0% found this document useful (0 votes)
59 views27 pages

Wsi Study 2022

Uploaded by

Aditya Chaudhary
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
59 views27 pages

Wsi Study 2022

Uploaded by

Aditya Chaudhary
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 27

Open FOAM R

Journal
Volume 2 [Review Papers], Pages 116–142
ISSN: 2753-8168

A REVIEW ON THE MODELLING OF WAVE-STRUCTURE INTERACTIONS


BASED ON OPENFOAM

LUOFENG HUANG1,∗ , YUZHU LI2 , DANIELA BENITES-MUNOZ3 , CHRISTIAN WINDT4 , ANNA FEICHTNER5 ,
SASAN TAVAKOLI6 , JOSH DAVIDSON7 , RUBEN PAREDES8 , TADEA QUINTUNA9 , EDWARD RANSLEY10 ,
MARCO COLOMBO11 , MINGHAO LI12 , PHILIP CARDIFF13 , GAVIN TABOR14 ,

1 Cranfield University, UK
2 National University of Singapore, Singapore
3 University College London, UK
4 Technische Universität Braunschweig, Germany
5 University of Exeter, UK
6 Aalto University, Finland
7 Budapest University of Technology and Economics, Hungary
8 Escuela Superior Politécnica del Litoral, Ecuador
9 University of Liège, Belgium
10 University of Plymouth, UK
11 The University of Sheffield, UK
12 Chalmers University of Technology, Sweden
13 University College Dublin, Ireland
14 University of Exeter, UK

Email address: [email protected], [email protected], [email protected],


[email protected], [email protected], [email protected], [email protected],
[email protected], [email protected], [email protected],
[email protected], [email protected], [email protected], [email protected]

DOI: 10.51560/ofj.v2.65
Version(s): –
Repo: –

Abstract. The modelling of wave-structure interaction (WSI) has significant applications in under-
standing natural processes as well as securing the safety and efficiency of marine engineering. Based
on the technique of Computational Fluid Dynamics (CFD) and the open-source simulation framework
- OpenFOAM, this paper provides a state-of-the-art review of WSI modelling methods. The review
categorises WSI scenarios and suggests their suitable computational approaches, concerning a rigid, de-
formable or porous structure in regular, irregular, non-breaking or breaking waves. Extensions of WSI
modelling for wave-structure-seabed interactions and various wave energy converters are also introduced.
As a result, the present review aims to help understand the CFD modelling of WSI and guide the use
of OpenFOAM for target WSI problems.

1. Introduction
1.1. Background. Wave-Structure Interaction (WSI) is defined as ocean waves applying forces to one or
multiple solid bodies and the solid’s dynamic response changes the surrounding wave field simultaneously.
WSI processes are ubiquitous in both marine environment and engineering, e.g. influencing the distribu-
tion of sea ice and vegetation, and dictating the safety and performance of ships and offshore installations.
Thereby, the modelling of WSI is of great importance for marine science, design and operations.
WSI modelling has been categorised into one-way coupling and two-way coupling. In one-way cou-
pling, the fluid impacts the structure, but not the other way around. In two-way coupling models, the
fluid and the structure impact each other. The modelling started with analytical approaches. Initially,
mathematical equations were developed to describe ocean waves as a periodic movement along a timeline.
Those wave equations were used to calculate the harmonic wave loads on a structure, or wave-induced
structural motions. Subsequently, the wave diffraction by a structure was also formulated. Such an-
alytical solutions for WSI have been developed since the early 20th century and approached a degree
of maturity around the 1990s, with a wide range of problems addressed by the Morison equation and
∗ Corresponding author

Received: 1 October 2021, Accepted: 17 August 2022, Published: 28 August 2022

© 2022 The Authors. This work is an open access article under the CC BY-SA 4.0 license

116
A REVIEW ON THE MODELLING OF WAVE-STRUCTURE INTERACTIONS BASED ON OPENFOAM 117

Figure 1. CFD simulation example of waves interacting with a realistic structure [10]

the potential flow theory coupled with structural solutions [1–5]. However, most analytical WSI works
were limited to linear wave conditions and inviscid fluids to obtain closed-form solutions, only valid for
small-amplitude waves.
Since the 2000s, nonlinear wave modelling has become popular using the potential flow theory in com-
bination with higher-order Boundary Element Methods (BEM). This BEM approach allows the prediction
of high-steepness waves and the corresponding response of structures, which has shown good accuracy
against experiments so long as the wave does not break by itself or by the structure [6, 7]. However, the
BEM approach is based upon dimensionality reduction as the discretisation is applied on a boundary
rather than on a volume. This limits the applicable structures to simple geometries, such as a cuboid, a
cylinder or a sphere. Overall, the BEM approach still cannot model WSI phenomena that involve strong
nonlinearities.
With the advances in modern computational techniques, computational fluid dynamics (CFD) models
have been developed to address more realistic WSI problems. In CFD, the physical domain is represented
by a corresponding computational domain, in which a solid geometry can be inserted. The space between
the surface of the structure and the domain boundaries is the fluid field, which can be discretised and
use the Navier-Stokes equations to obtain highly-nonlinear wave profiles.
There are two main branches of CFD, mesh-free methods and mesh-based methods. The compu-
tational speed has been a significant challenge for mesh-free methods, such as the Smoothed Particle
Hydrodynamics (SPH), which represents the fluid and structure as particles, e.g. [8]. In addition, to
develop boundary conditions such as inlet/outlet and the method requires artificial inputs for viscous
effects [9]. These are significant challenges for WSI modelling, in which inlet/outlet and fluid viscosity
are essential. Thus the WSI applicability of mesh-free methods is limited.
In terms of mesh-based methods of CFD, for example the Finite Volume Method (FVM), the solid
geometry is modelled as a closed surface that is in contact with numerous computational cells to represent
its structural complexity. In addition, a specific boundary-layer mesh can be built around the geometry
to account for fluids’ boundary-layer effect. Local mesh refinements are mature with FVM, allowing
high-resolution mesh to be applied in significant regions, e.g. around the free surface and the structure.
These features render mesh-based CFD a suitable approach to model and analysing WSI problems, with
the level of fidelity sufficiently high that the simulation can closely reproduce what happens in real sea
states (see Figure 1 for an example).
118 HUANG, LI, BENITES-MUNOZ, WINDT, FEICHTNER, TAVAKOLI, DAVIDSON, PAREDES, QUINTUNA, ET AL.

Figure 2. CFD software used for WSI simulations of WECs and their relative popu-
larity among approximately 200 reviewed studies [21]

1.2. Choices of CFD software. Different software is available for WSI simulations, ranging from com-
mercial and open-source software to in-house developments. Commercial tools commonly feature Graph-
ical User Interfaces (GUIs) and specific numerical methods, e.g. for multi-physics analysis. However, the
often significant license fees and, to some extent, the restricted access to the source code, possibly hinders
the usage of commercial software for WSI applications.
By avoiding license fees and providing access to source code, open-source CFD software has gained
popularity and is commonly backed by active communities. For OpenFOAM, dedicated workshops and
user group meetings facilitate knowledge exchange within the community, driving new developments. In
addition, a few private companies, such as Engys or WIKKI, are heavily involved in code development.
Drawbacks of using open-source software may arise from steep learning curves for beginners. Furthermore,
developments of new (advanced) numerical tools lack profit as a motivator. However, unlimited access
to the source code enables custom code development, facilitating the open-source software to be applied
in a wide range of applications. Apart from OpenFOAM, there is also other open-source software that
has made significant contributions to WSI modelling, such as REEF3D and Gerris [11, 12].
In-house codes are either driven by specific (physical) problems or the desire to include advanced
numerical algorithms. Their nature makes them unavailable to a large community. Examples of such
codes can be found in [13, 14].
The wide range of different CFD software suites raises the question of which one to select. Decision
drivers of choosing a specific CFD software for WSI are diverse, often being the available structural
solutions or numerical wave generation and absorption. Additional decision drivers may include project
time frame, budget, and user experience.
In principle, there is not a function that is implemented in commercial software that cannot be im-
plemented in OpenFOAM, and vice versa. The main differences, on one hand, include that OpenFOAM
enables more possibilities for learning and research purposes, as it is completely free to view and develop
the code. On the other hand, the learning curve for OpenFOAM is known to be longer than commercial
software, making it more challenging for a beginner who may have limited time for a specific project.
Some studies can be found which perform comparative analyses of different CFD software. Based
on the analysis of extreme wave loading, Westphalen et al. [15] consider STAR-CCM+, ANSYS CFX,
the in-house Cartesian cut-cell solver AMAZON-SC 3D, and an SPH solver. Sjökvist et al. [16, 17]
present a comparison of OpenFOAM with ANSYS Fluent. More recently, the CCP-WSI Blind Test
Series 1-3 [18–20] aimed at delivering comprehensive code-to-code and code-to-experiment comparisons
of different modelling approaches and dedicated experimental validation data. By way of example, Figure
2 shows the CFD software used for WSI simulations of Wave Energy Converters (WECs) and their relative
popularity in the literature.

1.3. Scope of this paper. Although CFD simulations have been shown capable of modelling WSI,
the quality of the simulations is dependent on the user’s computational setups. The capabilities of
CFD provide abundant options on numerical models and setups to account for the free surface, wave
A REVIEW ON THE MODELLING OF WAVE-STRUCTURE INTERACTIONS BASED ON OPENFOAM 119

modelling, turbulence effect, and structural response, among others. When inappropriate setups are
used, simulations can crash or lead to unphysical/erroneous results. For this reason, review articles
can bring significant contributions by helping users understand CFD and providing recommendations on
suitable modelling branches for specific problems.
In order to provide a comprehensive review of CFD modelling of WSI problems, the present work selects
OpenFOAM as the framework. Based on the open-source code, primary WSI modelling approaches are
reviewed and their applicability and limitation for various physical problems are discussed. This allows
readers to refer to corresponding source code, gaining a deeper understanding than appreciating black-box
functions in commercial CFD tools.
The remainder of this paper is organised as follows: Section 2 introduces relevant fluid modelling
approaches, i.e. modelling different types of ocean surface waves. Section 3 reviews WSI modelling
concerning various structural characteristics, for 3.1 Rigid fixed structures, 3.2 Rigid floating structures,
3.3 Deformable structures, 3.4 Porous structures; furthermore, Section 3.5 provides an extension of WSI
modelling that couples with seabed response. Section 4 presents the modelling of various types of WECs,
where the modelling methods of mooring lines and power-take-off systems are also introduced. Finally,
Section 5 summarises this work with its main conclusions.

2. Wave flow
2.1. Free surface modelling. A key requirement for simulating water waves in OpenFOAM is modelling
the evolving free surface interface. The most popular method for achieving this is the Volume of Fluid
(VOF) approach, detailed in Section 2.1.1, which can simulate complex WSI problems such as those
involving wave breaking, overtopping and slamming. Alternative methods are also available, as discussed
in Section 2.1.2. These methods are not widely applicable as VOF but can be capable of modelling a
restricted range of WSI conditions with potential computational savings.
2.1.1. The VOF method. The VOF method tracks the relative volumetric percentage of different fluids
in each cell. For the case of wave simulations, the VOF method typically considers two fluid phases:
water and air. This is depicted in Figure 3a, where the volume fraction of water is displayed - a value of
1 corresponds to a cell containing 100% water, a value of 0 corresponds cells with 100% air, and a value
between 0 and 1 represent the cells around the free surface interface.
The most commonly used OpenFOAM solver for wave modelling is interFoam, which considers two
fluid phases (e.g. air and water) and that the fluids are incompressible, immiscible, and viscous. However,
other solvers, shown in Figure 3, might be more efficient for specific cases e.g. simulating an oil spill [23],
or to include air compressibility e.g. calculating wave impact/slamming loads [24–26] and simulating
an oscillating-water-column WEC [27, 28]. In addition, it is possible to decrease the complexity of the
model by neglecting the effect of viscosity e.g. to reduce the computational burden of fully resolving the
boundary layers [29].
A variant of the interFoam solver, interIsoFoam, has been developed and is gaining popularity due to its
improved ability to handle the free surface interface. The interIsoFoam solver employs the novel geometric
VOF algorithm, isoAdvector [30], which maintains a sharp interface by constructing and advecting an
isosurface inside the cells around the interface position. By contrast, the original interFoam solver
utilises the multi-dimensional limiter for explicit solution (MULES) method, which employs an artificial
compression velocity term to bound the interface [31]. Considering wave propagation, interIsoFoam has
been shown to maintain a sharper interface than interFoam [32] and significantly reduce the excessive
wave dissipation inherent in traditional VOF methods [33], despite the considerably higher computational
cost.
To tackle the spurious velocities and, thus, allow lager time steps and better performance of turbu-
lence models at the interface, recently, the Ghost Fluid Method (GFM) has been implemented within
OpenFOAM [34]. The GFM ensures a continuous velocity field across the interface by coupling the two
phases (high and low density) via specific jump conditions at the interface.
In a comparative study, Peltonen et al. [35] investigated the performance of the GFM (based on [34])
and the traditional VOF for a number of marine-related test cases, i.e. two-dimensional inviscid flow
over a step, two-dimensional interface shear layer, and ship wave generation for a Wigley hull and the
Hamburg test case. The authors found rather subtle differences for the specific test cases and suggest a
further improvement by coupling GFM with geometric interface capturing methods, such as isoAdvector.
Recently, MULES has been improved with the Piecewise-Linear Interface Calculation (PLIC) and
Multicut PLIC (MPLIC) [36]. PLIC represents an interface by surface cuts that split each cell to match
the volume fraction of the phase in that cell. The surface cuts are oriented according to the point field of
the local phase fraction. The phase fraction on each cell face, i.e. the interpolated value, is then calculated
120 HUANG, LI, BENITES-MUNOZ, WINDT, FEICHTNER, TAVAKOLI, DAVIDSON, PAREDES, QUINTUNA, ET AL.

