Anda di halaman 1dari 12

Ocean Engineering 156 (2018) 306–317

Contents lists available at ScienceDirect

Ocean Engineering
journal homepage: www.elsevier.com/locate/oceaneng

Wave interaction with a tethered buoy: SPH simulation and


experimental validation
David F. Gunn a, Murray Rudman a, *, Raymond C.Z. Cohen b
a
Department of Mechanical and Aerospace Engineering, Monash University, Clayton, VIC, Australia
b
CSIRO Data61, Clayton, VIC, Australia

A R T I C L E I N F O A B S T R A C T

Keywords: Understanding the interaction between surface waves and floating, moored structures is an important problem for
SPH the design of offshore structures. Experimental analysis has traditionally been used to study these problems, but
Validation data using computational modelling could potentially reduce the number of experimental iterations required. In this
Wave-structure interaction paper we investigate the use of Smoothed Particle Hydrodynamics as a method for modelling these problems and
Floating body thus reducing the number of experiments required. Validation data is obtained from generic experiments that
Moored body
contain the important features of wave-structure interaction. A moored spherical buoy oscillating both through
and across a free surface, as well as interacting with an incident sinusoidal wave train are discussed and com-
panion SPH simulations are run. We show here that SPH is capable of achieving accurate predictions of the buoy's
motion when interacting with the free surface with appropriate particle resolutions. An initial transient motion
was observed in the experiments as a long period oscillation of the buoy at its surge natural frequency. This
transient motion was also predicted by SPH, but at reduced amplitude. Overall SPH was determined to be a good
choice for modelling surface and structure interactions and we provide guidelines on appropriate particle reso-
lution for such interactions.

1. Introduction capable of moving with 6 degrees of freedom, their interaction with large
non-linear surface waves is difficult to computationally model, thus a
The impact of large waves on ships and offshore structures is an sophisticated modelling technique is required. A technique that is
important aspect of maritime design, and much work has been conducted capable of modelling these interactions is Smoothed Particle Hydrody-
to investigate methods of accurately determining the loads induced by namics (SPH) (Monaghan, 1992, 2005) and is the focus of this study.
such impacts (Clauss, 2002; Schellin and El Moctar, 2007; Van Paepegem One of the strengths of SPH is the natural modelling of fluid bound-
et al., 2011; Gui et al., ; Rudman and Cleary, 2013; Zhao et al., 2014; aries. The technique has been utilised for modelling dam-breaking
Rudman and Cleary, 2016). Presently, experimental testing of prototypes problems (Cummins et al., 2012), sloshing problems (Souto-Iglesias
remains the most reliable method for determining the feasibility of a et al., 2006), breaking surface waves (Khayyer et al., 2008, 2009), and
design to withstand significant wave impacts. However, repeated many others (Doring et al., 2004; Bouscasse et al., 2013; Basic et al.,
experimental testing and design iteration can be expensive to conduct as 2014). This is due to its meshless Lagrangian formulation that is inher-
each iteration needs to be prototyped. In contrast to experimental testing, ently able to model changing fluid boundaries associated with object
numerical simulations can be performed to narrow the range of design motion without the need for any special treatment, or remeshing, that
options, resulting in fewer experimental tests needing to be performed. mesh-based methods require. Despite the importance of surface wave
For numerical simulations to be worthwhile, the numerical technique interactions with moored, floating structures, there is a lack of published
used needs to be both accurate and efficient to run. studies that investigate this interaction, and there is little or no data
In all but shallow water, offshore structures are typically moored suitable for validating computational models.
using tethering cables and chains (for example weather buoys, Tension Le Touze et al. (Le Touze et al., 2010) investigated a large wave
Leg Platforms (TLP), Floating Production Storage and Offloading (FPSO) impact on an FPSO vessel, providing a comparison between SPH simu-
vessels and floating offshore wind turbines). Since these structures are lations and the experimental work of Yang et al. (2007). In their work

* Corresponding author.
E-mail address: Murray.Rudman@monash.edu (M. Rudman).

https://doi.org/10.1016/j.oceaneng.2018.03.001
Received 22 October 2017; Received in revised form 2 March 2018; Accepted 2 March 2018
Available online 16 March 2018
0029-8018/Crown Copyright © 2018 Published by Elsevier Ltd. All rights reserved.
D.F. Gunn et al. Ocean Engineering 156 (2018) 306–317

