteborg, Sweden, 1996 142 [167] Poore, R. and Lettenmaier, T. Alternative Design Study Report: WindPACT Advanced Wind Turbine Drive Train Designs Study. NREL/SR-500-33196, National Renewable Energy Laboratory, Colorado, USA, August 2003 142, 145

259 [168] Wildi, T. Electrical Machines, Drives, and Power Systems. Prentice Hall, Pearson Education Inc., New Jersey, 2002 143 [169] Selmon, G. Electric Machines and Drives. Addison-Wesley Publishing Company Inc., USA, 1992 143 [170] Dubois, M., Polinder, H., and Ferreira, J. Comparison of Generator Topologies for Direct-Drive Wind Turbines. In Nordic Countries Power and Industrial Electronics Conference (NORPIE)2000, (pp. pp. 2226). Technical University of Delft, Electrical Power Processing Group, Aalborg, Denmark, June 2000 145 [171] Nygaard, T. A. Three-Dimensional Euler Flow-Field Computations Through a Wind Turbine Rotor. In European Wind Energy Conference, (pp. 117 120). Nice, France, March 1-5 1999 154 [172] Srensen, J. N., Shen, W. Z., and Munduate, X. Analysis of Wake States by a Full-eld Actuator Disc Model. Wind Energy, 1: pp. 7388, 1998 154 [173] Miller, R. H. The Aerodynamics and Dynamic Analysis of Horizontal Axis Wind Turbines. Journal of Wind Engineering and Industrial Aerodynamics, 15: pp. 329340, 1983 154 [174] Munro, D. H. The Production of Sound by Moving Objects. PhD, Physics, Massachusetts Institute of Technology, Cambridge, 1980 154 [175] Wood, D. H. On Wake Modelling at High Tip Speed Ratios. Wind Engineering, 16(5): pp. 291303, 1992 154 [176] Corten, G. P. Novel Views on the Extraction of Energy from Wind: Heat Generation and Terrain Concentration. In EWEC 2001, Copenhagen, 2001 154 [177] Hand, M. M., Simms, D. A., Fingersh, L. J., Jager, D. W., Cotrell, J. R., Schreck, S., and Larwood, S. M. Unsteady Aerodynamics Experiment Phase VI: Wind Tunnel Test Congurations and Available Data Campaigns. National Renewable Energy Laboratory, Golden, Colorado, 2001 161, 213, C-4, C-5 [178] Simms, D., Schreck, S., Hand, M., and Fingersh, L. J. NREL Unsteady Aerodynamics Experiment in the NASA-Ames Wind Tunnel: A Comparison of Predictions to Measurements. National Renewable Energy Laboratory, Golden, Colorado, 2001 162, C-4

260

Bibliography

[179] Gerber, B. S., Tangler, J. L., Duque, E. P. N., and Kocurek, J. D. Peak and Post-Peak Power Aerodynamics from Phase VI NASA Ames Wind Turbine Data. Journal of Solar Energy Engineering, 127, 2005 162 [180] TPI Composites Inc. Innovative Design Approaches for Large Wind Turbine Blades. SAND2003-0723, Sandia National Labs, 2003 192, 202 [181] Jamieson, P. and Rawlinson-Smith, R. High Lift Aerofoils for Horizontal Axis Wind Turbines. In European Union Wind Energy Conference, Goteborg, Sweden, 1996 192 [182] Martins, J. R. R. A., Sturdza, P., and Alonso, J. J. The Connection Between the Complex-Step Derivative Approximation and Algorithmic Dierentiation. In AIAA Conference, Reno, NV, 2001 195 [183] Belessis, M. A., Stamos, D. G., and Voutsinas, S. G. Investigation of the Capability of a Genetic Optimization Algorithm in Designing Wind Turbine Rotors. National Technical University of Athens, Fluids Section, Athens, Greece, 2005 196 [184] Gill, P. E., Murray, W., and Wright, M. H. Practical Optimization. Academic Press, London, 1981 196, 200 [185] Gano, S. E. Simulation-Based Design Using Variable Fidelity Optimization. PhD, Aerospace and Mechanical Engineering, University of Notre Dame, Notre Dame, Indiana, 2005 196 [186] Bulder, B. H. and Schepers, J. G. Numerical Optimization for Wind Turbine Design, Based on Aero-Elastic Analysis. Netherlands Energy Research Foundation ECN, Petten, 2001 196 [187] Samareh, J. A. A Survey of Shape Parameterization Techniques. In