(a) interFoam and interIsoFoam (The VOF method)

(b) potentialFreeSurfaceFoam

(c) shallowWaterFoam

Figure 3. Depiction of the methodologies of different OpenFOAM solvers for free-


surface flow [22]

from the amount submerged below the surface cut. MPLIC is for handling multiple surface cuts, where
a single cut in one cell is insufficient, e.g. the water volume in one cell is between two separated air
volumes. PLIC and MPLIC thus have improved the modelling of interface/free surface, e.g. achieved
high-resolution modelling of water splash [37].

2.1.2. Alternative methods. Schmitt et al. [22] explored the choice of solvers in OpenFOAM other than
those based on the VOF method, which could be utilised to implement numerical wave tanks, namely:
potentialFreeSurfaceFoam and shallowWaterFoam. While hundreds of publications have utilised the VOF
method for WSI, Schmitt et al. [22] found that less than a dozen have used shallowWaterFoam and only
A REVIEW ON THE MODELLING OF WAVE-STRUCTURE INTERACTIONS BASED ON OPENFOAM 121

Figure 4. Plunging-type breaking wave simulation [43], where k/ων indicates the tur-
bulence level.

two have used potentialFreeSurfaceFoam. The potentialFreeSurfaceFoam solver is a single-phase solver,


as depicted in Figure 3b, that calculates the free surface as a single-valued function, thus cannot simulate
effects such as wave breaking. PotentialFreeSurfaceFoam is based on the pimpleFoam solver, which is
widely used in a broad range of applications for incompressible, single-phase, transient flows. A special
boundary condition is added to the pimpleFoam solver, called waveSurfacePressure, which calculates the
change in surface elevation at each timestep based on the volume flux for the cells at the top boundary.
The shallowWaterFoam solver employs simplified equations that are valid in shallow water conditions.
The shallow water equations [38] enable the velocity to be represented by a depth-averaged horizontal
component, eliminating the requirement to solve for the vertical direction. Therefore, massive reductions
in the overall cell count are achieved since the domain only needs to be discretised using one cell in the
vertical direction, as depicted in Figure 3c. The shallow water equations only consider the water phase,
thus shallowWaterFoam is a single-phase solver, where the surface elevation is directly available as a
simulation variable. Another alternative solver is interTrackFoam, which applies an interface tracking
that considers single-phase by enforcing the free surface boundary conditions on it [39]. However, this
method is not applicable in a situation where the water can be discontinuous due to the structure, e.g.
waves bring partial water to go on top of a structure [40], because in this scenario the free surface is
no longer a continuous boundary. More examples of using alternative models to VOF can be found in
Schmitt et al. [22].

2.2. Wave modelling.

2.2.1. Variety of waves. There are a variety of wave theories that can be used to generate progressive
ocean surface waves. The waves can be divided into regular (linear and nonlinear), irregular waves and
focussed waves.
For regular wave modelling the incident waves are periodic. To represent more realistic wave conditions,
various regular waves of different wave heights and frequencies can propagate in a certain region, known as
irregular waves (also sometimes called random waves). The spectrum of irregular waves can be obtained
through on-site measurements and inputted into simulations. Commonly used wave spectra include the
JONSWAP, the Pierson-Moskowitz and the Bretschneider types [41,42], in which JONSWAP is considered
the most widely used.
Focussed waves are normally considered when assessing extreme wave loads, where a single high-crest
wave is formed due to the accumulation of wave components (although there are also trough-focussed or
focussed at any particular phase). The focussed crest waves are normally formed through the delicate
superposition of multiple wave peaks.
When water waves approach the coast, the water depth decreases to shallow enough that the wave
may feel the presence of seabed and change its shape. This action induces increasing wave height and
decreasing wavelength, which will cause the wave to break at a certain point. This process is known as
surf zone breaking waves, shown in Figure 4. The bathymetry will influence the wave breaking type from
spilling, plunging, collapsing, or surging. The modelling of breaking waves is different from modelling
non-breaking surface waves (for non-breaking waves potential-flow theory or a laminar model can be
used).
In the deep-water regime of offshore/coastal regions, where the surface wave is not influenced by
seabed, breaking waves can also occur due to the instability of wave trains subject to initially small
perturbations. The resonant interaction between the carrier wave trains and the small perturbations can
lead to an exponential growth of the wave amplitudes for the side bands, and further wave breaking. A
122 HUANG, LI, BENITES-MUNOZ, WINDT, FEICHTNER, TAVAKOLI, DAVIDSON, PAREDES, QUINTUNA, ET AL.

study by Li and Fuhrman [44] simulated the so-called Class I (Benjamin-Feir instability) and Class II
(crescent waves) deep-water wave instabilities involving wave breaking with the Reynolds stress turbulence
model.
In open seas, focussed wave groups can also form breaking waves. Resulting from the superposition
of multiple waves or a storm, such rogue waves can be formed as a random event [45]. Bredmose and
Jacobsen [46] modelled breaking waves generated from focused wave groups using the interFoam solver and
investigated the breaking wave interaction with an offshore wind turbine foundation in the intermediate
water. As strong turbulence is produced in wave breaking, this will require turbulence modelling that
will be discussed in Section 2.2.3.

2.2.2. Wave generation and absorption. To model waves numerically, it is essential to consider the gener-
ation of the desired wave at the inlet of the Numerical Wave Tank (NWT) and the absorption at the outlet
to avoid reflected waves. Numerical wavemakers can be implemented by different methods, namely: (a)
including a physically similar wave generator/paddle by making the inlet boundary move - this method is
similar to wave tank experiments [47,48]; (b) implementing dictating functions to force the CFD solutions
to present a target wave condition, where the relaxation method and the static boundary method are
two of the maturest [49, 50]. For (a), additional boundary condition modifications and dynamic mesh
treatment are required, therefore (b) is the typical approach is NWT modelling.
In the case of the relaxation technique, the generation of the wave is initialised by using the analytical
solution of the wave theory applied in the inlet zone. Normally the solutions are also artificially changed
in the outlet zone to absorb the wave thus modelling a far-field condition, as shown in Figure 5a. This
method has been implemented in a toolbox named waves2Foam [51].
In the relaxation approach, a user-defined function is introduced to modify the solution of a computa-
tional variable ξ (e.g. velocity or volume fraction of water) of each timestep by blending a target value,
as ξ = χ ξcomputed +(1 - χ) ξtarget , where χ is the relaxation coefficient taking values between 0 and 1,
as shown in Figure 5b. With this method applied, theoretical target waves are forced at the inlet and
outlet boundaries (where χ = 0) and unmodified computed results are obtained in the middle of the
NWT, where χ = 1, with the inlet and outlet zones gradually changing ξ to relax the solution change.
The relaxation technique has been proven effective and stable, mostly able to generate and absorb waves
with negligible error [52–54]. Nonetheless, the size of the computational domain tends to increase due to
adding the relaxation zones.

(a) Inlet and outlet relaxation zones (grey) in a numerical wave tank

(b) The value of the spatial weighting factor

Figure 5. Illustration of the relaxation method [55]

As an alternative, the static boundary method that uses boundary conditions to control the generated
wave was developed, and it is available in OpenFOAM as IHFoam and olaFlow [10,56]. At the wave gen-
eration boundary, the boundary conditions are obtained from wave theory. For the absorption boundary,
a correction velocity is applied to cancel out the incoming waves. This method has also been successfully
A REVIEW ON THE MODELLING OF WAVE-STRUCTURE INTERACTIONS BASED ON OPENFOAM 123

used across the WSI community [57–59], and further information on the developer’s toolbox can be found
in [51, 56, 60].
Compared to the relaxation method, the static boundary method can avoid the use of relaxation zones
that increases the computational cost. Nonetheless, it has been highlighted that wave absorption using
the static boundary method is a challenge, which could cause undesirable reflection as also pointed out
by Jacobsen [61]. The wave field near the outlet boundary is normally disturbed by a structure, i.e. it
is not pre-known as an incident wave field, so measuring the flow before the outlet is essential for the
cancelling out. However, the measured flow near the outlet is a combination of incident and reflected
wave, which means wave filters must be used to derive a cancelling-out wave train from the outlet. The
wave filtering technique strongly relies on wave theories. Previously, this approach was only valid for
linear wave theory in shallow water [62]. Further development has been made by Higuera [63] to enable
the absorption of deep waters, but the author still highlighted the incapability of absorbing nonlinear
waves. Recently, Borsboom and Jacobsen [64] have made an advancement by analytically decomposing
the measured flow into multiple modes. This significantly improved the absorption of nonlinear waves
and the undesired reflection was reported to be less than 5% of the incident waves.
Windt et al. [50] used different test cases to assess available OpenFOAM methods for NWT. Their
comparison suggested that the static boundary method is the most computationally efficient for wave
generation, while the wave absorption part may be better replaced by using a numerical beach method,
which is implemented as a dissipating region nearby the outlet.
For either the relaxation method or the static boundary method, it is worthwhile to note that at least
10 cells per wave height and 100 cells per wavelength should be used to generate an accurate target wave.
This mesh density generally gives a deviation of less than 1% against theoretical waves, provided that
other setups are reasonable [65]. More information on the different wave generation methodologies can
be found in the reviews of Schmitt and Elsaesser [49].

2.2.3. Turbulence modelling. WSI can be accompanied by significant turbulent behaviours that dissipate
the kinetic energy and change the fluid behaviours and structural loads. Direct Numerical Simulation
(DNS) can accurately replicate turbulent flows, but this requires solving the WSI problems with an
extremely high mesh density that most computing facilities can unlikely afford. As a result, assumption-
based modelling is commonly required. Such assumptions have categorised turbulence modelling strate-
gies into several groups, known as Reynolds-Average Navier-Stokes equations (RANS), Large-Eddy Sim-
ulation (LES) and their combination (Hybrid). These methods apply certain numerical treatments to
account for the turbulent effects in the simulation, allowing savings in cell amount. More details of dif-
ferent turbulence modelling approaches can be seen in [66]. Figure 6 illustrates the approaches’ levels of
fidelity and the computational demand.
In general, most of the studies of WSI have employed a two-equation model, e.g. the k − ω or k − ϵ
model (where k is the turbulence kinetic energy, ω or ϵ indicates the dissipation rate to close the RANS
equations for the turbulent flow [68]. This can be corroborated by Windt et al. [69] on the modelling
of WEC (Figure 7). Besides two-equation models, there are also other RANS-based turbulence models,
e.g. Algebraic models, one-equation models, Reynolds stress models (RSM). There are two main reasons
for RANS-based two-equation models to be the mainstream choice: (a) the computational resources
required for RANS-based two-equation models are much lower than Hybrid or LES, which makes RANS
simulations affordable by contemporary computing facilities. (b) RANS-based two-equation models are
seen to provide sufficient accuracy for most 3D WSI applications, so there is normally no necessity to
apply a higher-order turbulence modelling approach. Validation examples of RANS-based turbulence
modelling in various WSI applications will be included in Section 3.
Two-equation turbulence models have also been widely used in the past two decades for simulating
surf zone breaking waves. However, seemingly all the simulations with two-equation models (both k − ω
and k − ϵ types) have shown a tendency to overestimate turbulence levels for breaking waves, both in
the pre- and post-breaking regions. For example, the results presented in Brown et al. [70] performed
with OpenFOAM using different turbulence models have shown over-production of turbulence even in
the pre-breaking regions (in this region, the waves can be modelled with potential-flow theory and the
turbulence is nearly zero). The same phenomenon has also been observed in the simulations by Devolder
et al. [71] who added buoyancy production terms to the k − ω branches to eliminate turbulence pollution
from the air to the water phase.
Recently, Larsen and Fuhrman [72] have proved that this problem is due to the unconditional instability
of two-equation turbulence closure models (both k − ω and k − ϵ types) in the potential flow core region
beneath surface waves. A method for formally stabilising two-equation models was derived in their work,
and it was shown that stabilised two-equation models lead to pronounced improvement in the predicted
124 HUANG, LI, BENITES-MUNOZ, WINDT, FEICHTNER, TAVAKOLI, DAVIDSON, PAREDES, QUINTUNA, ET AL.