they found that the green water effects of the wave impact were
reasonably well modelled using SPH in 2D, but in 3D the results were in
error primarily due to the poorer particle resolution. However, the mo-
tion of the FPSO in their simulations was prescribed and the structure
may respond differently if allowed to dynamically respond to the
incoming wave - an inherently important requirement for any realistic Fig. 1. Schematic of the Wave Tank used. This section view is along the central
simulation technique developed to aid in the design process. plane of the tank. The red dot and axes represent the position of the coordinate
Water entry and exit validations have been performed for non- system origin. Not to scale. (For interpretation of the references to colour in this
tethered cylinders, boxes and wedges using SPH (Doring et al., 2004; figure legend, the reader is referred to the Web version of this article.)
Shao, 2009; Campbell et al., 2009; Basic et al., 2014), although these
studies have predominantly been 2D simulations. Bouscasse et al. (2013) first 20.5 m of which is the working section of the tank. A porous, wave
undertook a range of different 2D simulations of wave diffraction, and damping beach in placed into the tank, commencing at the end of the
complex free-body interactions with fluid and waves, showing good working section. The beach extends upwards at an angle of 18 and the
agreement with experimental data. Ren et al. (2015) also presented 2D end wall of the tank is located shortly after the beach has risen 2 m above
simulations of wave interactions with freely floating bodies with good the floor.
agreement with experimental data. Many of these studies were focussed For all experiments and simulations, the x-coordinate is defined in the
on developments of the SPH algorithm to improve boundary treatments lengthwise direction of the tank (towards the beach and away from the
and other aspects of the implementation, rather than on the physics of piston), the y-coordinate in the upwards vertical direction, and the z-
wave-structure interaction. When the wave is not 2D or the structure is coordinate across the tank (‘out of the page’ with respect to Fig. 1). The
not a quasi-2D object and instead has six degrees of freedom, 3D effects origin of the coordinate system is set so that x ¼ 0 is at the start of the
are an essential aspect to consider. working section, y ¼ 0 is at the floor of the working section, and z ¼ 0 is
Omidvar et al. (2013) proposed a variable particle mass SPH meth- at the mid-plane of the tank (z ¼ 0:5 corresponds to the side walls). The
odology and showed results for an untethered “Manchester Bobber” in a coordinate system origin is shown in Fig. 1 by the red axes. A water depth
Gaussian wave packet that yielded quite good comparison to experi- of 80 cm, measured in the working section, is used in all test cases.
mental measurements. Pan et al. (2016) undertook 3D SPH simulations A spherical, tethered buoy is used in the experiments performed in
of a quasi-2D experiment with two degrees of freedom and obtained good this study. It has a mass of 1.745 kg and radius of 101.5 mm. The buoy
results. They further simulated the interaction of a solitary wave on a was made from a hollow sphere of high density expanded polystyrene
floating tethered platform, however had no data against which to vali- (density of 29.5 kg/m3) with 40 mm thickness. To ensure that the centre
date this case. of mass was located below the centroid (guaranteeing a stable sitting
Rudman and Cleary (2013, 2016) investigated a fully 3D rogue wave orientation in the water) a 38 mm thick mild steel disk of 40 mm radius
impact on a moored TLP using SPH, but no validation for this interaction was placed in the lower half of the buoy as detailed in Fig. 2 (a). This steel
was presented and this remains a weakness of many publications in this disk provided the majority of the mass to the buoy and offset the centre of
area. mass to 27.0 mm below the centroid. Using a CAD software (SolidWorks)
Gunn et al. (2014) presented a preliminary study of the interaction representation of the buoy, the Ixx , Iyy , and Izz components of the moment
between a tethered spherical buoy and a fluid free surface, providing a of inertia matrix were found to be 1.7352, 1.7764, and 1.7352 respec-
comparison between experimental data and SPH simulations. Their study tively, and the off-diagonal moments were 0.
showed promising results for the use of SPH to model tethered objects, The buoy was painted with a black and white octant pattern (Fig. 2b)
but insufficient particle resolution meant that no quantitative conclu- to improve visibility in the motion capture process. Additional markers
sions could be drawn. were added to the white sections to provide additional tracking points for
To fill the gap in the open literature for studies of wave interaction the motion capture process, and thus more reliable processing. The de-
with floating, moored structures that are not quasi-2D interactions, the tails of this process are discussed in Section 2.2.
current study presents results of a series of experiments of the interaction Offshore structures in deep water can be tethered by cables that, due
between a free surface and a spherical buoy. This data is appropriate for to their length, are noticeably elastic. In order to represent a similar
validating any computational method, and its provision is the first aim of scenario to offshore structures, the buoy in the experiments was tethered
the study. Although appropriate for any simulation technique, here we to the tank floor using a cable (1 mm diameter) and a spring (stiffness k ¼
use it to compare to predictions from the SPH method that has been used 30:88 N/m) system shown in Fig. 3. A coil spring is required at the
in previous studies (Cummins et al., 2012; Rudman and Cleary, 2013, laboratory scale as the cable length is not sufficient to be noticeably
2016). This comparison comprises of two parts: first with the buoy elastic. A potential issue is that the coil spring could create additional
oscillating vertically across or horizontally along the free surface of a unknown damping when expanding or contracting underwater, which is
body of water, and second the dynamic response of the buoy to sinusoidal difficult to implement into the SPH model. Consequently a pulley system
wave trains with varying periods. The experiments are simulated in SPH was used near the tank floor that allowed the spring to be located above
and the comparison is used to determine the simulation parameters the water. The length of the cable and spring pre-tension are set so the
required for obtaining acceptably accurate predictions of the buoy dy- resting position of the buoy is half submerged. The resistance of the
namics. This determination is the second aim of the study. pulley system was negligible compared to the buoyancy forces on the
buoy and is neglected.
2. Methodology As mentioned, two different types of experiments were undertaken:
oscillation tests and wave-train interactions. Both vertical and horizontal
2.1. Experimental setup oscillation tests were performed. For oscillation tests in the vertical di-
rection, the spherical buoy was pulled under the surface so that top of the
The experiments performed in this study were conducted in the buoy was completely submerged by 15 mm. Once submerged to the
Monash Wave Tank. The tank is a 1 m wide flume that extends for correct depth, the cable was pressed against a fixed wedge above the
approximately 35 m, and a side view schematic is presented in Fig. 1. A tank, as shown in the schematic in Fig. 4. A mark on the cable (shown in
hydraulically driven, vertical-face, piston-type wave generator is present the figure as the red mark) identified the initial location of the point of
at the left end of the tank, which moves parallel to a 30 sloped floor. the wedge, ensuring a consistent release position. Once the submerged
Following the piston region, the tank floor inclines upwards by 0.4 m buoy was in a static equilibrium, the cable was released and the buoy
over 6 m. The tank floor is then flat for the remainder of the tank. The allowed to oscillate and eventually return to its static equilibrium

307
D.F. Gunn et al. Ocean Engineering 156 (2018) 306–317

Fig. 2. The spherical buoy used in the experiments. (a) Cross


section schematic of the buoy. The buoy and steel disk are
cylindrically symmetric about the vertical line of symmetry,
(b) The octant pattern painted on the buoy.

camera used in the experiment is calibrated using OpenCV (www.opencv.


org) to determine the intrinsic parameters of the cameras (such as the
focal length, principal point, and distortion factors). A chessboard
pattern, shown in Fig. 6, is used to perform this calibration. The intrinsic
parameters are then combined with landmarks in the view to determine
the cameras extrinsic properties (location and orientation) precisely.
Four Sony HDR-CX220E video cameras were used to record the buoy
motion in the experiments. The cameras are placed at four locations
around the buoy to allow a motion tracking technique to be used to
determine the 3D coordinates of the buoy's centroid. For the experiments
all cameras were set to the same quality settings, which included a frame
rate of 50 frames per second. During the experiments a red light (see
Fig. 5) is flashed in the field of view of all cameras, allowing the cameras
to be synchronised during post-processing.
For each camera the intrinsic and extrinsic properties that describe
the transformation from world coordinates ðxw ; yw ; zw Þ to pixel co-
ordinates ðu; vÞ need to be known in order to give the estimates for the
marker's position. The transformation between world and pixel co-
ordinates is given by the following equation:
Fig. 3. Schematic of the experimental setup with the floating tethered buoy at 2 3
2 3 xw
its resting position. Note that the figure is not to scale. uw 6 7
4 vw 5 ¼ C 6 yw 7 (1)
4 zw 5
w
1