CEAS/AIAA/ICASE/NASA Langley International Forum on Aeroelasticity and Structural Dynamics, Williamsburg, VA, June 1999 197 [188] Willcox, K. and Wakayama, S. Simultaneous Optimization of a MultipleAircraft Family. AIAA Journal of Aircraft, 40(4), 2003 200 [189] Selig, M. S. and Tangler, J. L. Development and Application of a Multipoint Inverse Design Method for Horizontal Axis Wind Turbines. Wind Engineering, 19(2): pp. 91105, 1995 201 [190] TPI Composites Inc. Innovative Design Approaches for Large Wind Turbine Blades: Final Report. SAND 2004-0074, Sandia National Labs, May 2004

261 202 [191] Gipe, P. AWT 26 Generates Noise Complaint. Year. Retrieved from http:// www.wind-works.org/articles/lg_awt26noise.html September 2006 229 [192] Thresher, R. W. and Dodge, D. M. Trends in the Evolution of Wind Turbine Generator Congurations and Systems. Wind Energy, 1(S1): pp. 7085, 1998 234 [193] Proven, G. and Derrick, A. Regulation and Application of the Proven 2kW Wind Turbine. In European Union Wind Energy Conference, Goteborg, Sweden, 1996 234, D-1 [194] Hansen, C. NWTC Design Codes: AirfoilPrep. 2005. Retrieved

from http://wind.nrel.gov/designcodes/preprocessors/airfoilprep/ December 2005 C-1 [195] Simms, D., Schreck, S., Hand, M., Fingersh, L., Cotrell, J., Pierce, K., and Robinson, M. Plans for Testing the NREL Unsteady Aerodynamics Experiment 10-m Diameter HAWT in the NASA Ames Wind Tunnel. National Renewable Energy Laboratory, Golden, Colorado, 1999 C-4 [196] Jonkman, J. M. Modelling of the UAE Wind Turbine for Renement of FAST AD. National Renewable Energy Laboratory, Golden, Colorado, 2003 C-4 [197] White, F. M. Fluid Mechanics. WCB/McGraw-Hill, Boston, 4th edition, 1999 C-7 [198] Bertagnolio, F., Sorensen, N., Johansen, J., and Fuglsang, P. Wind Turbine Airfoil Catalogue. Riso-R-1280(EN), Ris National Laboratory, Roskilde, Denmark, 2001 C-12

Appendices

Appendix

Geometric Transformations
A common notation is used for the 4x4 geometric transformation matrices used throughout. The geometric translation matrix 1 0 T (x, y, z) = 0 0 is dened as: 0 0 x 1 0 y 0 1 z 0 0 1

(A.1)

Matrices dening a rotation about the x, y and z axes are dened respectively as:

1 0 0 0 C S R x () = 0 S C 0 0 0 C S 0 S C 0 R y () = 0 0 1 0 0 0 C 0 S 0 1 0 R z () = S 0 C 0 0 0 where S = cos and C = cos . To transform a coordinate point x, y, z, 1 transformation order as:
T

0 0 0 1 0 0 0 1 0 0 0 1

(A.2a)

(A.2b)

(A.2c)

through a rotation about the x

axis and then about the z axis, the matrices must be premultiplied in reverse x y x = R z () R x () z 1

(A.3)

A vector x may be transformed by setting the fourth element of the vector to zero. Note that translations then have zero eect. A-1

Appendix

Cross Section Analysis


This appendix describes the computation of cross-sectional properties for a generalized composite beam section. It has been adapted from the authors previous work [152]. The basic idea is to map the coordinates of the triangle from the real space x, y to a convenient unit vector space u, v , as shown in B.1.