Figure 6. Illustration of the fidelity and computational requirement for different tur-
bulence modelling methods [67]

Figure 7. Turbulence models used for CFD modelling of WECs [69]

turbulence and undertow velocity profiles prior to breaking and in the outer surf zone. Fuhrman and
Li [73] analysed a more complicated but popular turbulence model – the realisable k − ϵ model. They
found that this model is conditionally unstable in the potential flow region beneath surface waves. They
likewise stabilised the realisable k − ϵ model so that the initial conditions would not affect the stability
(without exponential growth of turbulence beneath waves in the potential flow region). However, even
the stabilised two-equation models in Larsen and Fuhrman [72] and Fuhrman and Li [73] were still rather
inaccurate in the inner surf zone (i.e. closer to the shoreline), thus seemingly requiring yet more advanced
methods of achieving turbulence closure.
Li et al. [43] analysed more advanced RANS-based Reynolds Stress turbulence Models (RSMs) which
solve all components of the Reynolds stress and break free from the Boussinesq approximation (which is
a basis for all RANS-based two-equation models). The RSMs were proved to be reasonable for simulating
A REVIEW ON THE MODELLING OF WAVE-STRUCTURE INTERACTIONS BASED ON OPENFOAM 125

non-breaking progressive wave trains without having the problem of over-production of turbulence in the
potential flow region beneath surface waves. The RSM model has also achieved good accuracy in the
prediction of coastal breaking waves on a sloping beach, especially the undertow velocity, as presented in
the work of Li et al. [43,74]. The model has also been applied to simulate the deep-water wave instabilities
involving wave breaking, as presented in Li and Fuhrman [44]. The RSM model in terms of the stress-ω
model is publicly shared through Li [75].
Nonetheless, some WSI studies performed with the laminar model (assuming no turbulence effect) have
also shown good accuracy, e.g. [76]. This is probably due to the fact that the waves are non-breaking
and the inertial effect in the problem is stronger than the viscous effect. Therefore, whether turbulence
modelling is required for solving a WSI problem depends on the physics of the fluid flow in the particular
case. For example, Li et al. [76] examined the Keulegan-Carpenter (KC) number for their study case on
wave interaction with a gravity-based foundation. They found that the turbulence effect is negligible as
the KC number is relatively low, in which case a turbulence modelling approach may not be needed.

2.2.4. Air entrainment. Air entrainment can be driven by turbulent motion near the free surface, e.g.
in high-velocity open channel flows. High-fidelity studies such as DNS have been conducted on air
entrainment in relatively small-scale problems such as in the case of a stationary turbulent hydraulic jump
[77]. For larger-scale problems such as spillways, DNS is less practical due to the vast computational cost.
Rather, RANS turbulence models have been widely applied. For example, Lopes et al. [78] implemented
the entrained air flux estimator of Ma et al. [79] in interFoam, coupled with the k − ω SST turbulence
model to simulate the stepped spillway. Air entrainment is also commonly considered in breaking waves.
Tomaselli [80] numerically investigated the air entrainment in breaking waves and their interaction with
a monopile also using interFoam, combined with an LES turbulence approach. The air phases in the
aforementioned studies were considered incompressible. Air compressibility has also been considered and
implemented in OpenFOAM as compressibleInterFoam. Relevant studies can be found such as Ferrer et
al. [26] and López et al. [28] who studied dam break, plunging wave impact at a vertical wall, and an
oscillating water column.

3. WSI modelling
3.1. Wave interaction with rigid fixed structures. Wave interactions with rigid fixed structures are
commonly seen in coastal defence and offshore wind turbines. Figure 8 provides an example of this type
of WSI. As the structure is fixed, the modelling is relatively simple, as the fixed structure can be treated
as a wall boundary condition that does not involve any structural solutions or dynamic mesh treatment.
Therefore, successful modelling for this type of WSI mainly relies on accurate representation of the fluid
behaviours, i.e. the wave or turbulence modelling, which has been introduced in Section 2.
Extensive studies have used OpenFOAM to assess coastal and offshore engineering problems with a
fixed structure. Since a sea wall or a breakwater normally expands extensively in the profile direction,
2D simulations have been widely used to perform analyses and agree well with wave flume experiments.
For example, Morgan and Zang [81] demonstrated OpenFOAM’s ability to produce accurate results for a
wide variety of 2D wave conditions and geometries. They considered submerged breakwater cases where
it was found that the numerical predictions agree with the experimental data, in terms of waveforms and
wave amplitude spectra. Their setup can be used to investigate the dependence of the wave transmission
and reflection properties of the breakwater. Chen et al. [82] performed 2D simulations to predict wave
overtopping on a dike structure. Their prediction of overtopping water level agrees with the experimental
measurements in the time domain. However, the accuracy is based on a stabilised k − ω model developed
by Larsen and Fuhrman [72], which is essential for accurately modelling wave propagation in the coastal
region and breaking on a structure.
In order to account for 3D structures in uni-directional waves but avoid the computational cost of
modelling a whole computational domain in 3D, a 2D-3D coupling model was developed by Di Paolo et
al. [83, 84], where 2D was used to solve wave generation from the inlet, stable propagation towards the
structure, and absorption near the outlet; 3D was used only around the structure. Both one-way and
two-way approaches are available. Validation against experimental data demonstrated that this approach
can provide accurate predictions when considering uni-directional wave conditions.
Full 3D simulations are required for the 3D multi-directional wave conditions as in real sea states.
However, the wave modelling may be simplified by the spectral approach, which provides boundary
conditions for a small CFD computational domain near the structure. An example of this treatment
is HOS-NWT [85], which has been validated against experiments [86]. Decorte et al. [87] used the
126 HUANG, LI, BENITES-MUNOZ, WINDT, FEICHTNER, TAVAKOLI, DAVIDSON, PAREDES, QUINTUNA, ET AL.

Figure 8. Uni-directional waves interacting with a fixed cylinder [41]

HOS-NWT approach to assess the wave load on a fixed wind turbine, according to the statistics in non-
Gaussian seas. In addition, the coupling of Machine Learning with CFD can also significantly speed up
the computational investigation, in particular when a large amount of simulations is required [88, 89].

3.2. Wave interaction with rigid floating structures. The interaction of waves with a rigid floating
body is a two-way process. Waves can induce body motions, and meanwhile those motions change the
surrounding wave field. An example is shown in Figure 9. To model this type of problem, one of the
multiphase solvers of OpenFOAM is needed to be used, e.g. interFoam. However, as the floating body is
moving, simulations need to be carried out with dynamic mesh. The particular module for dynamic mesh
can be sixDoFRigidBodyMotion. This module enables us to model the six Degrees of Freedom (DoF)
motion under the action of fluid and body forces. The motions in different directions can be restrained.
In summary, sixDoFRigidBodyMotion is used for wave-induced motions. In another scenario where the
body motion is prescribed rather than induced, another module called solidBody may be used. solidBody
allows to model either prescribed linear or angular harmonic motions. Such motions might help to model
the radiation problem from the prescribed motions of a floating object. Simulating prescribed harmonic
motions can also help measure the hydrodynamic coefficients of a floating object.
The main challenge in modelling the motion of an object is handling the dynamic mesh. Two different
methods can be used for 6DoF simulations: dynamicMotionSolverFvMesh and dynamicOversetFvMesh,
known as the morphing-mesh method and the overset method.
dynamicMotionSolverFvMesh morphs the mesh over every time step, where there are different libraries
for single-body, sixdofrigidbodymotion, and for multi-body, rigidbodymotion. The number of cells in the
meshing methodology is not changed over time, but the cells might get distorted or flexed under the
action of the motion of the solid body. An inner and an outer distance around the body are identified.
The changes in the cells occur in the area located in between these two distances. When the motion
of the solid body moving under the action of forces is relatively large, simulations might crash. This
commonly happens due to large distortion of the cells, which needs to be treated by additional steps,
such as reconfiguring the mesh after a certain simulation duration, sliding the mesh, or adding a dedicated
motion interpolating scheme [40, 47, 91].
Extensive work has been carried out by using the dynamicMotionSolverFvMesh. In Palm et al. [92],
CFD simulations were performed to model the dynamic response of WEC exposed to waves. In Islam et
al. [93], CFD simulations were run to numerically replicate the wave-induced motion of a floating barge.
More examples of the morphing-mesh technique in WSI can be found in [59, 94–96].
dynamicOversetFvMesh, known as the overset approach, divides the computational mesh into two
parts: background mesh and overset mesh (where multiple overset regions are possible). The background
A REVIEW ON THE MODELLING OF WAVE-STRUCTURE INTERACTIONS BASED ON OPENFOAM 127

Figure 9. Wave interaction with a rigid floating plate [90]

mesh denotes the fluid domain which is a fixed Eulerian framework; while the overset mesh attaches to
the structure, moving together with the structure based on its Lagrangian framework. In this way, both
parts of the mesh are not distorted. This method is particularly used for cases with large motions to
overcome the limitations of the dynamicMotionSolverFvMesh for those cases, such as a swinging WEC,
large structural motions induced by rogue waves, planning hulls, or a water-entry process [45, 97, 98]. At
the present stage, the overInterDyMFoam solver is used for overset simulations. The application of the
overset method for wave-structure interaction problems is well documented in [99]. The main challenge of
overset, however, is inaccuracies resulting from the interpolation that is needed to couple the background
and overset regions. Due to this, overset may violate mass conservation, and the pressure equation
needs to be adjusted implicitly to guarantee the transfer of fluxes correctly between overlapping regions.
Corresponding mitigations may be seen in [100]. Another potential solution is to build the overset region
sufficiently large to avoid communication locations near the area of interest [98], i.e. farther from the
structure, which might help in predicting the fluid load/impact.
Simulations using the overset technique for WSI are also widely conducted. Wu et al. [101] simulated
the dynamic motion of a small ice floe. They successfully modelled the wave-induced drift motion of ice
floe, having oscillatory motions in all six degrees of freedom. The model was found to have a great level
of accuracy in the prediction of the motion, following validation against experiments. Benites-Munoz et
al. [97] have used the overset technique to model a WEC rotating with a large angle, and good agreement
with experiments was reported. A power-take-off model is incorporated into the motion solver as a
spring-damping effect. Other overset applications in WSI can be found in [21, 102, 103].
The modelling of wave interaction with multiple floating bodies containing drift and collisions are still
in early-stage in OpenFOAM, while substantial progress might be seen in the near future. The coupling of
Discrete Element Method with wave modelling is suitable to model numerous scattered bodies interacting
wiht waves [104, 105].
3.3. Wave interaction with deformable structures. WSI can also involve a deforming structure.
For example, flexible wave energy converters that deform with waves are shown to have higher efficiency
and lower structural risk than rigid ones [106]; deforming breakwaters demonstrate evidently better
wave-attenuation performance than rigid ones [107]; natural features such as sea ice and vegetations also
can deform with waves [108, 109]. To include the deformation significantly increases the computational
128 HUANG, LI, BENITES-MUNOZ, WINDT, FEICHTNER, TAVAKOLI, DAVIDSON, PAREDES, QUINTUNA, ET AL.

challenge, because CFD itself does not contain structural deformation. Therefore, CFD is required to
be coupled with an additional set of computational solid mechanics (CSM) equations to calculate the
structural deformation. If the CSM solution is not fed back to CFD, this mechanism is known as one-way
coupling, and this is commonly applied for stress analyses, where the solid deformation is inconspicuous
while the structural internal stress can be critical, e.g. [110]. However, for WSI processes, deforming
structures may also significantly change the surrounding fluid field, if the structural deformation is large.
In order to accurately simulate the fluid field, the CSM solution is required to be fed back to CFD to
enable a two-way Fluid-Structure Interaction (FSI) process.
Waves’ interactions with flexible vegetation such as seagrass have also been simulated using Open-
FOAM. Chen et al. [111] presented a coupled wave-vegetation interaction model for simulating flexible
vegetation such as seagrass. The wave hydrodynamics is modelled with IHFoam, and the vegetation
motion is solved with the finite element method. The wave and vegetation models are coupled with an
immersed boundary method. The above work used a combination of solvers, e.g. interFoam for CFD and
an external solver for CSM, which requires a third code for coupling, data interpolation and simulation
management. This usually results in the FSI being one-way coupling, thus not fully representing the
physics.
The realisation of a two-way coupling FSI simulation can be achieved through a partitioned approach.
For every time step of the FSI simulation, (i) the solution of the fluid part is obtained first, and then the
fluid solution is passed as a force on the fluid-solid interface, so that the solid part can solve the solid
deformation. (ii) The deformed solid mesh will also induce the fluid mesh to deform. In weakly coupled
FSI, the solver goes to the next timestep immediately after step (ii), i.e. without further iterations
between fluid and solid domains. This may be acceptable when the solid deformation is small, as the
influence of mesh deformation on the fluid velocity field can be negligible. However, when the solid
deformation is large, weakly coupled FSI will induce errors i.e. inconsistent dynamic and kinematic
features between fluid and solid at the interface, and the errors can be accumulated over each timestep,
leading to unreliable results. To enable a strongly or fully coupled FSI, after step (ii), the fluid solution
(i.e. velocity and pressure of the fluid domain) is updated again in the current timestep, because the
fluid domain has changed with the mesh deformation. If the fluid part is solved again, likewise, the solid
solution will need to be updated again. This can require a large number of iterations until the fluid and
solid parts are fully coupled, i.e. with consistent dynamic and kinematic features at the interface), while
this demonstrates a higher level of accuracy than one-way coupling or weakly two-way coupling. Such a
process that solves fluid and solid separately is therefore called a partitioned approach. Fluid and solid
can also be implicitly solved together, but this FSI approach is still under development in OpenFOAM,
with some early progress shown, such as [112–114].
Based on the partitioned approach, OpenFOAM has made substantial progress in terms of FSI simula-
tions. Tukovic et al. [115, 116] developed an FSI code based on OpenFOAM (fsiFoam solver) with strong
two-way coupling as introduced above. An advantage of this OpenFOAM approach is that it employs
the finite-volume method (FVM) for both fluid and solid domains [115, 117]. Most current FSI works
involve a combination of solvers, usually with a finite-volume solver for the fluid flow and a finite-element
solver for the structural analysis, e.g. [118], which requires a third code for coupling, data interpolation
and simulation management. Thus, the combined alternative approaches for the fluid and solid domains
will tend to increase computational costs and imposes limitations on the coupling method. In contrast,
the entirely FVM approach of Tukovic et al. [116] makes an all-in-one solver under the framework of
OpenFOAM.
One limitation of fsiFoam was that it could only be applied to single-phase fluid modelling. Therefore,
it could not be used for multi-phase applications, e.g. WSI containing both air and water. In order
to simulate hydroelastic problems within OpenFOAM, Huang [119] incorporated fsiFoam with the VOF
approach to model multiphase flows, furtherly, with waves2foam to model WSI (named waveFsiFoam). In
this way, simulations were enabled for hydroelastic interactions of waves with a large elastic ice sheet [108]
and with an elastic breakwater [107], and the accuracy was validated against experiments. Figure 10
gives an example of highly-nonlinear breaking waves interacting with a seawall that undertakes large
deformations.
In recent years, the development of fsiFoam has been combined with a FSI toolbox, solids4foam,
led by Cardiff et al. [121]. solids4foam has made a significant advancement particularly in structural
solutions, by enabling the FSI simulations of viscoelastic, thermoelastic, and poroelastic solids [121, 122].
solids4foam is currently deemed to be a long-term FSI framework of the OpenFOAM family and a robust
solver for structural analyses. More applications of solids4foam can be found in [123–125].
A REVIEW ON THE MODELLING OF WAVE-STRUCTURE INTERACTIONS BASED ON OPENFOAM 129