where w is a scaling factor, C ¼ A½R T, A is the 3  3 intrinsic parameter


matrix (camera, lens, and other distortions), and R and T are the extrinsic
parameters describing the rotational and translational transformations
between world coordinates and 3D camera coordinates. The checker-
board pattern calibration was used to determine the matrix A for each
camera. The checkerboard is placed in front of the camera at different
positions and slopes, and each image is input into an OpenCV calibration
application, some example images of this are shown in Fig. 6.
Numerous landmarks are marked on the walls of the tank, whose
coordinates are known exactly. The cameras used for the experiments are
Fig. 4. The release mechanism for the Vertical Oscillation tests. The red mark placed so that there are as many landmarks in the view as possible to
on the cable is pressed against the point of the wedge to ensure that the release
ensure the extrinsic properties are more accurately obtained. The land-
height remains consistent between repetitions. (For interpretation of the refer-
marks are then identified in the camera view and used to determine the
ences to colour in this figure legend, the reader is referred to the Web version of
this article.)
½R T matrix using the OpenCV calibration application. Fig. 7 shows the
identification of the landmarks (the inside corners of the L shapes for the
landmarks partway up the wall, and where the edge between the red and
position. This method resulted in more consistent results than those
yellow tape meets the wall top for the top landmarks).
previously reported (Gunn et al., 2014).
In the motion capture process the pixel coordinates of a point can be
known for a single camera view, but this single view is not sufficient to
2.2. Motion tracking determine the points’ world coordinates. This is due the fact that each
image pixel can correspond to any point along a ray originating from the
Motion capture software (SwiMCap, developed at CSIRO) is used to camera, so it is impossible to determine exactly which world coordinate
measure the trajectory of the buoy in the experiments. The software is correct. In order to identify the location of a particular marker, it must
utilises multiple synchronised camera views to determine the 3D location be visible in at least two camera views and the lines of possible points
of an object in the fields of view. In the experiments the cameras are
must intersect. However, there are some potential errors that can be
located above the water surface and look down at the buoy, so no introduced in the calibration process and also when the user is identi-
refraction effects are present except for the lens distortion. Consequently
fying the ðu; vÞ coordinates (i.e. inaccurate clicking on the markers in the
the camera calibration for such cases could be performed ex situ. Each

308
D.F. Gunn et al. Ocean Engineering 156 (2018) 306–317

Fig. 5. Camera views from the four cameras used to


record the buoy motion. The red light used to syn-
chronise timing is seen in each image at the end of
the pole. (For interpretation of the references to
colour in this figure legend, the reader is referred to
the Web version of this article.)

Fig. 6. In the calibration method a chessboard


pattern is moved in front of the camera in a range of
positions and orientations. The calibration software
uses the distortion of the pattern to determine the
intrinsic properties of the camera.

software), thus a least squares method is used to estimate the world co- the error of the sphere fit was then calculated computationally.
ordinates of the point that is closest to the lines of possible solutions by
minimising the sum of the square of the distances from the world coor- 2.3. SPH methodology
dinate to the rays.
A total of 11 markers on the upper hemisphere of the buoy were In SPH, the fluid is represented by a collection of interpolation points,
tracked during the experiments: 3 markers in each of white octants, and typically referred to as “particles”, that replace the computational mesh
the 5 intersections of black and white octants. Fig. 8 shows an example present in traditional methods. The discretised equations of motion are
view of the buoy in one of the experiments, with 6 markers present in the solved for each of these particles, using a Lagrangian description of the
particular view. Since the markers are on the surface of a sphere of known equations of motion. Each particle is influenced by the neighbouring
radius, RT , the location of the sphere's centroid can be calculated by particles at any given time, but there is no explicit connectivity that keeps
defining a function: certain particles nearby for the entire simulation. The implementation
X used in this study is identical to that used in Rudman and Cleary (2013)
fE ðxÞ ¼ ðjx  xi j  RT Þ2 (2)
i
and by Cummins et al. (2012). Consequently, details of the equations,
boundary conditions, and other simulation details are not repeated here
where xi is calculated position of marker i. The function fE essentially and only a brief overview of the method is provided.
tries to fit a sphere of radius RT , centred at x, to the set of marker loca- General details of the SPH method are well described by numerous
tions and returns an estimate of the error. The centroid that minimised authors, including Monaghan, 1992, 2005, and Cleary (1998). There are
two main variants of SPH for incompressible flow simulation – Weakly

309
D.F. Gunn et al. Ocean Engineering 156 (2018) 306–317

Fig. 7. Locating the landmarks in the field of view.


The vertices of the pink lines indicate the location of
the landmark on the screen. The green lines are a
reprojection of the pink lines, giving a visual indica-
tion whether the calibration is accurate or not. A
zoomed view has been edited into this image showing
a landmark at ðx; y; zÞ ¼ ð0:864; 1:107;  0:5Þ. (For
interpretation of the references to colour in this figure
legend, the reader is referred to the Web version of
this article.)

Fig. 8. Identification of the markers on the surface of


the buoy. Six of the 11 markers are visible in this
frame and the user marks them with a circle. The
crosses represent the estimated marker location based
on this camera and other cameras where the marker
is visible and are included in views where the marker
is not present (e.g. the yellow marker is not visible in
this view but its estimate is included here). Only 8
identifier colours are available, so blue, green and red
are repeated.

Compressible SPH (WCSPH) which was the method originally proposed (Monaghan, 1994):
by Monaghan (1994) for incompressible flow and Incompressible SPH
dr i X 2mj    
(ISPH), first introduced by Cummins and Rudman (1999). In WCSPH an ¼ vi þ ε v  vi W r i  r j ; h (3)
artificial, stiff equation of state is used in which small changes in density dt ρi þ ρj j
result in large changes in pressure, hence forcing the flow to be almost !
incompressible. The main advantage of WCSPH is a simple algorithm dvi X Pi Pj  
¼ Fi  mj 2 þ 2 þ Πij rW r i  r j ; h (4)
which relates pressure and density via the equation of state function. The dt j
ρi ρj
main disadvantage of WCSPH is that limiting density fluctuations to less
pffiffiffiffiffiffiffiffiffiffiffiffi
than ε times the fluid density requires the sound speed (cs ¼ ∂P=∂ρ) to d ρi X    
pffiffiffi ¼ mj vi  vj ⋅rW r i  r j ; h
be of order 1= ε times the typical velocity of the flow. The Courant dt
(5)
j
time-step restriction applies also to the sound speed (i.e. pressure waves
must be resolved by the method). Thus to enforce density fluctuations of
where 0  ε  1 is the XSPH term introduced by Monaghan (1992). This
less than 1%, the sound speed must be 10 times higher than the flow
keeps the particles orderly in the absence of viscosity and ε is chosen as
speed, meaning the time-step is ten times more restrictive than a truly
0.5 in this research. F i is the sum of the external forces on particle i, Pi is
incompressible method. In contrast, ISPH enforces incompressibility in a
the pressure of particle i, and Πij is the viscous term given by
more traditional way via a Poisson pressure equation (PPE) for the
pressure (i.e. there is no equation of state). Hence the Courant time-step is   
ξ 4μi μj vi  vj ⋅ r i  r j
based purely on the flow speed allowing order of magnitude larger Πij ¼ ⋅ ⋅  (6)
ρi ρj μi þ μj r i  r j j2 þ β2
time-steps. However the algorithmic complexity of ISPH is significant
and the implicit solution of the PPE is computationally quite expensive.
where ξ ¼ 5:75 for the Wendland kernel, and β is a small parameter that
In our experience ISPH does not offer significant time savings and the  
ensures singularities are avoided when r i  r j  ¼ 0 (typically β ¼ 0:01δx
small density fluctuations inherent in WCSPH do not adversely affect the
where δx is the mean particle spacing). The pressure terms (Pi and Pj ) in
quality of the resulting solution. Consequently WCSPH has been chosen
Eq. (4) are given by the weakly compressible equation of state:
here as the simulation method.
In this paper, an improved Euler time-stepping scheme is used to solve  γ
ρi
the equations of motion in SPH, including a mass continuity equation Pi ¼ P0 1 (7)
ρ0