Figure B.1 Section integral coordinate transform The mapping function may be dened as: x = u = Ax By Ay By u v (B.1)

for an arbitrary vector x in x, y to u in u, v , both lying in the triangular integration region. The local and absolute centroid may be computed from: xc,local = A+B 3 xc = O + xc,local B-1 (B.2a) (B.2b)

The area Atri of the element, and an elemental area dA may be computed from: Atri = 1/2 A B dA = |dx dy| = |a0 a3 a1 a2 | dudv = dudv (B.3a) (B.3b)

The rst (S) and second (I) moments of area for triangular domain, computed in the real domain, may then be derived from the above transformation and quantities as: Sx = Sy = Ixx = Iyy = Ixy = ydA xdA y 2 dA x2 dA xydA = = = = By Ay + 6 6 Ax Bx + 6 6 + Atri (xc,y xc,local,y ) + Atri (xc,x xc,local,x ) + Atri x2 x2 c,y c,local,y (B.4a) (B.4b) (B.4c) (B.4d) (B.4e) (B.4f) (B.4g)

2 A2 By Ay By y + + 12 12 12

Ax Bx A2 B2 + x + x + Atri x2 x2 c,x c,local,x 12 12 12 Ax By + Bx Ay Ax Ay Bx By = + + + 24 12 12 Atri (xc,x xc,y xc,local,x xc,local,y )

Jz =

(x2 + y 2 )dA = Ixx + Iyy

The formulas are applied to each triangular element, and their contributions aggregated as modulus (EA, ES, EI, GJ) and density (A, S, J) weighted sectional quantities. The over-tilde indicate nal aggregate quantities. The nal sectional mass properties are then found from: mps = A xCG = [S y , S x ]/mps Izz = J mps |xCG |2 (B.5a) (B.5b) (B.5c)

where mps is the mass/unit-span of the section and the stiness properties from: xEA = [ES y , ES x ]/EA EIx = EI x EAx2 EA,y EIy = EI y EAx2 EA,x EIxy = EI xy EAxEA,x xEA,y GJ = GJ GA |xEA |2 B-2 (B.6a) (B.6b) (B.6c) (B.6d) (B.6e)

The angle and magnitude of the principle axes may be determined from: = 1/2 arctan 2EIxy EIy EIx (B.7a) (B.7b) (B.7c)

EIprin,1 = EIx cos2 + EIy sin2 2EIxy sin() cos() EIprin,2 = EIx sin2 + EIy cos2 2EIxy sin() cos()

The coordinate denition and nature of blade proles means that the rst principle axis is invariably chordwise, and the second edgewise. Given the total twist of the section , including pitch set angle, active pitch angle, and section twist, the OP and IP bending stinesses may be computed as: EIOP = EIIP EIx + EIy EIx + EIy + EIxy sin 2 2 2 cos 2 EIx + EIy EIx + EIy = + EIxy sin 2 2 2 cos 2 (B.8a) (B.8b)

B-3

Appendix

Reference Machine Specifications


This appendix contains the details for the reference machines used in this thesis. Where required, airfoil data has been extended in AOA using NRELs AirfoilPrep worksheet [194], which implements Viternas method [109] for post-stall, with a max cd of 1.29 and an

A of 10.