Figure 10. An elastic seawall hit by highly nonlinear breaking waves [120]

3.4. Wave interaction with porous structures. The two main methods to model wave interaction
with porous structures are the micro- and the macro-scale approaches. With the micro-scale approach,
the detailed geometry of the structure is resolved. Accurate geometry and fluid flow resolution, however,
require a large number of mesh cells, and thus, a high computational demand. With the macro-scale
approach, the effect of the porous structure on the fluid is applied by means of its bulk effects. Here,
the porous structure is not explicitly resolved but represented by a geometrically defined porous region
in which a volume-averaged macro-scale model is added to the momentum equation through a source
term. This approach cannot reproduce the exact fluid flow inside or close to a porous structure but can
provide a sufficiently accurate replication of the mean flow for many WSI problems. The advantage of this
approach is a smaller computational demand than the micro-scale approach. A comprehensive review on
the mathematical foundations and solution techniques of macro-scale approaches in the context of coastal
structures is given in Losada et al. [126].
Porous structures can be categorized into large volumetric structures such as breakwaters or vegetation,
and thin porous structures such as perforated or slotted barriers. Depending on the overall aim of the
simulation, either the micro- or the macro-scale approach can be more suitable. For thin porous structures,
the micro-scale approach has been used in most of the literature, where common objectives are to derive
hydrodynamic coefficients, as for instance by Mentzoni and Kristiansen [127] for an oscillating perforated
sheet or by George and Cho [128] for a sloshing tank with a baffle; or to validate simpler (e.g. analytical)
models, as for example by Poguluri and Cho [129] for a vertically slatted barrier. Conversely, the macro-
scale approach has mainly been applied for thin porous structures when it is challenging to geometrically
resolve the structure or when the overall forces are of interest rather than details of the flow. Examples are
studies on fluid interaction with fish cages by Shim et al. [130] and Chen and Christensen [131] and with
circularly perforated structures by Feichtner et al. [132]. For large volumetric structures, it is common
to use a macro-scale approach when the bulk effects are to be investigated. This avoids challenges in
resolving the often complex and irregular geometries of the structures, and reduces the computational
demand significantly whilst providing sufficiently accurate fluid flow replication. Examples are simulations
of wave interaction with breakwaters by Jensen et al. [133] and Higuera et al. [134], or with vegetation,
for instance by Hadadpour et al. [135]. In contrast though, Maza et al. [136] represented a mangrove
forest microscopically by means of a cylinder array and studied tsunami wave interaction.
Within a macro-scale approach, the general formulation of the pressure drop across a porous structure,
commonly called the extended Darcy-Forchheimer law, ∆p/x = a × u + b/2 × u|u| + c × δu/δx, where a, b
and c are porosity coefficients that are added to the momentum equation to account for the effect of the
porous structure on the fluid. The linear term represents viscous effects, dominant for flow regimes with
small Re-numbers; the quadratic term represents turbulent effects with large Re-numbers; the transient
term represents effects due to fluid flow acceleration across the porous voids. The relative importance of
the viscous, inertial and transient terms and the related formulations of the porosity coefficients (a, b,
c) are highly problem-specific and key to accurately representing the properties of the structure and the
130 HUANG, LI, BENITES-MUNOZ, WINDT, FEICHTNER, TAVAKOLI, DAVIDSON, PAREDES, QUINTUNA, ET AL.

corresponding flow regime. For applications to volumetric granular material of coastal structures, options
of formulations are provided in the review by Losada et al. [126]. A comparison of common formulations
for the drag coefficient, b, for slatted barriers can be found in the work by Poguluri and Cho [129].
The porous pressure-drop equation can be implemented in three ways. One option is the implemen-
tation of volumetric porous media with either isotropic or anisotropic characteristics. This works for
both thin and large structures. An additional option for thin structures with negligible thicknesses, is
the porous baffle implementation where a pressure drop is applied across a surface. An overview of the
different types of implementation including illustrations can be found in Feichtner et al. [137], where
the three types have been compared for simulations of wave interaction with thin perforated sheets and
cylinders. The porous baffle or pressure-jump condition is a standard implementation in OpenFOAM and
can be used with interFoam or any other application solver. It introduces the linear and quadratic term
of the formulation in the pressure-drop equation but neglects the transient term, and the input requires
the thickness of the porous structure and the coefficients a and b (in OpenFOAM referred to as D and
F for “Darcy” and “Forchheimer”). A porosity solver for two-phase flow named porousInterFoam was
developed but this did not account for the limited amount of fluid inside the porous structure which can
lead to mass conservation problems. This mass conservation issue was addressed by Jensen et al. [133] and
Higuera et al. [134] through implementing macro-scale porosity based on the Volume-Averaged RANS
(VARANS) equations, in combination with the wave modelling toolboxes waves2Foam [51] (with the
porousWavesFoam solver by Jensen et al., [133]) and OlaFlow/IHFoam [134]. Both use porosity coeffi-
cients (a, b, c) based on granular material by default but it is straightforward to transfer the input to
also model thin porous structures, as in Feichtner et al. [137]. Figure 11 provides an example of the
macro-scale simulation approach. The mass conservation issue in porous materials was also addressed by
Romano et al. [138]; their development was combined together with overset and successfully applied to
landslide problems, which could potentially be adaptable to WSI problems.
A typical benchmark problem for fluid flow across a porous medium but without wave interaction
is the 2D porous dam-break problem. Also Jensen et al. [133] and Higuera et al. [140] have used it
for the validation of their porosity solvers where they followed the setups by Lin [141] and Liu et al.
[142]. Additional examples of CFD modelling for wave interaction with microscopically resolved porous
structures are works by Wu and Hsiao [143] on solitary wave interaction with a submerged slotted barrier,
and by Lee et al. [144] on wave interaction with a circular perforated caisson breakwater. Further examples
of wave interaction with porous structures represented by their macro-scale effects are studies by Srineash
et al. [145] in the context of dams, by Brito et al. [146] on submerged vegetated floodplains and work by
Zhao et al. [147] on a vertical net panel.
In the case of micro-scale porosity modelling, the challenges lie in the accurate representation of the
geometry of the structure, the correspondingly large number of mesh cells that are required and the
relatively high computational demand that follows as a consequence. The biggest challenge with the
macro-scale approach is that the pressure-drop model and the formulation of the porosity coefficients (a,
b, c) respectively are highly application dependent. Since no universal formulation exists, the models
rely on coefficients derived from experiments or high-fidelity micro-scale models. The solution of a
specific engineering problem in the context of wave interaction with porous structures generally requires
a combination of experiments, simpler models, and models with higher fidelity. Another topic, where
further studies are required, is turbulence modelling for macro-scale porosity representations. Opinions
differ on whether a model is required at all or whether turbulence effects are already accounted for by the
pressure-drop formulation. Losada et al. [126] provide a useful overview of the viewpoints and arguments.
They also identify future research needing to be done to answer the unresolved problems.
Maza et al. [136] used IHFoam to investigate tsunami wave interaction with mangrove forests. They
compared two different conceptual approaches to model the mangrove forests: one is to directly simulate
the detailed geometries of the rigid cylinder array; the other way is to consider the mangrove forests as a
porous rigid media and apply volume-averaging to model the momentum damping created by the plants.
The second approach can be more efficient; however, it was found that the maximum wave-exerted forces
were underestimated.

3.5. WSI with seabed response. For offshore and coastal structures that are built upon the seabed,
the investigation of the interaction between waves, the structure and the soil is critical for preventing
future structure failures. In conventional geotechnical wave-soil interaction modelling, analytical wave
pressure fields derived from linear or higher-order wave theories are usually applied as an external force for
solving soil responses. However, the analytical solutions are often not accurate or not applicable to predict
the wave-induced seabed responses in the presence of marine structures. When marine structures with
complicated geometries are installed above the seabed, it will alter the wave field and the forces on the
A REVIEW ON THE MODELLING OF WAVE-STRUCTURE INTERACTIONS BASED ON OPENFOAM 131

(a) The actual geometry

(b) Simulation

Figure 11. Wave interaction with a porous structure (macro-scale approach) [139]

surrounding seabed. In order to achieve better predictions for wave-structure-seabed interaction problems,
multi-physics numerical models have been developed, where the interaction between nonlinear waves and
soil (in the presence of the structure or not) is modelled. SedFoam is a well-known OpenFOAM solver
that includes seabed (or riverbed etc.) as an additional continuum phase, which has shown prominent
performance in the modelling of sedment-transport problems [148].
Liu et al. [149] made the first effort to investigate the seabed response in waves using an integrated CFD
approach. They applied a solver in the OpenFOAM framework for two immiscible incompressible fluids
(water and air) to produce a wave field with free surface. They also implemented a poro-elastic soil solver
by discretizing the Biot’s equations [150] for modelling the seabed. Later, Tang and Johannesson [151]
extended their work into an anisotropic model in the quasi-static form in the OpenFOAM framework.
The quasi-static anisotropic poro-elastic solver by Tang and Johannesson [151] was also validated and
applied in the work of Li et al. [76] in which the anisotropic consideration was proven to be practical
for modelling the seabed of medium and coarse sand. Li et al. [152] further developed the soil model
in the partial-dynamic form and incorporated a liquefaction module considering different liquefaction
criteria. The model is opensource at [153], and an example simulation is shown in Figure 12. Lin et
al. [154] used this approach to investigate the nonlinear wave-induced seabed response around mono-pile
foundation. Celli et al. [155] studied the effect of a submerged berm on the liquefaction near a rubble
mound breakwater in regular and irregular waves. Liang and Jeng [156] studied seafloor liquefaction
132 HUANG, LI, BENITES-MUNOZ, WINDT, FEICHTNER, TAVAKOLI, DAVIDSON, PAREDES, QUINTUNA, ET AL.

Figure 12. WSI simulation coupled with the response of seabed (the lower part) [152]

around a pipeline under combined JONSWAP wave spectrum and current conditions. A variation of
the poro-elastic model has been used by Ye et al. [157] to predict the subsidence of a rubble mound
breakwater that exists in real life. Plasticity modelling of the soil in OpenFOAM is also seen in the
literature. Tang et al. [122] implemented an FVM-based code of poro-elasto-plasticity soil model. The
model was built based on the Biot’s consolidation theory and combined with a perfect plasticity Mohr-
Coulomb constitutive relation. Elsafti and Oumeraci [158] implemented a multi-yield surface plasticity
model to simulate plastic soil response under cyclic loads in their CFD solver geotechFoam, also developed
in the OpenFOAM framework. A recent work from Shanmugasundaram et al. [159] has presented a
new development to calculate stress fields in seabed, which is capable of predicting residual-liquefaction
problems.
The soil models for wave-structure-soil interactions have been built based on the constitutive models
developed for onshore geotechnical engineering. Most studies have been performed based on Biot’s
[150] consolidation model and its extensions. In fact, the seabed is under seawater and subjected to
complicated environmental loads such as waves, current and seismic loadings. To date, a generally
appropriate seabed constitutive model for marine geotechnical engineering is not yet available in the
public literature. Whether the constitutive relations for onshore can be applied to offshore remains
to be investigated. Meanwhile, most of the existing works have also been limited to an uncoupled
approach or semi-coupled approach, rather than a fully-coupled approach. As the seabed deformed
under strong environmental loading or the structure moved its place, it will affect the wave fields and
loading distributions in return. Under circumstances such as complete liquefaction failures and large
deformations, the soil can no longer sustain the structure and the structure will move significantly.
Therefore, a fully-coupled model with strong two-way interaction of the wave-structure-soil system is
necessary to be developed.