310
D.F. Gunn et al. Ocean Engineering 156 (2018) 306–317

where γ is chosen to equal 7 (Cleary, 1998), ρ0 is the reference density 0.2 m above the tank floor to the still water level in the y  direction and
(chosen as 1000 kg/m3 for water), and P0 is the overall dynamic pressure its length in the x  direction depends on the wavelength. In all cases, an
scale given by unforced region in the x direction of approximately 1.5 m was left to
allow the buoy to react freely to incoming waves. The total length of the
γP0 SPH domain in the x  direction is set to be an integer number of
¼ c2s (8)
ρ0 wavelengths, ensuring that waves reentering the wavemaker region
(after passing through the periodic boundary) are in phase with the
where cs is the speed of sound. desired velocity field and require less artificial forcing.
There are a number of different ways of implementing solid object This momentum source approach is conceptually similar to the one
boundary conditions in SPH, and this area has received attention in a discussed by Liu et al. (2015), however it is intuitively simpler. Instead of
number of recent studies. Bouscasse et al. (2013) used a ghost particle basing the forcing on the temporal wave amplitude function, it is based
approach (i.e. several rows of fictitious particles inside the boundary) on on the velocity deviation and unlike that of (Liu et al., 2015) can handle
which ghost particle pressures were set via the Neumann pressure con- any arbitrary 3D wave form provided an analytic expression for the ve-
dition and ghost particle velocities calculated from neighbouring fluid locity field is available. This, of course, is not always the case, but for the
particles to satisfy both the divergence-free condition (in part of the tests discussed in this study it is available. Because the domain length
calculation) and the no-slip condition in the viscous sub-step. Zheng et al. must be an integer number of wavelengths to accommodate periodicity,
(Zheng et al., 2018) used ISPH and the mirror particle approach, however this approach cannot be used to handle more general wave spectra.
replaced the PPE on the boundary with an explicit expression for the However the periodicity means that there is no need of a wave damper,
pressure gradient. This resulted in more accurate pressure estimates on and the waves that pass out one end of the domain and recycle back into
the surface of a submerged object although similar estimates of motions the wave maker region are rapidly forced back to the desired profile.
to those obtained from a more traditional mirror particle approach except Calibration tests of the wavemaker, conducted in the Ph.D. thesis of
for the puzzling result of different motion periods. Gunn (see Section 4.2.3 of Gunn, 2017), showed that the amplitude of the
Here we use the repulsive forcing approach originally given in generated wave is sensitive to a number of input parameters, specifically
(Monaghan, 1994) and more recently used in (Omidvar et al., 2013) the permeability factor K (see Rudman and Cleary, 2016), and the di-
where it was found to give better results than the particle boundary mensions of the wavemaker region. Although effort was made to ensure
forcing method of Kajtar and Monaghan (2009). When used with that the amplitude of the generated waves match those of the experi-
WCSPH, we have found this boundary technique provides results for the ments, an error of up to 5% was observed. This discrepancy is accounted
integrated forces on the object surface that are equivalently good to more for in the analysis described below.
complex techniques, even though the individual particle pressures can be
noisy. This apparent contradiction arises because forces on the object are 3. Results: oscillation tests
determined from pressure gradients that in our experience are
well-estimated using the repulsive forcing method which is significantly The first validation study considers the behaviour of a tethered
simpler to implement in practice than ghost and mirror particles. spherical buoy released from a non-equilibrium position that subse-
quently oscillates near an otherwise flat free surface. Experiments are
2.4. Wave generation performed where the buoy described in Section 2.1 is displaced either
vertically or horizontally and released. These experiments are then
In the second part of this study, surface waves will be produced using reproduced in a number of SPH simulation to investigate the capability of
SPH and compared to experimentally produced waves. However, repli- SPH to predict the buoyancy, drag, and spring forces on the buoy, and to
cating the entire wave flume (in 3D) using SPH requires an infeasibly understand the requirements for obtaining accurate results.
large number of particles, including large numbers to approximate parts
of the flume far away from the surface interaction with the buoy which is 3.1. Vertical oscillation of the buoy
the only region of interest. To reduce the number of particles required,
the domain was shortened with boundary conditions in the x-direction The vertical oscillation experiment in which the buoy was initially
assumed to be periodic. A spatially varying momentum source (called submerged by 15 mm and released was performed six times and an
here a “wavemaker region”) was utilised to generate the wave. The average trajectory calculated. The runs were quite consistent, with each
general methodology for the wavemaker region is described in Rudman trajectory deviating less than 3% of the average trajectory. The natural
and Cleary (2016), but it is used here in a different manner. The wave- frequency of this system is difficult to theoretically predict, so it was
maker region remains stationary and applies a force to fluid particles measured by finding the average time between consecutive peaks and
inside its region of influence that aims to ensure their velocity matches a consecutive troughs. It was found that the period of oscillation was
specified velocity field. This velocity field can be any valid fluid velocity approximately 0.6 s, corresponding to a natural frequency of 1.67 Hz.
field, but in this paper we use the intermediate depth Airy wave solution The mean experimental trajectory is shown as the solid black line in
(Young, 1999): Fig. 9.
The experiments were reproduced using SPH simulations with
agk cosh ky
vx ¼ sinψ (9) different particle spacings: 10 mm, 5 mm, and 2.5 mm, equivalent to
ω cosh kd approximately 20, 40, and 80 particle spacings across the buoy diameter.
The buoy was placed at the same initial depth as the experiments with the
agk sinh ky
vy ¼ cosψ (10) same spring stiffness and initial spring extension. The simulations were
ω cosh kd
performed in a 1 m  1 m wide tank in order to reduce the total number
of particles, and hence computational expense. Fig. 9 shows a comparison
vz ¼ 0: (11)
between the vertical trajectories of the centre of the sphere in these
Here a is the wave amplitude (mean-to-peak), g is gravitational ac- simulations and the experimental mean trajectory.
celeration, k is the wave number ( ¼ 2π =λ), ω is the angular frequency ( ¼ With a particle spacing of 10 mm the agreement is shown to be poor.
2π =τ), d is the depth to the mean surface level, and ψ ¼ kx  ωt. The The amplitude of the first oscillation is in error by 35%, and the period
forcing is an increasing function of the difference between a particle's has approximately 15% error. Reducing the particle spacing to 5 mm
velocity and the desired velocity at that spatial location and time. (also shown in Fig. 9) improves the period agreement significantly (to
The wavemaker region spans the entire z  dimension, extends from within the experimental variation) and the amplitude error decreases to

