You are on page 1of 28

energies

Article
A Numerical Investigation of the Hydrodynamic Performance of
a Pitch-Type Wave Energy Converter Using Weakly and Fully
Nonlinear Models
Sunny Kumar Poguluri 1 , Dongeun Kim 2 and Yoon Hyeok Bae 1, *

1 Department of Mechanical and System Design Engineering, Hongik University,


Seoul 04066, Republic of Korea; sunnykumar@hongik.ac.kr
2 Multidisciplinary Graduate School Program for Wind Energy, Jeju National University,
Jeju 63243, Republic of Korea; d1259@jejunu.ac.kr
* Correspondence: yhbae@hongik.ac.kr

Abstract: In this study, the performance of a wave energy converter (WEC) rotor under regular and
irregular wave conditions was investigated using 3D nonlinear numerical models. Factors such as
the power take-off (PTO) load torque, wave periods, spacing of multiple WEC rotors, and wave
steepness were analyzed. Two models were employed: a weakly nonlinear model formulated by
incorporating the nonlinear restoring moment and Coulomb-type PTO load torque based on the
potential flow theory, and a fully nonlinear model based on computational fluid dynamics. The
results show that the average power estimated by both numerical models is consistent, with a wave
steepness of 0.03 for the range of one-way and two-way PTO load torques, except for the deviations
observed in the long range of the one-way PTO load torque. Furthermore, the average power of the
WEC rotor under the applied PTO load torque exhibits a quadratic dependency, regardless of the
wave steepness. In addition, adopting a one-way PTO load torque was more efficient than adopting a
two-way PTO load torque. Therefore, the fully nonlinear model demonstrated its ability to handle a
high degree of nonlinearity, surpassing the limitations of the weakly nonlinear model, which was
limited to moderate wave steepness.

Keywords: WEC rotor; nonlinear restoring moment; coulomb type PTO load torque (one-way and
Citation: Poguluri, S.K.; Kim, D.; Bae,
two-way); potential flow theory; computational fluid dynamics; absorbed power
Y.H. A Numerical Investigation of the
Hydrodynamic Performance of a
Pitch-Type Wave Energy Converter
Using Weakly and Fully Nonlinear
1. Introduction
Models. Energies 2024, 17, 898.
https://doi.org/10.3390/en17040898 Ocean energy resources have attracted increasing attention in recent decades and
are now prioritized worldwide because they can help reduce environmental impacts and
Academic Editor: Patrick G. Verdin
global warming. Wave energy converters (WECs) are important in the renewable energy
Received: 19 January 2024 sector owing to their predictability and stability of wave energy [1]. Wave excitation can
Revised: 8 February 2024 lead to nonlinear forces and large body motion in the design of WECs with maximum
Accepted: 11 February 2024 power absorption. The accurate estimation of wave loads and large motions is essential
Published: 15 February 2024 and requires nonlinear analysis. Among the existing WECs, the Salter’s duck (WEC
rotor) exhibits the highest efficiency in extracting power from two-dimensional regular
waves [2]. The majority of the initial studies were based on linear potential theory, and
their estimations were performed according to the linear superposition of wave diffraction
Copyright: © 2024 by the authors.
and radiation using Laplace’s equation [3–12]. These linear solutions can be frequency-
Licensee MDPI, Basel, Switzerland.
dependent or time-domain models [11]. Reproducing the behavior of a WEC rotor using
This article is an open access article
linear models may not be sufficient to address large WEC motions and accurate power
distributed under the terms and
production over the entire range of sea conditions. Various studies on WEC have shown
conditions of the Creative Commons
Attribution (CC BY) license (https://
that experimental testing followed by prototype testing in seas is inevitable; however, with
creativecommons.org/licenses/by/
the increase in computational capabilities over the years, nonlinear models can now be
4.0/). adopted [11,13].

Energies 2024, 17, 898. https://doi.org/10.3390/en17040898 https://www.mdpi.com/journal/energies