4. Wave energy converters


In recent years, marine renewable energy systems, such as WECs, support global efforts to adopt clean
energies. WECs feature unique characteristics, such as enlarging body motions to extract maximum
energy. The large wave conditions and system dynamics of WECs involve significant non-linear effects,
which means linear or weakly non-linear potential flow methods are not suitable. Therefore, OpenFOAM
WSI developments are particularly valuable for the modelling of WECs.
4.1. WEC types. Different devices have been envisaged to exploit and harness the wave energy re-
source. Figure 13 shows some of the most prominent. Based on different modelling approaches, the
following section will divide WECs into four categories, i.e. Point Absorbers (PAs), Oscillating Wave
Surge Converters (OWSCs), Oscillating Water Columns (OWCs), and flexible WECs. For a specific
review of various aspects of WEC modelling in CFD, the interested reader is referred to [160].
4.1.1. Point absorbers. PAs are characterised by their small size relative to the incident wavelength.
The large device motion due to the action of energy maximising control systems [161] challenges the
A REVIEW ON THE MODELLING OF WAVE-STRUCTURE INTERACTIONS BASED ON OPENFOAM 133

Figure 13. Schematic of different WEC types: (a) PA; (b) Terminator; (c) Attenuator;
(d) OWSC; (e) OWC; (f) Pressure differential [160]

Figure 14. Wave interaction with an oscillating wave energy converter [97]

dynamic mesh motion methods [162]. Numerous studies can be found analysing PAs in the OpenFOAM
framework. Studies are conducted in order to, e.g., validate numerical models [19,20,163], evaluate viscous
drag effects [164, 165], analyse the device performance [166], investigate device scaling [167], assess WEC
control performance [59], or for conceptual device design [168].

4.1.2. Oscillating water columns. Within OWCs, the oscillating water inside a chamber forces entrapped
air to rotate a turbine. When modelling these devices in a CFD framework both highly non-linear
free surface deformations within the air chamber (i.e. sloshing) [169] and air compressibility [27] may
need to be considered. Furthermore, the implementation of the Power Take-Off (PTO) system is of
relevance, which can be achieved, e.g., by modelling a narrow orifice [53, 170, 171] or implementing a
porous material in a specific zone of the domain [172, 173] to model the pressure drop over the turbine of
the real system. CFD models can then be used for performance assessment [174] or geometry optimisation
studies [175, 176].

4.1.3. Oscillating wave surge converters. OWSCs rotate around a hinge point, extracting power from the
surging wave motion. While PAs are often characterised by multiple operational DoFs, OWSCs feature
large excursions in a single DoF, pitch. For the modelling of these devices in the OpenFOAM framework,
special treatment of the body motion is required, e.g. by using arbitrary mesh interfaces [47] An example
of the use of mesh morphing during the analysis of an OWSC can be found in [177]. Figure 14 shows
Benites-Munoz et al.’s work modelling the large rotation of an OWSC using the overInterDymFoam [97].
134 HUANG, LI, BENITES-MUNOZ, WINDT, FEICHTNER, TAVAKOLI, DAVIDSON, PAREDES, QUINTUNA, ET AL.

4.1.4. Flexible WECs. The above rigid types of WECs have suffered from structural problems induced
by large wave loads and did not prove to have sufficient energy output [178]. One potential solution can
be applying flexible materials for WECs, since the flexibility can effectively mitigate wave loads and the
wave-induced deformation can be utilised to generate significant power [106]. Some examples of flexible
WEC devices are shown in Figure 15. The functionality of flexible WECs involves complex structural
deformations that are based on fully-coupled FSI, i.e., any deformation of the structure triggers a response
of the wave flow and vice versa; in the WEC field, a modelling method with such capability has not been
seen in literature, which is pointed out as a crucial gap by the recent review of Collins et al. [106].
Fully-coupled FSI can be achieved through CFD+CSM, as introduced in Section 3.3. Solids4foam may
be a good platform to develop models for flexible WECs; this is demonstrated through a recent work of
Huang et al. using fully-coupled CFD+CSM to simulate a “strip shape” WEC [179]. Figure 16 shows
the solver’s capability of coupling the wave flow and the WEC deformation.

(a) Anaconda, UK (b) SBM S3, France

(c) AWS-III, UK (d) Wave Carpet, USA

Figure 15. Flexible WECs: existing examples

4.2. Sub-system modelling. In addition to the (floating) structure of a WEC, other sub-system com-
ponents should ideally be included in the hydrodynamic modelling framework. The most prominent
sub-systems are the PTO system (often in conjunction with the energy maximising control system) and
the mooring system. When including these sub-systems within the CFD-based NWT, high-fidelity models
are desirable to avoid spoiling the underlying fidelity of the (costly) hydrodynamic model.
4.2.1. Power take-off and control modelling. Including a PTO system within the NWT is inherently
required in order to assess the power output of a device. Hence, PTO systems have historically been the
first sub-systems to be included within NWTs. As stated in Section 4.1.3, when modelling OWCs, the
PTO can simply be mimicked by incorporating a narrow orifice or porous media. For PAs or OWSCs, a
simple implementation of a PTO can be realised by means of a linear spring-damper system, examples
of which can be found in, e.g. [103, 180, 181]. Non-linear PTO representation (with or without the
inclusion of energy maximising control systems) is more challenging in terms of their implementation.
Coulomb type dampers are implemented in [17, 71]. Penalba et al. [182] present the coupling between an
OpenFOAM-based NWT with a fully-nonlinear PTO model, running in MATLAB.
4.2.2. Mooring modelling. The inclusion of mooring systems within NWTs is attracting increased atten-
tion. As for the PTO models, different mooring models, with different levels of fidelity and different
levels of associated computational cost, are available. A relatively simple implementation of a mooring
system can be realised through linear springs, i.e. by simply adding a force that is proportional to the
structure displacement [19,20]. For more sophisticated, non-linear mooring models, coupling of the NWT
to external toolboxes is commonly used, examples of which can be found in [92, 183]. Available mooring
toolboxes for OpenFOAM include MooDy [164] and MoorDyn [184]. Chen and Hall [185] compared these
two toolboxes and found their accuracies are very similar. However, MooDy applies a finite-element
A REVIEW ON THE MODELLING OF WAVE-STRUCTURE INTERACTIONS BASED ON OPENFOAM 135

(a) Floating case

(b) Submerged case

Figure 16. Wave interaction with a deformable wave energy converter [179]

approach that is slower than MoorDyn, which applies a lumped mass approach. Apart from considering
the structural response in waves, MooDy also provides the capability to assess the snap loads in mooring
cables, important for the robustness of mooring systems themselves [186]. Barajas et al. [187] presented
a study using MooDy and overset to simulate the motions of a WEC design (without PTO).

5. Conclusions
The present paper has reviewed the progress of various WSI modelling approaches in OpenFOAM.
The numerical modelling of ocean surface waves and their interactions with fixed, floating, deformable or
porous structures have been described in detail. Various available modelling approaches are suggested,
and their pros/cons are discussed. The main conclusions of the review are recounted below in Table 1.
This paper is the first review of WSI modelling for OpenFOAM that has covered such comprehensive
topics. To the best knowledge of the authors, this is also currently the widest WSI review in the entire
CFD community. This contribution is expected to help readers better understand the relevant CFD
methodologies and provide directions on the selection of models as well as the future development of the
field.
Author Contributions: writing—original draft preparation, L.H., Y.L., D.B.M., C.W., A.F., S.T., J.D., R.P.,
T.Q.; writing—review and editing, E.R., M.C., M.L., P.C., G.T.; project administration, L.H.; All authors have
read and agreed to the published version of the manuscript.

References
[1] A. E. Heins, “Water waves over a channel of finite depth with a submerged plane barrier,” Canadian Journal of
Mathematics, vol. 2, pp. 210–222, 1950.
[2] J. Morison, J. Johnson, and S. Schaaf, “The force exerted by surface waves on piles,” Journal of Petroleum Technology,
vol. 2, no. 05, pp. 149–154, 1950.
[3] J. B. Keller and F. C. Karal Jr, “Surface wave excitation and propagation,” Journal of Applied Physics, vol. 31, no. 6,
pp. 1039–1046, 1960.
[4] R. Rainey, “A new equation for calculating wave loads on offshore structures,” Journal of Fluid Mechanics, vol. 204,
pp. 295–324, 1989.
[5] O. M. Faltinsen, J. Newman, and T. Vinje, “Nonlinear wave loads on a slender vertical cylinder,” Journal of Fluid
Mechanics, vol. 289, pp. 179–198, 1995.
[6] B. Z. Zhou and G. X. Wu, “Resonance of a tension leg platform exited by third-harmonic force in nonlinear regular
waves,” Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences, vol.
373, no. 2033, p. 20140105, 2015.
[7] B. Z. Zhou, G. X. Wu, and B. Teng, “Fully nonlinear wave interaction with freely floating non-wall-sided structures,”
Engineering Analysis with Boundary Elements, vol. 50, p. 117–132, 2015.
136 HUANG, LI, BENITES-MUNOZ, WINDT, FEICHTNER, TAVAKOLI, DAVIDSON, PAREDES, QUINTUNA, ET AL.

Table 1. Summary on the progress of contemporary WSI developments in OpenFOAM

topic key points


Wave generation and absorp- The relaxation method shows overall stable performance, but it
tion causes additional computational costs. The static boundary method
is relatively computationally cheap, but its wave absorption function
is complicated and undesirable wave reflection may remain.
Turbulence modelling Laminar and RANS are suitable for most WSI problems, while cases
involving breaking waves are turbulence-demanding. Recent ad-
vances have been made in the turbulence modelling of breaking waves
with RANS turbulence models.
WSI - rigid bodies WSI with a single solid body is relatively mature while with multiple
solid bodies is underdeveloped. A recommendation is to couple with
low-order wave modelling approaches to save computational costs
and include real sea states. Most existing studies have considered
regular waves – more research works on irregular waves are desired
in the future.
WSI - deformable bodies Two-way fully coupling has been achieved, while the computation
can become slow due to FSI iterations. Speeding up computation is
a key future development in this field.
WSI - porous structures Fully resolving the porous structure is computationally prohibitive.
Using an empirical approach to mimic porous layers is acceptable
but highly depends on the accurate selection of empirical values.
WSI with seabed The one-way coupling has been relatively mature, but the two-way
coupling is lacking.
WECs PTO is currently incorporated in simulations using oversimplified
ways, although it is complicated in real WEC systems.

[8] Y. Sun, G. Xi, and Z. Sun, “A fully lagrangian method for fluid–structure interaction problems with deformable
floating structure,” Journal of Fluids and Structures, vol. 90, p. 379–395, 2019.
[9] M. S. Shadloo, G. Oger, and D. Le Touzé, “Smoothed particle hydrodynamics method for fluid flows, towards industrial
applications: Motivations, current state, and challenges,” Computers & Fluids, vol. 136, p. 11–34, 2016.
[10] P. Higuera, “olaflow: Cfd for waves [software].” 2017. [Online]. Available: https://doi.org/10.5281/zenodo.1297013
[11] H. Bihs, W. Wang, C. Pakozdi, and A. Kamath, “Reef3d:: Fnpf—a flexible fully nonlinear potential flow solver,”
Journal of Offshore Mechanics and Arctic Engineering, vol. 142, no. 4, 2020.
[12] T. R. Keen, T. J. Campbell, J. D. Dykes, and P. J. Martin, “Gerris flow solver: Implementation and application,”
NAVAL RESEARCH LAB STENNIS DETACHMENT STENNIS SPACE CENTER MS OCEANOGRAPHY DIV,
Tech. Rep., 2013.
[13] D. M. Greaves, “Viscous waves and wave-structure interaction in a tank using adapting quadtree grids,” Journal of
fluids and structures, vol. 23, no. 8, p. 1149–1167, 2007.
[14] J. Sanders, J. E. Dolbow, P. J. Mucha, and T. A. Laursen, “A new method for simulating rigid body motion in
incompressible two-phase flow,” International Journal for Numerical Methods in Fluids, vol. 67, no. 6, p. 713–732,
2011.
[15] J. Westphalen, D. M. Greaves, A. Raby, Z. Z. Hu, D. M. Causon, C. G. Mingham, P. Omidvar, P. K. Stansby, and
B. D. Rogers, “Investigation of wave-structure interaction using state of the art cfd techniques,” Open Journal of
Fluid Dynamics, vol. 4, no. 01, p. 18, 2014.
[16] L. Sjökvist and M. Göteman, “Peak forces on wave energy linear generators in tsunami and extreme waves,” Energies,
vol. 10, no. 9, p. 1323, 2017.
[17] L. Sjökvist, J. Wu, E. Ransley, J. Engström, M. Eriksson, and M. Göteman, “Numerical models for the motion and
forces of point-absorbing wave energy converters in extreme waves,” Ocean Engineering, vol. 145, p. 1–14, Nov 2017.
[18] E. Ransley, S. Yan, S. A. Brown, T. Mai, D. Graham, Q. Ma, P.-H. Musiedlak, A. P. Engsig-Karup, C. Eskilsson,
and Q. Li, “A blind comparative study of focused wave interactions with a fixed fpso-like structure (ccp-wsi blind test
series 1),” International Journal of Offshore and Polar Engineering, vol. 29, no. 02, p. 113–127, 2019.
[19] E. Ransley, S. Yan, S. Brown, M. Hann, D. Graham, C. Windt, P. Schmitt, J. Davidson, J. Ringwood, and P.-H.
Musiedlak, “A blind comparative study of focused wave interactions with floating structures (ccp-wsi blind test series
3),” International Journal of Offshore and Polar Engineering, vol. 30, no. 01, p. 1–10, 2020.
[20] E. J. Ransley, S. A. Brown, M. Hann, D. M. Greaves, C. Windt, J. Ringwood, J. Davidson, P. Schmitt, S. Yan, J. X.
Wang, J. H. Wang, Q. Ma, Z. Xie, G. Giorgi, J. Hughes, A. Williams, I. Masters, Z. Lin, H. Chen, L. Qian, Z. Ma,
Q. Chen, H. Ding, J. Zang, J. van Rij, Y.-H. Yu, Z. Li, B. Bouscasse, G. Ducrozet, and H. Bingham, “Focused wave
interactions with floating structures: a blind comparative study,” Proceedings of the Institution of Civil Engineers -
Engineering and Computational Mechanics, vol. 174, no. 1, p. 46–61, Mar 2021.
A REVIEW ON THE MODELLING OF WAVE-STRUCTURE INTERACTIONS BASED ON OPENFOAM 137