311
D.F. Gunn et al. Ocean Engineering 156 (2018) 306–317

Fig. 9. Buoy displacement in the vertical direction


for the vertical oscillation test.

15%. Further reduction to 2.5 mm shows that the amplitude is within 2% An average period was estimated similarly to the vertical oscillation test
and the period remains within experimental variation. However, each of and was found to be 5.2 s, corresponding to a frequency of oscillation of
the simulations shows an increased damping of the motion after the first 0.19 Hz. The experimental data is shown as the black curve in Fig. 10.
oscillation that is greater than that measured experimentally. An un- Two SPH simulations were performed, one with 10 mm particle
published CSIRO report has suggested that particle disorder in the vi- spacing, and another with 5 mm particle spacing. In these simulations a
cinity of a moving object can dissipate significant amounts of energy in domain of length 2 m, width 1 m, and fluid depth of 0.8 m was modelled.
SPH simulations. It is hypothesised that this is the likely the cause of the Since the domain has been doubled in length compared to the vertical
additional damping observed in these simulations. However further oscillation tests, there are double the number of SPH particles. Addi-
focussed study is needed to prove this hypothesis. tionally the tests here require simulation times of at least 12 s of simu-
It is clear that improving the particle resolution gives significantly lation time (nearly 5 times as long as the previous case), so the
better results. However, the 2.5 mm particle spacing simulation (in which simulations require approximately 10 times as much computational time
the buoy diameter is approximately 80 particles) required in excess of 50 to run. For these reasons, the 2.5 mm particle spacing was omitted from
million particles for just this small test case. This simulation required this test case.
over 11 weeks of walltime using 12 processors (with an OpenMP The mean trajectory in the experiments and the simulated trajectories
implementation) and 48 GB of memory on a Dual Xeon 8-core E5-2650 using SPH are shown in Fig. 10. For a mean particle spacing of 10 mm the
Compute Node. The computational expense at such resolution is not motion is rapidly damped indicating that, like the vertical tests, the
feasible for most applications, and larger simulations become impractical damping is over predicted in SPH. For the 5 mm particle spacing simu-
even for research purposes until computing hardware developments have lation the amplitude of the initial oscillation is in very good agreement,
been made or more massively parallel implementations of the SPH but the amplitude of the first trough is approximately 15% in error. The
method are developed. In summary, if an object is discretised with 40 period of the first full oscillation is under-predicted by around 6%. The
SPH particles across the key dimension, for an unforced, tethered oscil- magnitude of the errors resulting from this damping is consistent with the
lation we expect good amplitude estimates, although errors of order 15% errors observed in the vertical oscillation tests for the same particle
in frequency predictions. spacing. If the 2.5 mm particle spacing case were feasible to perform, it is
likely that the errors would improve to less than 2%, within the experi-
3.2. Horizontal oscillation of the buoy mental measurement error, as was observed in the vertical oscillation
case.
In the second oscillation test, the buoy was displaced horizontally
rather than vertically. Using the same experimental setup, the buoy was 4. Wave train tests
pulled 0.5 m horizontally across the surface of the water and kept half-
submerged. The buoy was then released and allowed to freely return to While the simple oscillation tests are useful for determining how well
its equilibrium position. The release point in this case was not as SPH can model the interaction with the fluid surface, the long term goal
consistent as in the vertical release experiments, and the mean starting of modelling the interaction of a ship with a rogue wave requires that the
position was measured from the video to be 50.82.0 cm from its equi- interaction between simple waves and the buoy be validated. To perform
librium position. As in the vertical oscillation tests, the mean trajectory of these validations and interactions with the buoy, the Monash wave tank
the buoy was measured across 6 runs. The variability of the starting was used to generate single frequency wave trains with periods of either
position caused the amplitude of oscillation to vary by around cm, but the 1, 2, or 3 s in water of 80 cm depth which are then compared with SPH
effect on the timing of the peaks and zero crossing times was negligible. simulations. The 1 s period wave is chosen as it provides waves that are

Fig. 10. Buoy displacement in the horizontal direc-


tion for the horizontal oscillation test.