C.1 Tjaereborg Machine


The Tjaereborg machine was an operational turbine in Esbjerg, Denmark. It was built by ELSAM in 1987, and decommissioned in 2001. The relevant data for the rotor, as supplied by Mikkelsen [89] and used in his thesis, is given in Table C.1 and Fig. C.1. Table C.1 Tjaereborg parameters Parameter Rated power Rotor diameter Rotor position Hub length (Rhinge ) Nominal cone angle Pitch set angle (pitch,set ) Fixed rotation speed () Nacelle tilt Nacelle oset (o) Hub height Tower station height(diameter) Value 2 61 Upwind 1.46 0 0.5 22.1 3 6.81 61 0.0 (7.25) 28.0 (4.75) 56.0 (4.25) = 0.3 ww = 2.50 lw = 3.25 MW m m

RPM

m m m m m

Assumed tower wake properties

C-1

12 10

Chord Twist Thk

35 30 25 20 15 10 5 0

Chord (m), twist (deg)

6 4 2 0

10

Distance from root (m)

15

20

25

30

(a) Blade geometry parameters


24 22 20 18 16 14 12 10 8 6 4 2 0
pitch

Pitch (deg)

V (m/s)

12

16

20

24

(b) Operating prole

Figure C.1 Tjaereborg blade details

C-2

Thickness (%)

2.0 1.5 1.0

1.2 1.0 0.8

12 15 18 21 24

0.0 -0.5 -1.0

15 18 21 24 0 20 40 60

cd

cl

0.5

12

0.6 0.4 0.2 0.0

-20

AOA (deg)

-20

AOA (deg)

20

40

60

(c) Airfoil lift coecient (NACA 4412)

(d) Airfoil drag coecient (NACA 4412)

Figure C.1 Tjaereborg blade details (cont.)

C-3

C.2 NREL UAE Phase IV


The UAE was a long-running program conducted by the National Renewable Energy Laboratory (NREL) to provide experimental data to validate analysis tools. The rst phases of the program consisted of eld trials, but due to atmospheric turbulence creating inow anomalies, a controlled wind tunnel experiment was conducted in phase VI. The wind tunnel at the National Aeronautics and Space Administration (NASA) Ames Research Centre was used to enable testing of a large scale machine, thereby avoiding aerodynamic scaling diculties. The NASA Ames tunnel test section is 24.4 m x 36.6 m, allowing a rotor of nominal 10 m diameter to be tested without wake blockage, wind shear or large-scale inow turbulence. The set-up for the test itself is well documented [113, 177, 178, 195, 196], so only a brief description is included here and in Table C.2. The UAE machine is a heavily modied Grumman Wind Stream-33 built to allow exible conguration. It employs a 2-bladed rotor that can be rigid, teetering or apping and operate upwind or downwind with full-blade independent pitch control. The nominal blade has a tip radius of 5.029m, and can be tted with extended or smoke tips. The drive-train has a gearbox and constant-speed induction generator, but may be controlled via power electronics. Telemetry provision is quite extensive, including instrumentation for loads (blade, hub, drive train and tower strain gauges/load cells/tunnel balance) and conguration (pitch, azimuth, yaw angles, etc.). The blades themselves are instrumented with pressure taps and pitot-tube probes at xed locations, and the tunnel itself has various inow measurements (wind speed, temperature, pressure taps, etc.). The data sets available are pre-corrected for the inuence of the overhung rotor weight on the shaft strain gauge measurements used to compute power. An important note when using the raw data is to correct the power reading for dynamic imbalance of the rotor. Imbalance in the telemetry boom led to rotating imbalance and in turn a cyclic error in the power [177, p. 75]. The blade data shown in Fig. C.2 for the 5.029 m radius rotor (i.e. without extended tip). Hand et al. [177] provides tabular data for the extended-tip blade that must be interpolated to derive these proles. The pitch setting pitch,tip of the experiments refers to the out-of-plane angle at the (short) blade tip. The blade pitch setting pitch,set is therefore: pitch,set = pitch,tip twist,tip C-4 (C.1)

The data provided by NREL has a fairly large range in cl and cd 1 , at dierent Re and from testing facilities. The airfoil dataset used here is the S809 OSU data [177] with AOA data to 26 , for best Re t among the various sets available. The cd curve was smoothed around 10 and extended beyond the provided data range using AirfoilPrep. Table C.2 NREL UAE parameters Parameter Rated power Rotor diameter Hub length (Rhinge ) Fixed rotation speed () Value 12 11 0.508 72 kW m m RPM