[21] C. Windt, J. Davidson, B. Akram, and J. V. Ringwood, “Performance assessment of the overset grid method for
numerical wave tank experiments in the openfoam environment,” in International Conference on Offshore Mechanics
and Arctic Engineering, vol. 51319. American Society of Mechanical Engineers, 2018, p. V010T09A006.
[22] P. Schmitt, C. Windt, J. Davidson, J. V. Ringwood, and T. Whittaker, “Beyond vof: alternative openfoam solvers
for numerical wave tanks,” Journal of ocean engineering and marine energy, vol. 6, no. 3, p. 277–292, 2020.
[23] H. Yang, J. Lu, and S. Yan, “Preliminary numerical study on oil spilling from a dht,” in The Twenty-fourth Interna-
tional Ocean and Polar Engineering Conference. OnePetro, 2014.
[24] F. Gao, Z. H. Ma, J. Zang, D. M. Causon, C. G. Mingham, and L. Qian, “Simulation of breaking wave impact
on a vertical wall with a compressible two-phase flow, model,” in The Twenty-fifth International Ocean and Polar
Engineering Conference. OnePetro, 2015.
[25] B. R. Seiffert, R. C. Ertekin, and I. N. Robertson, “Wave loads on a coastal bridge deck and the role of entrapped
air,” Applied Ocean Research, vol. 53, p. 91–106, 2015.
[26] P. M. Ferrer, D. M. Causon, L. Qian, C. G. Mingham, and Z. H. Ma, “Numerical simulation of wave slamming on a
flap type oscillating wave energy device,” Proceedings of the Twentysixth, 2016.
[27] I. Simonetti, L. Cappietti, H. Elsafti, and H. Oumeraci, “Evaluation of air compressibility effects on the performance
of fixed owc wave energy converters using cfd modelling,” Renewable Energy, vol. 119, p. 741–753, 2018.
[28] I. López, R. Carballo, F. Taveira-Pinto, and G. Iglesias, “Sensitivity of owc performance to air compressibility,”
Renewable Energy, vol. 145, p. 1334–1347, 2020.
[29] C. Eskilsson, J. Palm, J. P. Kofoed, and E. Friis-Madsen, “Cfd study of the overtopping discharge of the wave dragon
wave energy converter,” Renewable Energies Offshore, p. 287–294, 2015.
[30] J. Roenby, H. Bredmose, and H. Jasak, “A computational method for sharp interface advection,” Royal Society open
science, vol. 3, no. 11, p. 160405, 2016.
[31] S. S. Deshpande, L. Anumolu, and M. F. Trujillo, “Evaluating the performance of the two-phase flow solver interfoam,”
Computational science & discovery, vol. 5, no. 1, p. 014016, 2012.
[32] B. E. Larsen, D. R. Fuhrman, and J. Roenby, “Performance of interfoam on the simulation of progressive waves,”
Coastal Engineering Journal, vol. 61, no. 3, p. 380–400, 2019.
[33] V. Vukčević, J. Roenby, I. Gatin, and H. Jasak, “A sharp free surface finite volume method applied to gravity wave
flows,” arXiv preprint arXiv:1804.01130, 2018.
[34] V. Vukčević, H. Jasak, and I. Gatin, “Implementation of the ghost fluid method for free surface flows in polyhedral
finite volume framework,” Computers & fluids, vol. 153, p. 1–19, 2017.
[35] P. Peltonen, P. Kanninen, E. Laurila, and V. Vuorinen, “The ghost fluid method for openfoam: A comparative study
in marine context,” Ocean Engineering, vol. 216, p. 108007, 2020.
[36] D. Files. Interface capturing in openfoam. [Online]. Available: https://cfd.direct/openfoam/free-software/
multiphase-interface-capturing/
[37] J. S. Piña, D. Godino, S. Corzo, and D. Ramajo, “Air injection in vertical water column: Experimental test and
numerical simulation using volume of fluid and two-fluid methods,” Chemical Engineering Science, p. 117665, 2022.
[38] A. J.-C. de Saint-Venant, “Théorie du mouvement non-permanent des eaux, avec application aux crues des rivieres
eta l’introduction des marées dans leur lit,” CR Acad. Sci. Paris, vol. 73, no. 147–154, p. 5, 1871.
[39] Z. Tuković and H. Jasak, “A moving mesh finite volume interface tracking method for surface tension dominated
interfacial fluid flow,” Computers & fluids, vol. 55, p. 70–84, 2012.
[40] L. Huang and G. Thomas, “Simulation of wave interaction with a circular ice floe,” Journal of Offshore Mechanics
and Arctic Engineering, vol. 141, no. 4, p. 041302, 2019.
[41] L. F. Chen, J. Zang, A. J. Hillis, G. C. J. Morgan, and A. R. Plummer, “Numerical investigation of wave–structure
interaction using openfoam,” Ocean Engineering, vol. 88, p. 91–109, 2014.
[42] ITTC, “Seakeeping experiments,” Recommended Procedures and Guidelines, 2017, container-title: Recommended
Procedures and Guidelines.
[43] Y. Li, B. Larsen, and D. R. Fuhrman, “Reynolds stress turbulence modelling of surf zone breaking waves,” Journal
of Fluid Mechanics, 2022.
[44] Y. Li and D. R. Fuhrman, “Cfd simulation of deep-water wave instabilities involving wave breaking,” Journal of
Offshore Mechanics and Arctic Engineering, p. 1–14, 2021.
[45] S. Lyu, L. Huang, and G. Thomas, “Motions of a floating body induced by rogue waves,” The 14th OpenFOAM
Workshop, 2019.
[46] H. Bredmose and N. G. Jacobsen, “Breaking wave impacts on offshore wind turbine foundations: focused wave groups
and cfd,” in International Conference on Offshore Mechanics and Arctic Engineering, vol. 49118, 2010, p. 397–404.
[47] P. Schmitt and B. Elsaesser, “On the use of openfoam to model oscillating wave surge converters,” Ocean Engineering,
vol. 108, p. 98–104, 2015.
[48] P. Higuera, I. J. Losada, and J. L. Lara, “Three-dimensional numerical wave generation with moving boundaries,”
Coastal Engineering, vol. 101, p. 35–47, 2015.
[49] P. Schmitt and B. Elsaesser, “A review of wave makers for 3d numerical simulations,” VI International Conference
on Computational Methods in Marine Engineering (MARINE 2015), no. 2015, p. 1–10, 2015.
[50] C. Windt, J. Davidson, P. Schmitt, and J. V. Ringwood, “On the assessment of numerical wave makers in cfd
simulations,” Journal of Marine Science and Engineering, vol. 7, no. 2, 2019.

®
[51] N. G. Jacobsen, D. R. Fuhrman, and J. Fredsøe, “A wave generation toolbox for the open-source cfd library: Open-
foam ,” International Journal for Numerical Methods in Fluids, vol. 70, no. 9, p. 1073–1088, 2012.

®
[52] Z. Z. Hu, D. Greaves, and A. Raby, “Numerical wave tank study of extreme waves and wave-structure interaction
using openfoam ,” Ocean Engineering, vol. 126, p. 329–342, 2016.
[53] T. Vyzikas, S. Deshoulières, O. Giroux, M. Barton, and D. Greaves, “Numerical study of fixed oscillating water
column with rans-type two-phase cfd model,” Renewable energy, vol. 102, p. 294–305, 2017.
[54] N. Bruinsma, B. T. Paulsen, and N. G. Jacobsen, “Validation and application of a fully nonlinear numerical wave
tank for simulating floating offshore wind turbines,” Ocean Engineering, vol. 147, p. 647–658, 2018.
138 HUANG, LI, BENITES-MUNOZ, WINDT, FEICHTNER, TAVAKOLI, DAVIDSON, PAREDES, QUINTUNA, ET AL.

[55] N. Bruinsma, “Validation and application of a fully nonlinear numerical wave tank,” Master Thesis, TU Delft, 2016.
[56] IHCantabria, “Ihfoam manual,” 2014. [Online]. Available: https://ihfoam.ihcantabria.com/

®
[57] B. Devolder, P. Rauwoens, and P. Troch, “Application of a buoyancy-modified k-ω sst turbulence model to simulate
wave run-up around a monopile subjected to regular waves using openfoam ,” Coastal Engineering, vol. 125, p.
81–94, 2017.

®
[58] H. G. Guler, C. Baykal, T. Arikawa, and A. C. Yalciner, “Numerical assessment of tsunami attack on a rubble mound
breakwater using openfoam ,” Applied Ocean Research, vol. 72, p. 76–91, 2018.
[59] C. Windt, N. Faedo, M. Penalba, F. Dias, and J. V. Ringwood, “Reactive control of wave energy devices–the modelling
paradox,” Applied Ocean Research, vol. 109, p. 102574, 2021.

®
[60] P. Higuera, J. L. Lara, and I. J. Losada, “Realistic wave generation and active wave absorption for navier-stokes
models. application to openfoam .” Coastal Engineering, vol. 71, p. 102–118, 2013.
[61] N. G. Jacobsen, “Theoretical model of an absorbing wave boundary condition,” Tech. Rep. DOI: 10.13140/RG. 2.2.
22360.62726, Private work, Tech. Rep., 2021.
[62] H. A. Schäffer and G. Klopman, “Review of multidirectional active wave absorption methods,” Journal of waterway,
port, coastal, and ocean engineering, vol. 126, no. 2, p. 88–97, 2000.
[63] P. Higuera, “Enhancing active wave absorption in rans models,” Applied Ocean Research, vol. 94, p. 102000, 2020.
[64] M. Borsboom and N. G. Jacobsen, “A generating-absorbing boundary condition for dispersive waves,” International
Journal for Numerical Methods in Fluids, vol. 93, no. 8, p. 2443–2467, 2021.
[65] K. O. Connell and A. Cashman, “Development of a numerical wave tank with reduced discretization error,” in
2016 International Conference on Electrical, Electronics, and Optimization Techniques (ICEEOT). IEEE, 2016, p.
3008–3012.
[66] B. Pena and L. Huang, “A review on the turbulence modelling strategy for ship hydrodynamic simulations,” Ocean
Engineering, 2021.
[67] J. Davidson and R. Costello, “Efficient nonlinear hydrodynamic models for wave energy converter design—a scoping
study,” Journal of Marine Science and Engineering, vol. 8, no. 1, p. 35, 2020.
[68] D. Khojasteh, S. Tavakoli, A. Dashtimanesh, A. Dolatshah, L. Huang, W. Glamore, M. Sadat-Noori, and G. Iglesias,
“Numerical analysis of shipping water impacting a step structure,” Ocean Engineering, vol. 209, p. 107517, 2020.
[69] C. Windt, J. Davidson, and J. V. Ringwood, “Investigation of turbulence modeling for point-absorber-type wave
energy converters,” Energies, vol. 14, no. 1, p. 26, 2021.
[70] S. A. Brown, D. M. Greaves, V. Magar, and D. C. Conley, “Evaluation of turbulence closure models under spilling
and plunging breakers in the surf zone,” Coastal Engineering, vol. 114, p. 177–193, 2016.
[71] B. Devolder, V. Stratigaki, P. Troch, and P. Rauwoens, “Cfd simulations of floating point absorber wave energy
converter arrays subjected to regular waves,” Energies, vol. 11, no. 33, p. 641, Mar 2018.
[72] B. E. Larsen and D. R. Fuhrman, “On the over-production of turbulence beneath surface waves in reynolds-averaged
navier–stokes models,” Journal of Fluid Mechanics, vol. 853, p. 419–460, 2018.
[73] D. R. Fuhrman and Y. Li, “Instability of the realizable k–ϵ turbulence model beneath surface waves,” Physics of
Fluids, vol. 32, no. 11, p. 115108, 2020.
[74] Y. Li, M. B. Fredberg, B. E. Larsen, and D. R. Fuhrman, “Simulating breaking waves with the reynolds stress
turbulence model,” Coastal Engineering Proceedings, no. 36v, p. 17–17, 2020.
[75] Y. Li, The Reynolds Stress Turbulence Model. https://github.com/LiYZPearl/ReynoldsStressTurbulenceModels.:
github, 2021.
[76] Y. Li, M. C. Ong, and T. Tang, “Numerical analysis of wave-induced poro-elastic seabed response around a hexagonal
gravity-based offshore foundation,” Coastal Engineering, vol. 136, p. 81–95, 2018.
[77] M. Mortazavi, V. Le Chenadec, P. Moin, and A. Mani, “Direct numerical simulation of a turbulent hydraulic jump:
turbulence statistics and air entrainment,” Journal of Fluid Mechanics, vol. 797, p. 60–94, 2016.
[78] P. Lopes, J. Leandro, and R. F. Carvalho, “Self-aeration modelling using a sub-grid volume-of-fluid model,” Interna-
tional Journal of Nonlinear Sciences and Numerical Simulation, vol. 18, no. 7–8, p. 559–574, 2017.
[79] J. Ma, A. A. Oberai, D. A. Drew, R. T. Lahey Jr, and M. C. Hyman, “A comprehensive sub-grid air entrainment
model for rans modeling of free-surface bubbly flows,” The Journal of Computational Multiphase Flows, vol. 3, no. 1,
p. 41–56, 2011.
[80] P. D. Tomaselli, “A methodology for air entrainment in breaking waves and their interaction with a mono-pile,” PhD
Thesis, Technical University of Denmark Lyngby, Denmark, 2016.
[81] G. Morgan and J. Zang, “Application of openfoam to coastal and offshore modelling,” The 26th IWWWFB. Athens,
Greece, 2011.