312
D.F. Gunn et al. Ocean Engineering 156 (2018) 306–317

classified as deep water waves (kd > π ), while the 2 and 3 s period waves ζ ¼ a sinðkx  ωtÞ (13b)
provide intermediate depth waves (π =10 < kd  π ). A shallow water
For small values of kd, tanh kd  kd ¼ 2π d=λ. Consequently we
wave would require a 16 m wavelength (with a period of 5.8 s) which
choose to non-dimensionalise the measured trajectories by the co-
requires an infeasibly long piston stroke in the tank and a 16 m long SPH
efficients of the cos and sin in Eq. (13):
domain that requires too many particles to be practicle. The wave trains
interact with the buoy (Fig. 3) located 1.5 m from the piston end of the ðx  xe Þd y  ye t  t0
working section (see Fig. 1). The same spring is used (stiffness of x ¼ y ¼ t ¼ (14)
aλ a τ
30.88 N/s) as in the previous experiments, and the untensioned length is
set to ensure the buoy is half submerged at rest. The motion tracking where ðxe ; ye Þ is the resting position of the buoy, a is the amplitude of the
technique is applied to measure the response of the buoy to the waves. incident wave train, τ is the wave period, and t0 is the time that the first
fully developed crest reaches the buoy.
4.1. Non-dimensionalisation
4.2. Waves with 1 s period
Young (1999) gives a formulation of the trajectory of fluid particles
based on linear wave theory. The position of a fluid particle, ðχ ; ζÞ, The first wave train interaction considered is between the tethered
measured from its mean position ðx; yÞ is given as spherical buoy and waves with a period of 1 s. At the depth used in the
study, the wavelength was 1.556 m, which classified the waves as deep
agk cosh ky
χ¼ cosðkx  ωtÞ (12a) water waves. Experimental data for heave and surge motion of the buoy's
ω2 cosh kd
centre of mass are given as the black lines in Figs. 12 and 13.
agk sinh ky The SPH domain used in the simulations was 4.669 m, or 3 wave-
ζ¼ sinðkx  ωtÞ (12b) lengths, in the x-direction, and the wavemaker region was set to be 3.0 m
ω2 cosh kd
long. Two simulations were performed with different mean particle
where a is the mean-to-peak wave amplitude. Using the dispersion spacings: 10 mm requiring nearly 4 million particles, and 5 mm requiring
relationship, ω2 ¼ gk tanh kd, and measuring particles on the surface (i.e. approximately 30 million particles. Figs. 12 and 13 also show compari-
kd ¼ ky), these equations become sons of the heave and surge motion of the buoy's centre of mass. The
waves generated in the experiment had an amplitude of 5.8 cm. Although
a
χ¼ cosðkx  ωtÞ (13a) the amplitude parameter in the wavemaker region was set to match the
tanh kd experimental waves, the wavemaker error discussed earlier caused the
resulting waves to have amplitudes of approximately 5.5 cm. Thus, the

Fig. 11. Visualisation of the simulated buoy in a 1 s


wave train with 10 mm particle spacing. The figures
show a slice through the midplane of the tank. The
particles on the left have been coloured by their
speed in the x-direction, and the particles on the right
have been coloured by their pressure.

313
D.F. Gunn et al. Ocean Engineering 156 (2018) 306–317

non-dimensionalisation in Eqn. (14) uses the actual amplitude of the from vertical. The tension in the cable at this mean position is 4.535 N
waves and not the specified amplitude. A visualisation of the fluid flow with 0.566 N in the horizontal component. Since the buoy typically
around the buoy is shown in Fig. 11. moves in the direction of the wave when ascending the front of an
For the heave motion in Fig. 12, the amplitude of the simulated incoming wave, and opposite to the wave when descending down the
motion with 10 mm particle spacing is underpredicted by 15% over the back, it experiences the increasing frontal slope of the wave for over half
first two periods, increasing to 25% underprediction afterwards. Addi- of the wave period. The slope causes the buoyancy force on the buoy to
tionally, after t  4, a phase error of around 15% is introduced. For the not be vertical, and since the buoy observes the front slope more than the
5 mm simulation, the amplitude of motion is in better agreement with the back slope there will be a net horizontal buoyancy force on the buoy.
experiments, but some phase error is still observed. This phase error is Since the vertical buoyancy force on the buoy is 21.48 N, a net horizontal
likely to be due to the difference in wave generation mechanisms be- buoyancy of 0.566 N is not unreasonable.
tween experiments and simulations. A frequency analysis of the 1 s period case with 5.8 cm amplitude is
In the surge motion in Fig. 13, however, there is a long period performed over the full time series and shown in Fig. 15. The surge
modulation clearly evident in both experimental and simulation results, frequency spectra, shown in Fig. 15a, shows a large peak at 1 Hz, cor-
although more noticeable in the 5 mm simulation. To investigate the responding to the frequency of the incident wave train. There is also a
nature of this long period modulation, the experimental results are shown peak at approximately 0.19 Hz, which is the natural frequency of oscil-
in Fig. 14, over a longer period of time. The buoy trajectory shows that lation that was observed in the horizontal oscillation tests that were
this modulation significantly reduces in amplitude after a few oscilla- performed earlier (shown as the vertical dashed blue line in the figure).
tions, indicating that the modulation is an initial transient motion. This result indicates the modulation is an initial transient motion of the
In addition to the modulation, the surge position of the buoy does not buoy, caused by the initial few waves displacing the buoy horizontally
average to zero. Since x ¼ 0 corresponds to the buoy's initial resting with subsequent modulation coming from the buoy attempting to return
position, this indicates that there is a net, non-zero force on the buoy that to its horizontal equilibrium position. The heave frequency spectra, in
is offsetting it from its resting position. For the 5.8 cm amplitude wave Fig. 15b, shows the same large peak at the wave train frequency, however
case the buoy's mean position is located at x ¼ 0:5, which corresponds there is no peak at the natural frequency of vertical oscillation (1.67 Hz).
to a mean displacement of 8.86 cm and a cable angle of approximately 7 Small peaks are also observed at 2 and 3 Hz in both frequency spectra,

Fig. 12. Non-dimensionalised heave motion of the


buoy in the experiments and simulations with a 1 s
period wave train.

Fig. 13. Non-dimensionalised surge motion of the


buoy in the experiments and simulations with a 1 s
period wave train.

Fig. 14. The full non-dimensionalised time series of


the wave train case with 1 s period.

314
D.F. Gunn et al. Ocean Engineering 156 (2018) 306–317

Fig. 15. Frequency spectra of the 1s wavetrain case


with amplitude 5.8 cm in the (a) surge direction and
(b) heave direction. The vertical dashed blue lines
indicates the respective natural frequency obtained
from the oscillation tests performed in Section 3.