Wake decit and surface pressure integrations are both used for cd

C-5

0.8 0.7 0.6

Chord Twist Thk

120 100 80 60 40 20 0

Chord (m)

0.5 0.4 0.3 0.2 0.1 0.0 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0

-20

Distance from root (m)


1.2 1.0 0.8

(a) Blade geometry parameters

cl, cd

0.6 0.4 0.2 0.0 cl cd 0 20 40 60

-20

-0.2

AOA (deg)

(b) S809 airfoil data

Figure C.2 NREL UAE blade details

C-6

Twist (deg), thickness (%)

C.3 1.5 MW Reference Machine (REF-1500)


The REF-1500 was supplied by GH as a representative model of a current state-ofthe-art MW class machine. It has a custom controller for variable speed operation with pitch control (PTF). The relevant parameters for the machine are given in Table C.3, and the blade details in Fig. C.3. The tower is linearly tapered and rotor upwind of the tower. The airfoil data has been modied slightly around = 16 for the high Re number data to avoid extremely sharp stall in the GH data. A 100% thick airfoil (circle) has also been dened, with zero lift and constant cd = 0.3 for a cylinder at Re >1e6 White [197, p. 298].1 Table C.3 REF-1500 parameters Parameter Rated electrical power No-load constant power loss Generator eciency Rotor diameter Hub length (Rhinge ) Pitch set angle (pitch,set ) Rotation speed () Hub height Hub oset Nominal shaft tilt Tower height Tower top/bottom diameter Value 1.5 15 95 70 1.75 1 14.421 84 3.3 5 82 2.823/5.663 MW kW % m m

RPM m m

m m

The original data used cd = 1.0.

C-7

3.0 2.5 2.0

Chord Twist Thk

120 100 80 60 40 20 0

1.5 1.0 0.5 0.0

10

Distance from root (m)

15

20

25

30

35

(a) Blade geometry parameters


1.E+10 1.E+09 EIflap EIedge mps pitch axis 10000

EIflap, EIedge (Nm2)

1.E+08 1.E+07 1.E+06 1.E+05

1000

100

10

Distance from root (m)

15

20

25

30

35

10

(b) Blade stiness and mass parameters

Figure C.3 REF-1500 details

C-8

Mass/unit length (kg/m), pitch axis (%c)

Twist (deg), thickness (%)

Chord (m)

30

RPM, pitch angle (deg)

25 20 15 10 5 0 0 5

RPM Pitch angle

10

V (m/s)

15

20

25

(c) Operating prole


2.0 1.5 1.0 2.0 1.5 1.0

cl, cd

0.0 -0.5 -1.0

cl, cd

0.5 13 cl 17 cl 21 cl 13 cd 17 cd 21 cd 0

0.5 0.0 -0.5 -1.0 13 cl 17 cl 21 cl 13 cd 17 cd 21 cd 0

-20

-1.5

AOA (deg)

20

40

60

-20

-1.5

AOA (deg)

20

40

60

(d) Airfoil coecients for Re = 2e6 (LS1)

(e) Airfoil coecients for Re = 4e6 (LS1)

Figure C.3 REF-1500 details (cont.)

C-9

C.4 GH Demo Machine


The GH Demo Machine is supplied with BLADEDTM as an example set of data. The data relevant to this thesis are given in Table C.4 and Fig. C.4. Table C.4 Demo machine parameters Parameter Rated power Rotor diameter Hub length (Rhinge ) Value 2.2 MW 80 m 1.25 m

C-10

14 12

Chord Twist Thk

120 100 80 60 40 20 0

Chord (m), twist (deg)

10 8 6 4 2 0

10

Distance from root (m)

15

20

25

30

35

(a) Blade geometry parameters


1.E+10 1.E+09 EIflap EIedge mps pitch axis 10000

EIflap, EIedge (Nm2)