®
[82] W. Chen, J. J. Warmink, M. R. A. Van Gent, and S. Hulscher, “Numerical modelling of wave overtopping at dikes
using openfoam ,” Coastal Engineering, vol. 166, p. 103890, 2021.

®
[83] B. Di Paolo, J. L. Lara, G. Barajas, and I. J. Losada, “Wave and structure interaction using multi-domain couplings
for navier-stokes solvers in openfoam . part i: Implementation and validation,” Coastal Engineering, vol. 164, p.
103799, Mar 2021.

®
[84] B. Di Paolo, J. L. Lara, G. Barajas, and Í. J. Losada, “Waves and structure interaction using multi-domain couplings
for navier-stokes solvers in openfoam . part ii: Validation and application to complex cases,” Coastal Engineering,
vol. 164, p. 103818, 2021.
[85] G. Ducrozet, F. Bonnefoy, D. Le Touzé, and P. Ferrant, “A modified high-order spectral method for wavemaker
modeling in a numerical wave tank,” European Journal of Mechanics-B/Fluids, vol. 34, p. 19–34, 2012.
[86] B. Quinn, “Validation of the high order spectral (hos) method for extreme and breaking waves and coupling of the
hos-numerical wave tank model with openfoam,” Master’s Thesis, University of Stavanger, Norway, 2019.
[87] G. Decorte, A. Toffoli, G. Lombaert, and J. Monbaliu, “On the use of a domain decomposition strategy in obtaining
response statistics in non-gaussian seas,” Fluids, vol. 6, no. 1, p. 28, 2021.
[88] B. Pena and L. Huang, “Wave-gan: A deep learning approach for the prediction of nonlinear regular wave loads and
run-up on a fixed cylinder,” Coastal Engineering, vol. 167, p. 103902, 2021.
A REVIEW ON THE MODELLING OF WAVE-STRUCTURE INTERACTIONS BASED ON OPENFOAM 139

[89] S. Daniels, A. Rahat, G. Tabor, J. Fieldsend, and R. Everson, “Application of multi-objective bayesian shape opti-
misation to a sharp-heeled kaplan draft tube,” Optimization and Engineering, vol. 23, no. 2, pp. 689–716, 2022.
[90] S. Tavakoli and A. V. Babanin, “Wave energy attenuation by drifting and non-drifting floating rigid plates,” Ocean
Engineering, vol. 226, p. 108717, Apr 2021.
[91] J. Palm and C. Eskilsson, “Facilitating large-amplitude motions of wave energy converters in openfoam by a modified
mesh morphing approach,” in 14th European Wave and Tidal Energy Conference. European Tidal and Wave Energy
Conference, 2021.
[92] J. Palm, C. Eskilsson, G. M. Paredes, and L. Bergdahl, “Coupled mooring analysis for floating wave energy converters
using cfd: Formulation and validation,” International Journal of Marine Energy, vol. 16, p. 83–99, 2016.
[93] H. Islam, S. C. Mohapatra, J. Gadelho, and C. G. Soares, “Openfoam analysis of the wave radiation by a box-type
floating structure,” Ocean Engineering, vol. 193, p. 106532, 2019.
[94] M. Yan, X. Ma, W. Bai, Z. Lin, and Y. Li, “Numerical simulation of wave interaction with payloads of different
postures using openfoam,” Journal of Marine Science and Engineering, vol. 8, no. 6, p. 433, 2020.
[95] S. C. Mohapatra, H. Islam, and C. Guedes Soares, “Boussinesq model and cfd simulations of non-linear wave diffraction
by a floating vertical cylinder,” Journal of Marine Science and Engineering, vol. 8, no. 8, p. 575, 2020.
[96] M. Yousefifard and A. Graylee, “A numerical solution of the wave–body interactions for a freely floating vertical
cylinder in different water depths using openfoam,” Journal of the Brazilian Society of Mechanical Sciences and
Engineering, vol. 43, no. 1, p. 1–14, 2021.
[97] D. Benites-Munoz, L. Huang, E. Anderlini, J. R. Marı́n-Lopez, and G. Thomas, “Hydrodynamic modelling of an
oscillating wave surge converter including power take-off,” Journal of Marine Science and Engineering, vol. 8, no. 10,
p. 771, 2020.
[98] L. Huang, S. Tavakoli, M. Li, A. Dolatshah, B. Pena, B. Ding, and A. Dashtimanesh, “Cfd analyses on the water
entry process of a freefall lifeboat,” Ocean Engineering, vol. 232, p. 109115, 2021.
[99] H. Chen, L. Qian, Z. Ma, W. Bai, Y. Li, D. Causon, and C. Mingham, “Application of an overset mesh based numerical
wave tank for modelling realistic free-surface hydrodynamic problems,” Ocean Engineering, vol. 176, p. 97–117, 2019.
[100] D. D. Chandar, “On overset interpolation strategies and conservation on unstructured grids in openfoam,” Computer
Physics Communications, vol. 239, p. 72–83, 2019.
[101] T. Wu, W. Luo, D. Jiang, R. Deng, and S. Huang, “Numerical study on wave-ice interaction in the marginal ice
zone,” Journal of Marine Science and Engineering, vol. 9, no. 1, p. 4, 2021.
[102] S. Tavakoli, L. Huang, and A. V. Babanin, “Drift motion of floating bodies under the action of green water,” in
International Conference on Offshore Mechanics and Arctic Engineering, vol. 85161. American Society of Mechanical
Engineers, 2021, p. V006T06A013.
[103] C. Windt, J. Davidson, E. J. Ransley, D. Greaves, M. Jakobsen, M. Kramer, and J. V. Ringwood, “Validation of
a cfd-based numerical wave tank model for the power production assessment of the wavestar ocean wave energy
converter,” Renewable Energy, vol. 146, p. 2499–2516, 2020.

® ® ®
[104] A. Hager, C. Kloss, and C. Goniva, “Combining open source and easy access in the field of dem and coupled cfd-dem:
Liggghts , cfdem coupling and cfdem workbench,” in Computer Aided Chemical Engineering. Elsevier, 2018,
vol. 43, pp. 1699–1704.
[105] L. Huang, J. Tuhkuri, B. Igrec, M. Li, D. Stagonas, A. Toffoli, P. Cardiff, and G. Thomas, “Ship resistance when
operating in floating ice floes: A combined cfd&dem approach,” Marine Structures, vol. 74, p. 102817, 2020.
[106] I. Collins, M. Hossain, W. Dettmer, and I. Masters, “Flexible membrane structures for wave energy harvesting: A
review of the developments, materials and computational modelling approaches,” Renewable and Sustainable Energy
Reviews, vol. 151, p. 111478, 2021.
[107] L. Huang and Y. Li, “Design of the submerged horizontal plate breakwater using a fully coupled hydroelastic ap-
proach,” Computer-Aided Civil and Infrastructure Engineering, 2022.
[108] L. Huang, K. Ren, M. Li, Z. Tuković, P. Cardiff, and G. Thomas, “Fluid-structure interaction of a large ice sheet in
waves,” Ocean Engineering, vol. 182, p. 102–111, 2019, container-title: Ocean Engineering.
[109] N. G. Jacobsen, W. Bakker, W. S. Uijttewaal, and R. Uittenbogaard, “Experimental investigation of the wave-induced
motion of and force distribution along a flexible stem,” Journal of Fluid Mechanics, vol. 880, p. 1036–1069, 2019.
[110] M. Masoomi and A. Mosavi, “The one-way fsi method based on rans-fem for the open water test of a marine propeller
at the different loading conditions,” Journal of Marine Science and Engineering, vol. 9, no. 4, p. 351, 2021.
[111] H. Chen, Q. Zou, and Z. Liu, “A coupled rans-vof and finite element model for wave interaction with highly flexible
vegetation,” COASTAL ENGINEERING, p. 2, 2016.
[112] A. Karac, “Drop impact of fluid-filled polyethylene containers,” PhD Thesis, Imperial College London (University of
London), 2003.
[113] C.-G. Giannopapa, “Fluid structure interaction in flexible vessels,” PhD Thesis, University of London, 2004.
[114] C. J. Greenshields and H. G. Weller, “A unified formulation for continuum mechanics applied to fluid–structure inter-
action in flexible tubes,” International Journal for Numerical Methods in Engineering, vol. 64, no. 12, p. 1575–1593,
2005.
[115] Z. Tukovic and H. Jasak, “Updated lagrangian finite volume solver for large deformation dynamic response of elastic
body,” Transactions of FAMENA, vol. 31, no. 1, p. 55, 2007.
[116] Z. Tukovic, P. Cardiff, A. Karac, H. Jasak, and A. Ivankovic, “Openfoam library for fluid structure interaction,” in
9th OpenFOAM Workshop, vol. 2014, 2014.
[117] Z. Tuković, A. Ivanković, and A. Karač, “Finite-volume stress analysis in multi-material linear elastic body,” Inter-
national journal for numerical methods in engineering, vol. 93, no. 4, p. 400–419, 2013.
[118] J. McVicar, J. Lavroff, M. R. Davis, and G. Thomas, “Fluid–structure interaction simulation of slam-induced bending
in large high-speed wave-piercing catamarans,” Journal of Fluids and Structures, vol. 82, p. 35–58, 2018.
[119] L. Huang, “An opensource solver for wave-induced fsi problems,” In proceedings of CFD with OpenSource Software.
Edited by Nilsson. H., 2018.
140 HUANG, LI, BENITES-MUNOZ, WINDT, FEICHTNER, TAVAKOLI, DAVIDSON, PAREDES, QUINTUNA, ET AL.

[120] Y. Li, Z. Hu, and L. Huang, “Hydroelastic simulation of breaking wave impact on a flexible coastal seawall,” in The
41th International Conference on Ocean, Offshore & Arctic Engineering (OMAE). ASME: The American Society
of Mechanical Engineers, 2022.
[121] P. Cardiff, A. Karač, P. De Jaeger, H. Jasak, J. Nagy, A. Ivanković, and Z. Tuković, “An open-source finite volume
toolbox for solid mechanics and fluid-solid interaction simulations,” arXiv preprint arXiv:1808.10736, 2018.
[122] T. Tang, O. Hededal, and P. Cardiff, “On finite volume method implementation of poro-elasto-plasticity soil model,”
International journal for numerical and analytical methods in geomechanics, vol. 39, no. 13, p. 1410–1430, 2015.
[123] I. Oliveira, J. Gasche, and P. Cardiff, “Implementation and numerical verification of an incompressible three-parameter
mooney-rivlin model for large deformation of soft tissues,” in The 15th OpenFOAM Workshop, 2020.
[124] M. Girfoglio, A. Quaini, and G. Rozza, “Fluid-structure interaction simulations with a les filtering approach in
solids4foam,” arXiv preprint arXiv:2102.08011, 2021.
[125] M. Abusara, N. Hakan, P. Cardiff, and D. Randstrom, “Fluid-structure interaction on a fixed fan blade,” in The 16th
OpenFOAM Workshop, 2021.
[126] I. J. Losada, J. L. Lara, and M. del Jesus, “Modeling the interaction of water waves with porous coastal structures,”
Journal of Waterway, Port, Coastal, and Ocean Engineering, vol. 142, no. 6, p. 03116003, 2016.
[127] F. Mentzoni and T. Kristiansen, “A semi-analytical method for calculating the hydrodynamic force on perforated
plates in oscillating flow,” in International Conference on Offshore Mechanics and Arctic Engineering, vol. 58844.
American Society of Mechanical Engineers, 2019, p. V07AT06A015.
[128] A. George and I. H. Cho, “Anti-sloshing effects of a vertical porous baffle in a rolling rectangular tank,” Ocean
Engineering, vol. 214, p. 107871, 2020.
[129] S. K. Poguluri and I. H. Cho, “Liquid sloshing in a rectangular tank with vertical slotted porous screen: based on
analytical, numerical, and experimental approach,” Ocean Engineering, vol. 189, p. 106373, 2019.
[130] K. Shim, P. Klebert, and A. Fredheim, “Numerical investigation of the flow through and around a net cage,” in
International Conference on Offshore Mechanics and Arctic Engineering, vol. 43444, 2009, p. 581–587.
[131] H. Chen and E. D. Christensen, “Investigations on the porous resistance coefficients for fishing net structures,” Journal
of Fluids and Structures, vol. 65, p. 76–107, 2016.
[132] A. Feichtner, E. Mackay, G. Tabor, P. R. Thies, L. Johanning, and D. Ning, “Using a porous-media approach for cfd
modelling of wave interaction with thin perforated structures,” Journal of Ocean Engineering and Marine Energy,
vol. 7, no. 1, p. 1–23, 2021.
[133] B. Jensen, N. G. Jacobsen, and E. D. Christensen, “Investigations on the porous media equations and resistance
coefficients for coastal structures,” Coastal Engineering, vol. 84, p. 56–72, 2014.