however these are resonance frequencies of the wave train. 12%. Conversely, the heave motion, shown in Fig. 19, is more accurately
With the knowledge that the modulation is an initial transient mo- predicted, with an error of 12% in amplitude and less than 1% error in
tion, it can be concluded that it is more heavily damped in the SPH phase.
simulations, consistent with the additional damping observed in the The simulated wave train interactions with the spherical buoy here
horizontal oscillation cases. In the 5 mm simulations, the damping of indicate that the buoy's heave trajectory is more accurately predicted in
these motions is less, and in Fig. 13 the transient motion is in better SPH than the surge trajectory. The primary force that determines the
agreement with the experiments, with an error of approximately 10%. vertical position is the interplay between buoyancy and weight of the
Consequently, the disagreements that SPH has with the experimental buoy, which is reasonably accurately modelled using SPH. In the surge
results is due to the same damping that was observed earlier. motion, the force on the buoy is hydrodynamic and primarily arises from
the drag from the fluid. The first wave that interacts with the buoy im-
poses a transient response from the buoy as it is offset from the equi-
4.3. Waves with 2 and 3 s period
librium position. The horizontal oscillation case showed that SPH does
not accurately predict the initial transient motion at coarse particle
The interaction between the buoy and wave trains with periods of 2
spacings, and the surge trajectories of the buoy responding to the 2 and
and 3 s were similarly generated experimentally. In the experiments, the
3 s period waves also shows that the first peak is smaller than the
2 and 3 s waves had amplitudes of 4.0 cm and 7.1 cm, and wavelengths of
following oscillations. For increased resolutions (40–50 particles across
4.85 m and 7.9 m respectively. The experimentally measured surge and
the buoy diameter), better results are expected in line with the results for
heave motion is shown as the solid black lines in Figs. 16 and 17 for the
the initial transient in 1 s case. Clearly, particle resolution in SPH plays an
2 s wave and Figs. 18 and 19 for the 3 s wave. The start-up transient surge
important role in accurately predicting the initial transience of the buoy.
modulation observed in the 1 s wave was not observed experimentally or
At a particle spacing of 5 mm, or 1/40th of the buoy diameter, the tran-
numerically with the 2 or 3 s wave.
sient motion is reasonably predicted (within 10% error); however, if a
SPH simulations were run and similarly to the 1 s wave test, although
coarser particle spacing of 10 mm, or 1/20th of the buoy diameter must
the simulated waves were specified to have matching amplitudes, the
be used then errors of 20–30% can be expected in the transient motion of
resulting waves were smaller with amplitudes of 3.8 cm and 6.8 cm
the buoy.
respectively. Due to the large number of particles required to simulate
these cases with 5 mm particle spacing, only 10 mm simulations were
undertaken. 4.4. Second order stokes theory
A comparison between simulation and experiments for the 2 s wave is
shown in Fig. 16 for surge and Fig. 17 for heave motion of the buoy's Because the waves used above are not strictly first order waves, in an
centre of mass. The amplitude of the surge motion is underpredicted by attempt to improve the SPH predictions without refining the particle
25% with a 5% phase error. In contrast, the heave motion is in reasonable spacing, the velocity field was modified to follow 2nd Order Stokes
agreement with the experiments with an amplitude error of 15% and 2% theory. This change was observed to make small improvements to the
phase error. amplitude of the buoy's predicted motion, particularly in the 3 s period
Finally, waves with a period of 3 s are simulated in SPH and compared case, but no significant improvement of the phase error was noticed. For
to the experimental results. The amplitude of the surge motion, shown in the 2 s period case, the second order terms was 4 orders of magnitude
Fig. 18, is underpredicted by 21%, and the trajectory is out of phase by smaller than the first order terms and the observed predictions were only

315
D.F. Gunn et al. Ocean Engineering 156 (2018) 306–317

Fig. 16. Non-dimensionalised surge motion of the


buoy in the experiments and simulations with a 2 s
period wave train.

Fig. 17. Non-dimensionalised heave motion of the


buoy in the experiments and simulations with a 2 s
period wave train.

Fig. 18. Non-dimensionalised surge motion of the


buoy in the experiments and simulations with a 3 s
period wave train.

Fig. 19. Non-dimensionalised heave motion of the


buoy in the experiments and simulations with a 3 s
period wave train.

1% different from the previous simulations. The 1 s period cases studied interaction of tethered floating structures with ocean waves. This
had second order velocity terms 10 orders of magnitude smaller than the experimental data can also be used in other studies and is a potentially
first order, so the Stokes theory velocity field makes no difference to the valuable resource for others seeking to validate computational fluid dy-
predicted buoy motion. namic models in similar situations. It will be made available on request.
Specifically in regard to the SPH model used here, vertical and hor-
5. Conclusions izontal oscillation tests show that with sufficient resolution SPH simu-
lations can model the buoy's initial motion to within 2% error, however
The experimental data measured in this study was used to compare to even at quite high resolutions simulations possess additional damping of
the results of SPH simulation in order to make some clear statements the motion. As expected, as particle resolution is increased, the predicted
about the resolution requirements needed in SPH to accurately model the motion of the buoy becomes more accurate and the damping was