1.E+08 1.E+07 1.E+06 1.E+05

1000

100

10

Distance from root (m)

15

20

25

30

35

10

(b) Blade stiness and mass parameters

Figure C.4 Demo machine details

C-11

Mass/unit length (kg/m), pitch axis (%c)

Thickness (%)

C.5 CONE-450 Concept


The details for the CONE-450 nal design are given in Table C.5 and Fig. C.5 [66]. The airfoil coecients in Fig. C.5(d) were not included in the original report, but are taken from another GH report using same airfoil [58]. This data is used for the FX blade sets in Part III. Table C.5 CONE-450 parameters Parameter Rated power Rotor diameter Hub length (Rhinge ) Blade mass First mass moment (about hinge) Second mass moment (about hinge) Value 450 23.166 2.0 641 4509 45108 kW m (unconed) m kg kgm kgm2

C.6 Coning Rotor Study


The NACA 63(2)-4XX airfoil used in Part III is taken from the Ris airfoil library [198] experimental data, at an Re of 1.5e6. The 17%21% data is from Abbott, and the 30% data from Velux. Some data has been smoothed to avoid sharp spikes, as shown in Fig. C.6.

C-12

2.5 2.0 1.5 1.0 0.5 0.0

Chord Twist Thk

25 20 15 10 5 0

Distance from root (m)

10

12

14

16

18

20

22

(a) Blade geometry parameters


1.E+02

Mass/unit length (kg/m)

1.E+01

1.E+00

Distance from root (m)

10

12

14

16

18

20

22

(b) Blade mass parameter

Figure C.5 CONE-450 details

C-13

Twist (deg), thickness (%)

Chord (m)

35 30 25

RPM

RPM

15 10 5 0 0 5 10 15 20 25

cl, cd

20

V (m/s)

-20

1.2 1.0 0.8 0.6 0.4 0.2 0.0 -0.2 -0.4 -0.6

1.6 1.4

cl cd 0 20 40 60

AOA (deg)

(c) Operating prole

(d) Airfoil coecients (FX66-17AII-182)

Figure C.5 CONE-450 details (cont.)

2.0 1.5 1.0

cl, cd

0.5 0.0 -0.5 -1.0 15 cl 18 cl 21 cl 30 cl 15 cd 18 cd 21 cd 30 cd 0

-20

-1.5

AOA (deg)

20

40

60

Figure C.6 Airfoil coecients (NACA 63-4XX)

C-14

Appendix

Historical Machines
Table D.1 has been compiled from a number of sources [9, 15, 18, 66, 74, 193].

D-1

Table D.1 Soft machine history

Date

Rating (kW)

Diameter (m)

# Blades

Orientationa Yawb Hubc

Power Controld

Coning

Machine 1941 1960s 1980s 1980s 1980s 1980s 1987 1250 100 5 2500 4000 250 3200 53.3 34 5 91.4 78 20 97.5 2 2 2 2 2 2 2 D D D U D D U A ? P A A P A IF T IF T Td FB T FSP FSP S PSP FSP S PSP

Smith-Putnam at Grandpas Knob1 Htter-Allgaier2 u Martin Ryle (Cambridge)3 Boeing Mod-24 KKRV Sweden/Hamilton Standard (US) WTS-45 Carter6 Boeing Mod-5B7

-3 /6 7 0 15 ? >0 -

Orientation: D = downwind, U = upwind Yaw: A = active, P = Passive c Hub: IF = individual apping, T = teetering, Td = teetering with d3, Tp = teeter pitch coupling, FB = ex-beam, R = rigid d Power Control: FSP = full-span pitch, PSP = partial-span pitch, Y = yaw control, S = stall, C = cone e Tower: Tr = truss, T = tubular, TG = guyed tubular 1 Flapping hinges adopted to combat gyroscopic loading when yawing 2 Low solidity, high aspect blades led to utter problems 3 Only rigid stops on apping blades; Tapered untwisted blades 4 First soft design using experience and simulation; Higher than expected cyclic loading from inadequate wind models 5 Soft tower; Spring mounted gearbox 6 Flexible spar instead of ap hinges; Next generation machine: Constant-coning instead of pitch-ap coupling to provide smoother power, PTS & variable speed 7 First variable speed machine using cycloconverter