®
[134] P. Higuera, J. L. Lara, and I. J. Losada, “Three-dimensional interaction of waves and porous coastal structures using
openfoam . part i: Formulation and validation,” Coastal Engineering, vol. 83, p. 243–258, 2014.
[135] S. Hadadpour, M. Paul, and H. Oumeraci, “Numerical investigation of wave attenuation by rigid vegetation based on
a porous media approach,” Journal of Coastal Research, vol. 92, no. SI, p. 92–100, 2019.
[136] M. Maza, J. L. Lara, and I. J. Losada, “Tsunami wave interaction with mangrove forests: A 3-d numerical approach,”
Coastal Engineering, vol. 98, p. 33–54, 2015.
[137] A. Feichtner, E. Mackay, G. Tabor, P. R. Thies, and L. Johanning, “Comparison of macro-scale porosity implementa-
tions for cfd modelling of wave interaction with thin porous structures,” Journal of Marine Science and Engineering,
vol. 9, no. 2, p. 150, 2021.

®
[138] A. Romano, J. Lara, G. Barajas, B. Di Paolo, G. Bellotti, M. Di Risio, I. Losada, and P. De Girolamo, “Numerical
modelling of landslide-generated tsunamis with openfoam : a new approach,” Coastal Structures 2019, p. 486–495,
2019.
[139] A. Feichtner, G. Tabor, E. Mackay, P. Thies, and L. Johanning, “On the use of a porous-media approach for the
modelling of wave interaction with thin perforated cylinders,” in The 15th OpenFOAM Workshop (OFW15), Online,
2020.

®
[140] P. Higuera, J. L. Lara, and I. J. Losada, “Three-dimensional interaction of waves and porous coastal structures using
openfoam . part ii: Application,” Coastal Engineering, vol. 83, pp. 259–270, 2014.
[141] P. Lin, Numerical modeling of breaking waves. Cornell University, 1998.
[142] P. L.-F. Liu, P. Lin, K.-A. Chang, and T. Sakakiyama, “Numerical modeling of wave interaction with porous struc-
tures,” Journal of waterway, port, coastal, and ocean engineering, vol. 125, no. 6, p. 322–330, 1999.
[143] Y.-T. Wu and S.-C. Hsiao, “Propagation of solitary waves over a submerged slotted barrier,” Journal of Marine
Science and Engineering, vol. 8, no. 6, p. 419, 2020.
[144] K.-H. Lee, J.-H. Bae, S.-G. Kim, and D.-S. Kim, “Three-dimensional simulation of wave reflection and pressure acting
on circular perforated caisson breakwater by olafoam,” Journal of Korean Society of Coastal and Ocean Engineers,
vol. 29, no. 6, p. 286–304, 2017.
[145] V. K. Srineash, A. Kamath, K. Murali, and H. Bihs, “Numerical simulation of wave interaction with submerged
porous structures and application for coastal resilience,” Journal of Coastal Research, vol. 36, no. 4, p. 752–770, 2020.
[146] M. Brito, J. Fernandes, and J. B. Leal, “Porous media approach for rans simulation of compound open-channel flows
with submerged vegetated floodplains,” Environmental Fluid Mechanics, vol. 16, no. 6, p. 1247–1266, 2016.
[147] Y.-P. Zhao, C.-W. Bi, Y.-X. Liu, G.-H. Dong, and F.-K. Gui, “Numerical simulation of interaction between waves
and net panel using porous media model,” Engineering Applications of Computational Fluid Mechanics, vol. 8, no. 1,
p. 116–126, 2014.
[148] J. Chauchat, Z. Cheng, T. Nagel, C. Bonamy, and T.-J. Hsu, “Sedfoam-2.0: a 3-d two-phase flow numerical model
for sediment transport,” Geoscientific Model Development, vol. 10, no. 12, p. 4367–4392, 2017.
[149] X. Liu, M. H. Garcı́a, and R. Muscari, “Numerical investigation of seabed response under waves with free-surface
water flow,” International Journal of Offshore and Polar Engineering, vol. 17, no. 02, 2007.
[150] M. A. Biot, “General theory of three-dimensional consolidation,” Journal of applied physics, vol. 12, no. 2, p. 155–164,
1941.
A REVIEW ON THE MODELLING OF WAVE-STRUCTURE INTERACTIONS BASED ON OPENFOAM 141

[151] T. Tang and B. Johannesson, “An integrated fvm simulation of wave-seabed-structure interaction using openfoam,”
in 9th OpenFOAM Workshop, 23-26 June 2014 in Zagreb, Croatia, 2014.

®
[152] Y. Li, M. C. Ong, and T. Tang, “A numerical toolbox for wave-induced seabed response analysis around marine
structures in the openfoam framework,” Ocean Engineering, vol. 195, p. 106678, 2020.
[153] Y. Li, A CFD toolbox for wave-structure-soil interaction simulation and liquefaction assessment.
https://github.com/LiYZPearl/wssi.: github, 2021.
[154] Z. Lin, D. Pokrajac, Y. Guo, D.-s. Jeng, T. Tang, N. Rey, J. Zheng, and J. Zhang, “Investigation of nonlinear
wave-induced seabed response around mono-pile foundation,” Coastal Engineering, vol. 121, p. 197–211, 2017.
[155] D. Celli, Y. Li, M. C. Ong, and M. Di Risio, “The role of submerged berms on the momentary liquefaction around
conventional rubble mound breakwaters,” Applied Ocean Research, vol. 85, p. 1–11, 2019.
[156] Z. Liang and D.-S. Jeng, “Poro-fssi-foam model for seafloor liquefaction around a pipeline under combined random
wave and current loading,” Applied Ocean Research, vol. 107, p. 102497, 2021.
[157] J. Ye, K. He, and L. Zhou, “Subsidence prediction of a rubble mound breakwater at yantai port: A application of
fssi-cas 2d,” Ocean Engineering, vol. 219, p. 108349, 2021.
[158] H. Elsafti and H. Oumeraci, “A numerical hydro-geotechnical model for marine gravity structures,” Computers and
Geotechnics, vol. 79, p. 105–129, 2016.

®
[159] R. khumar Shanmugasundaram, H. Rusche, C. Windt, O. Kirca, M. Sumer, and N. Goseberg, “Towards the numerical
modelling of residual seabed liquefaction using openfoam,” OpenFOAM Journal, vol. 2, p. 94–115, 2022.
[160] C. Windt, J. Davidson, and J. V. Ringwood, “High-fidelity numerical modelling of ocean wave energy systems: A
review of computational fluid dynamics-based numerical wave tanks,” Renewable and Sustainable Energy Reviews,
vol. 93, p. 610–630, Oct 2018.
[161] G. Giorgi and J. Ringwood, “Consistency of viscous drag identification tests for wave energy applications,” in Pro-
ceedings of the 12th European Wave and Tidal Energy Conference 27th Aug-1st Sept 2017, no. 643. European Wave
and Tidal Energy Conference 2017, 2017, p. 1–8.
[162] C. Windt, J. Davidson, D. D. Chandar, N. Faedo, and J. V. Ringwood, “Evaluation of the overset grid method for
control studies of wave energy converters in openfoam numerical wave tanks,” Journal of Ocean Engineering and
Marine Energy, vol. 6, no. 1, p. 55–70, 2020.
[163] E. J. Ransley, D. M. Greaves, A. Raby, D. Simmonds, M. M. Jakobsen, and M. Kramer, “Rans-vof modelling of the
wavestar point absorber,” Renewable Energy, vol. 109, p. 49–65, Aug 2017.
[164] J. Palm and C. Eskilsson, “Moody user manual: version 1.0. 0,” Chalmers University of Technology, Department of
Mechanics and Maritime Sciences, Tech. Rep., 2018.
[165] B. W. Schubert, F. Meng, N. Y. Sergiienko, W. Robertson, B. S. Cazzolato, M. H. Ghayesh, A. Rafiee, B. Ding, and
M. Arjomandi, “Pseudo-nonlinear hydrodynamic coefficients for modelling point absorber wave energy converters,”
in Proceedings of the 4th Asian Wave and Tidal Energy Conference, Taipei, Taiwan, 2018, p. 9–13.
[166] F. Meng, A. Rafiee, B. Ding, B. Cazzolato, and M. Arjomandi, “Nonlinear hydrodynamics analysis of a submerged
spherical point absorber with asymmetric mass distribution,” Renewable Energy, vol. 147, p. 1895–1908, Mar 2020.
[167] C. Windt, J. Davidson, and J. V. Ringwood, “Numerical analysis of the hydrodynamic scaling effects for the wavestar
wave energy converter,” Journal of Fluids and Structures, vol. 105, p. 103328, Aug 2021.
[168] H. Akimoto, K. Tanaka, and Y. Y. Kim, “Drag-type cross-flow water turbine for capturing energy from the orbital
fluid motion in ocean wave,” Renewable Energy, vol. 76, p. 196–203, Apr 2015.
[169] C. Xu and Z. Huang, “Three-dimensional cfd simulation of a circular owc with a nonlinear power-takeoff: Model
validation and a discussion on resonant sloshing inside the pneumatic chamber,” Ocean Engineering, vol. 176, p.

®
184–198, Mar 2019.
[170] A. Iturrioz, R. Guanche, J. L. Lara, C. Vidal, and I. J. Losada, “Validation of openfoam for oscillating water
column three-dimensional modeling,” Ocean Engineering, vol. 107, p. 222–236, Oct 2015.
[171] Z. Deng, C. Wang, P. Wang, P. Higuera, and R. Wang, “Hydrodynamic performance of an offshore-stationary owc
device with a horizontal bottom plate: Experimental and numerical study,” Energy, vol. 187, p. 115941, Nov 2019.
[172] A. Dimakopoulos, M. Cooker, E. Medina-Lopez, D. Longo, and R. Pinguet, “Flow characterisation and numerical
modelling of owc wave energy converters,” in 11th European Wave & Tidal Energy Conference (EWTEC). Nantes,
France, 2015.
[173] J. Gadelho, C. G. Soares, G. Barajas, and J. L. Lara, “Cfd analysis of the pto damping on the performance of an
onshore dual chamber owc,” Trends in Maritime Technology and Engineering Volume 2, p. 381–389, 2022.
[174] K. Rezanejad, J. F. M. Gadelho, and C. Guedes Soares, “Hydrodynamic analysis of an oscillating water column wave
energy converter in the stepped bottom condition using cfd,” Renewable Energy, vol. 135, p. 1241–1259, May 2019.
[175] I. Simonetti, L. Cappietti, H. El Safti, and H. Oumeraci, “Numerical modelling of fixed oscillating water column
wave energy conversion devices: Toward geometry hydraulic optimization,” in International Conference on Offshore
Mechanics and Arctic Engineering, vol. 56574. American Society of Mechanical Engineers, 2015, p. V009T09A031.
[176] I. Simonetti, L. Cappietti, H. Elsafti, and H. Oumeraci, “Optimization of the geometry and the turbine induced
damping for fixed detached and asymmetric owc devices: A numerical study,” Energy, vol. 139, p. 1197–1209, Nov
2017.
[177] L. F. Chen, L. Sun, L. Zang, A. J. Hillis, and A. R. Plummer, “Numerical study of roll motion of a 2-d floating
structure in viscous flow,” Journal of Hydrodynamics, Ser. B, vol. 28, no. 4, p. 544–563, 2016.
[178] E. Renzi, S. Michele, S. Zheng, S. Jin, and D. Greaves, “Niche applications and flexible devices for wave energy
conversion: A review,” Energies, vol. 14, no. 20, p. 6537, 2021.
[179] L. Huang, Z. Hu, Y. Li, and G. Thomas, “Fully-coupled cfd+csm analysis on an elastic floating/submerged plate
for wave energy harvest,” in 9th International Conference on HYDROELASTICITY IN MARINE TECHNOLOGY,
Rome, Italy, 2022.
[180] T. T. Loh, D. Greaves, T. Maeki, M. Vuorinen, D. Simmonds, and A. Kyte, “Numerical modelling of the waveroller
device using openfoam,” in Proceedings of the 3rd Asian Wave & Tidal Energy Conference, 2016.
142 HUANG, LI, BENITES-MUNOZ, WINDT, FEICHTNER, TAVAKOLI, DAVIDSON, PAREDES, QUINTUNA, ET AL.

[181] A. Rafiee and J. Fiévez, “Numerical prediction of extreme loads on the ceto wave energy converter,” in Proceedings
of the 11th European Wave and Tidal Energy Conference, Nantes, France, no. 09A1-2, 2015.
[182] M. Penalba, J. Davidson, C. Windt, and J. V. Ringwood, “A high-fidelity wave-to-wire simulation platform for wave
energy converters: Coupled numerical wave tank and power take-off models,” Applied Energy, vol. 226, p. 655–669,
Sep 2018.
[183] C. Jiang, O. el Moctar, G. Moura Paredes, and T. E. Schellin, “Validation of a dynamic mooring model coupled with
a rans solver,” Marine Structures, vol. 72, p. 102783, Jul 2020.
[184] M. Hall, “Moordyn v2: New capabilities in mooring system components and load cases,” in International Conference
on Offshore Mechanics and Arctic Engineering, vol. 84416. American Society of Mechanical Engineers, 2020, p.
V009T09A078.
[185] H. Chen and M. Hall, “Cfd simulation of floating body motion with mooring dynamics: Coupling moordyn with
openfoam,” Applied Ocean Research, vol. 124, p. 103210, 2022.
[186] J. Palm, C. Eskilsson, and L. Bergdahl, “An hp-adaptive discontinuous galerkin method for modelling snap loads in
mooring cables,” Ocean Engineering, vol. 144, p. 266–276, 2017.
[187] G. Barajas, J. L. Lara, B. Di Paolo, and I. J. Losada, “Analysis of a floating wave energy converter interaction
with waves using the overset framework,” in The 9th Conference on Computational Methods in Marine Engineering
(Marine 2021), 2022.

You might also like