316
D.F. Gunn et al. Ocean Engineering 156 (2018) 306–317

lessened. These tests suggest that to maintain accuracy of 10% or less in References
heave and surge predictions, approximately 40–50 particles are required
to represent the typical length scale of a floating object (in the case here, Basic, J., Degiuli, N., Werner, A., 2014. Simulation of water entry and exit of a circular
cylinder using the ISPH method. Trans. FAMENA 38 (1), 45–62.
the diameter of a spherical buoy). Bouscasse, B., Colagrossi, A., Marrone, S., Antuono, M., 2013. Nonlinear water wave
When observing the buoy's response to a wave train, a large ampli- interaction with floating bodies in SPH. J. Fluid Struct. 42, 112–129.
tude transient modulation in the surge trajectory was observed for the 1 s Campbell, J.C., Vignjevic, R., Patel, M., Milisavljevic, S., 2009. Simulation of Water
Loading on Deformable Structures Using SPH. Tech Science Press, pp. 1–21, 49(1).
period case, but not in the 2 or 3 s period cases. This transient motion was Clauss, G.F., 2002. Dramas of the sea: episodic waves and their impact on offshore
under-predicted by the SPH simulations with a coarse particle spacing structures. Appl. Ocean Res. 24 (3), 147–161.
corresponding to approximately 20 particles across a diameter (10 mm). Cleary, P.W., 1998. Modelling confined multi-material heat and mass flows using SPH.
Appl. Math. Model. 22 (12), 981–993.
Increasing the resolution to that which corresponds to around 40 parti- Cummins, S.J., Rudman, M., 1999. Using an approximate projection method to model
cles across a diameter (5 mm), led to much better prediction of the initial incompressible flows in SPH. J. Comp. Physiol. 152, 584–607.
transients. The smaller amplitude of the modulation in the SPH simula- Cummins, S.J., Silvester, T.B., Cleary, P.W., 2012. Three-dimensional wave impact on a
rigid structure using smoothed particle hydrodynamics. Int. J. Numer. Meth. Fluid.
tions indicates that they are less sensitive to initial transient motions, and
68 (12), 1471–1496.
this is also apparent in the 2 and 3 s period cases where the first surge Doring, M., Oger, G., Alessandrini, B., Ferrant, P., 2004. SPH simulations of floating
peak is under-predicted. For the 2 and 3 s period cases and only 20 bodies in waves. In: Proceedings of OMAE04, 23rd International Conference on
particles across the buoy's diameter, the amplitude of the surge motion Offshore Mechanics and Arctic Engineering. Vancouver, British Columbia, Canada,
June 20–25.
was under-predicted by up to 25% and there was a phase error of up to Q. Gui, S. Shao, P. Dong, Wave impact simulations by an improved ISPH model, J.
12%. In both cases the heave motions were more accurate with errors less Waterw. Port, Coast. Ocean Eng. 140(3).
than 15%. Improved resolution is expected to significantly reduce these Gunn, D.F., 2017. Using 3D SPH to Model the Interactions of Rogue Waves and Floating
Tethered Bodies. Ph.D. thesis. Monash University. URL. https://doi.org/10.4225/03/
discrepancies as was observed in the 1 s wave train simulations. 59e40c54c0791.
The simulations conducted show that SPH predictions of the buoy's Gunn, D.F., Rudman, M., Cohen, R.C.Z., 2014. Modelling water wave and tethered
heave motion after the initial transient phase are usually reasonably structure interactions using 3D smoothed particle hydrodynamics. In: Proceedings of
the 19th Australasian Fluid Mechanics Conference, pp. 8–11.
accurate. As the heave motion is dominated by the weight and buoyancy Kajtar, B., Monaghan, J.J., 2009. SPH particle boundary forces for arbitrary boundaries.
forces on the buoy, the simulations indicate that SPH is capable of Comput. Phys. Commun. 180, 1811–1820.
accurately modelling the buoyancy forces. In contrast, the surge motion Khayyer, A., Gotoh, H., Shao, S.D., 2008. Corrected incompressible SPH method for
accurate water-surface tracking in breaking waves. Coast Eng. 55 (3), 236–250.
of the buoy is not as well predicted by SPH at the resolution equivalent to Khayyer, A., Gotoh, H., Shao, S., 2009. Enhanced predictions of wave impact pressure by
20 particles across the buoy (10 mm). At a finer resolution of 40 particles improved incompressible SPH methods. Appl. Ocean Res. 31 (2), 111–131.
across the buoy (5 mm), the error in the surge motion of the buoy is Le Touze, D., Marsh, A., Oger, G., Guilcher, P.M., Khaddaj-Mallat, C., Alessandrini, B.,
Ferrant, P., 2010. SPH simulation of green water and ship flooding scenarios.
reduced to approximately 10% in the wave train case.
J. Hydrodynam., Ser. B 22 (5, Suppl. 1), 231–236.
The buoy used in this paper is a simple geometry intended for Liu, X., Lin, P., Shao, S., 2015. ISPH wave simulation by using an internal wavemaker.
providing a fundamental understanding of the wave-structure in- Coast Eng. 95, 160–170.
teractions. A comparison between experiments and simulations has Monaghan, J.J., 1992. Smoothed particle hydrodynamics. Annu. Rev. Astron. Astrophys.
30, 543–574.
shown that the SPH method is able to provide reasonably accurate pre- Monaghan, J.J., 1994. Simulating free surface flows with SPH. J. Comput. Phys. 110, 399.
dictions of the buoy's motion in response to incident waves, provided that Monaghan, J.J., 2005. Smoothed particle hydrodynamics. Rep. Prog. Phys. 68,
sufficient resolution is used. The recommendation here is to use a reso- 1703–1759.
Omidvar, P., Stansby, P.K., Rogers, B.D., 2013. SPH for 3D floating bodies using variable
lution of no less than 40 particle spacings across a representative mass particle distributions. Int. J. Numer. Meth. Fluid. 72, 427–452.
dimension of a floating object. As seen in the results presented within, the Pan, K., Ijzermans, R.H.A., Jones, B.D., Thyagarajan, A., van Beest, B.W.H., Williams, J.R.,
resolution required to obtain good results can often lead to large number 2016. Application of the SPH method to solitary wave impact on an offshore
platform. Comp. Part. Mech. 3, 155–166.
of particles that were not practical to undertake with the code available Ren, B., He, M., Dong, P., Wen, H., 2015. Non-linear simulations of wave-induced motions
to us. This strongly suggests the need to utilise massively parallel of a freely floating body using WCSPH method. Appl. Ocean Res. 50, 1–12.
implementations of SPH to make better use of existing technology. Rudman, M., Cleary, P.W., 2013. Rogue wave impact on a tension leg platform: the effect
of wave incidence angle and mooring line tension. Ocean Eng. 61, 123–138.
Because SPH has distinct advantages in dealing with interfaces and fluid- Rudman, M., Cleary, P.W., 2016. The influence of mooring system in rogue wave impact
structure interaction, another direction that should be further developed on an offshore platform. Ocean Eng. 115, 168–181.
is the implementation of dynamically variable-resolution SPH. Such a Schellin, T.E., El Moctar, O., 2007. Numerical prediction of impact-related wave loads on
ships. ASME J. Offshore Mech. Arct. Eng 129, 39–47.
methodology would use higher particle resolution near the interface and
Shao, S., 2009. Incompressible SPH simulation of water entry of a free-falling object. Int.
floating object, but coarser resolution further away, allowing larger do- J. Numer. Meth. Fluid. 59 (1), 91–115.
mains and good results to be obtained with more moderate numbers of Souto-Iglesias, A., Delorme, L., Peres-Rojas, L., Abril-Perez, S., 2006. Liquid moment
particles. These two enhancements would appear to be essential in the amplitude assessment in sloshing type problems with smooth particle
hydrodynamics. Ocean Eng. 33 (11–12), 1462–1484.
further development of SPH as a strong contender for simulating real- Van Paepegem, W., Blommaert, C., De Baere, I., Degrieck, J., De Backer, G., De Rouck, J.,
world applications. Degroote, J., Vierendeels, J., Matthys, S., Taerwe, L., 2011. Slamming wave impact of
The experimental data presented in this paper is available from the a composite buoy for wave energy applications: design and large-scale testing. Polym.
Compos. 32 (5), 700–713.
corresponding author for use in validating other computational codes Yang, C., Lu, H., L€ ohner, R., Liang, X., Yang, J., 2007. Numerical simulation of highly
and methods. nonlinear wave-body interactions with experimental validation. In: Proceedings of
the International Conference on Violent Flows, Fukuoka, Japan.
Young, I.R., 1999. Wind Generated Ocean Waves, vol. 2. Elsevier.
Acknowledgements Zhao, X., Ye, Z., Fu, Y., Cao, F., 2014. A CIP-based numerical simulation of freak wave
impact on a floating body. Ocean Eng. 87, 50–63.
The first author gratefully acknowledges the scholarships from Zheng, X., Lv, X., Ma, Q., Duan, W., Khayyer, A., Shao, S., 2018. An improved solid
boundary treatment for wave-float interactions using ISPH method. Int. J. Naval
Monash University and CSIRO that supported this research. Arch. Ocean Eng. https://doi.org/10.1016/j.ijnaoe.2017.08.001 (In Press).

317

Anda mungkin juga menyukai