Towere Tr TG T T T T T

D-2

Table D.1 Soft machine history (cont.) Rating (kW) Diamter (m) Orientationa

Yawb

Date

# Blades

Hubc

Power Controld

Coning

Machine 30 2000 450 0.615 600 12 ? 44 2.559 41/45 2 2 3 3 3/2 D U D D D P A P P P T T R IF FB S Y C, S S, C FSP,S,C

Ris Test Machine1 Gamma 602 CONE-4503 Proven4 MS4-6005

1991 1992 1994 1990 1997

5 0 /80 5 45 8 static

Highly instrumented test machine with variable springs and dampers used on the teeter DOF; Experiments concerned with free yaw performance and required free-teeter angle requirements Teetering hub permits high yaw rates facilitating reliance on yaw control for power limiting; Broad-range speed control; Soft tower; Low modulus blade material Hydraulic central actuator to control collective coning angle; Blades individually hinged, but essentially rigid due to collective coning mechanism Range of machine sizes; Constant-speed direct-drive generators; Zeebede exible inclined hinge (3 ) allows coning of blades, combining centrifugal force towards feather and pitch with cone angle towards stall Flex-beams allow 50% load reduction in high winds; full-span pitch towards stall; xed-speed; counter-balanced nacelle, laterally oset from tower to balance thrust and rotor torque component along tower from tilted nacelle (downwind end up); self-erecting; 2-bladed design for oshore; 45 m design for Class II sites

Towere T T TG T T

D-3

Table D.1 Soft machine history (cont.) Diamter (m)

Coning

Date

Rating (kW)

# Blades

Orientationa

Yawb

Hubc

Machine 1998 1999 2003 2003 250 15 500 275 40 12 33 30 2 2 2 2 D D U D A P A P IF T/IF Tp T FSP FSP FSP FSP

WTC1 Soft-Rotor Concept (Ris)2 Window 5003 Vergnet4

Power Controld

-5 /15 0 /90 > 0

Hydraulic dampers for semi-passive yaw; Hydraulically damped, independent blade apping Teetering hub with stiness; Blades not rigidly attached to teetered hub, but individually supported by hinge axis inclined to give 8 ap/edge coupling. A composite ex-beam across the hinge axis then provides individual blade apping stiness; No shear webs were used in the blades apart from the ex-beam; Blade pitching is unclear, but appears to envisaged as 15 with active stall control in future versions; Nacelle tilt exibility is provided Torque limiting gearbox; Novel pitch teeter coupling Tilt down for extreme storms (e.g. hurricanes)

Towere TG TG T TG1

D-4

Appendix

Material Specifications
These material values are taken from Grin [55], and are a result of test data and laminate calculations without safety factors. Unless otherwise stated, the materials are glass bres in an epoxy matrix.

E-1

Table E.1 Section layup material specications Density (kg/m3 )


ut uc

Name

Elastic Modulus E (GPa) 0.022 0.022 0.022 0.022 0.022 0.022 0.0105 0.0105 0.0105 0.0105 0.0105 0.0105 24.2 31.0 27.1 9.65 3.44 2.07 4.97 3.52 4.70 3.86 1.38 0.140 0.39 0.31 0.37 0.30 0.30 0.22

Torsional Modulus G (GPa)

Poissons Ratio -

Ultimate Strain Tensile Compressive

CDB340a A260b Spar cap mixc Random mat Gel coat Balsa 1700 1700 1700 1670 1230 144

E-2

Triaxial fabric, 25%:25%:50% distribution of +45 :-45 :0 bres b Uniaxial fabric c Alternating layers of CDB340 and A260 giving 70% uniaxial bres by mass

Anda mungkin juga menyukai