Energies 2024, 17, 898 2 of 28

Two main approaches can be used to address the nonlinear behavior of WEC: the
use of weakly nonlinear models and the use of fully nonlinear models. Both models
are typically assessed based on time-domain solutions [14]. In weakly nonlinear models,
the hydrodynamic coefficients can be obtained from a linear frequency domain solution,
whereas the force components, including radiation, wave excitation forces, external forces
(such as the power take-off system), and mooring, can be represented as nonlinear quan-
tities. Consequently, these models exhibit weakly nonlinear formulations, offering high
computing speed and low computational cost, [15–26] implemented weakly nonlinear
formulations from instantaneous hydrodynamic forces and moments (from buoyancy and
Froude–Krylov quantities), which were estimated at each position on the wetted surface of
the body during each simulation time step using a linear model. They concluded that for a
WEC device with large-amplitude motion, weakly nonlinear formulations are necessary
to capture the relevant physics. Van’t Hoff [27] simulated linear and weakly nonlinear
models, including linear and nonlinear viscous and power take-off (PTO) damping torques,
in the time-domain solution. High accuracy was obtained from the nonlinear interactions
between the WEC device and the wave. According to Folley et al. [28], incorporating a
nonlinear PTO damping torque, such as Coulomb damping, enables the specification of
forces at discrete time steps rather than relying on regular sinusoidal signals typically used
in frequency-domain models. This approach allows for a more accurate representation
of the system dynamics. Latching is a successful way of increasing the energy extracted
from WEC devices, and many researchers have addressed it using theoretical and experi-
mental bases [29–31]. Latching is the process of mechanically engaging and disengaging
the PTO system to extract wave energy while the WEC device interacts with the wave.
A considerable number of the advanced latching control techniques utilized in WECs is
dedicated to achieving suboptimal conditions. This involves manipulation of the WEC
phase, which refers to the alignment of the motion of the converter with the incoming
waves. Some researchers have implemented this by employing predictive methods [31–33],
whereas others have utilized wave period estimations [34,35]. The application of latching
to irregular wave conditions has not yet been satisfactorily addressed, but the use of a
hydraulic PTO system can provide a natural means of achieving latching [36–38]. In the
hydraulic PTO system, the motion of the WEC device is converted into a fluid flow at high
pressure by the hydraulic piston. This conversion can be made flexible (either unidirectional
or bidirectional) by controlling the piston motion. This type of PTO loading has several
advantages and can be naturally implemented theoretically and practically. Hydraulic PTO
systems can be implemented using Coulomb damping [39–43]. António [36] developed a
weakly nonlinear formulation by incorporating a damping mechanism based on Coulomb
damping. The results showed that energy extraction could be increased under regular and
irregular wave conditions by incorporating Coulomb damping. Furthermore, it is easier to
implement Coulomb damping than existing latching mechanisms. Energy extraction can be
remarkably increased in regular and irregular waves, and it is easier to implement the PTO
system compared with the existing latching mechanisms. Orszaghova et al. [43] studied
the application of linear and Coulomb-type damping forces on a vertically oscillating WEC
device to understand its low-frequency drifting behavior under extreme conditions using a
weakly nonlinear semi-analytical solution. The findings suggest that low-frequency offsets
are more common when the stiffness coefficients are low, affecting both the extreme and
mean behaviors. Additionally, the largest excursions typically occur when the Coulomb
damping is minimal and the spring stiffness is weak, although excessively high Coulomb
damping can also reduce the responses. The experimental data partially support the trends
observed in the simulations.
The aforementioned studies indicate that weakly nonlinear formulations can handle
moderate wave steepness interactions with WEC devices. However, some important physi-
cal phenomena, such as wave fragmentation (breaking), large WEC responses, slamming
effects, turbulent effects, and nonlinear WEC–WEC interactions, cannot be handled by
weakly nonlinear formulations [44,45]. Fully nonlinear formulations based on compu-
Energies 2024, 17, 898 3 of 28

tational fluid dynamics (CFD) may provide improved predictions owing to their ability
to handle most near-field effects and nonlinear wave–WEC interactions [43,46–53]. Ko
et al. [54] conducted a numerical investigation to determine the optimal PTO torque of
an asymmetric WEC for regular and irregular waves. The study was conducted in Open-
FOAM by applying the PTO torque to the WEC in both unidirectional and bidirectional
manners. They reported that unidirectional PTO torque was more efficient than bidirec-
tional PTO torque. Ha et al. [55] conducted a numerical investigation based on STAR-CCM+
to understand the nonlinear dynamic behavior of a WEC rotor. They concluded that at
a wave height of 0.11 m, the CFD results agree well with the experimental results; the
frequency-domain solution and the difference between the two were amplified as the wave
height increased. Wave run-up and slamming are nonlinear aspects that lead to an increase
in the difference.
Three-dimensional (3D) nonlinear numerical models are necessary, and these models
must account for physical flow phenomena that are close to real fluid interactions. In the
context of pitch-type WECs, the extant literature cannot provide a coherent understanding
of the effect of absorbed power owing to 3D nonlinear effects, which is the primary objective
of the present study. Furthermore, this study highlights two types of nonlinear analyses:
weakly nonlinear (a potential-based model using Orcaflex (software package)) and fully
nonlinear (a RANS-based model using CFD). The weakly nonlinear approach incorporates
the nonlinear restoring moment and Coulomb-type PTO load torque, whereas the fully
nonlinear model is based on 3D implicit, unsteady, and incompressible Reynolds-averaged
Navier–Stokes (RANS) equations solved using CFD. In this study, a fully nonlinear model
was used to comprehensively address the hydrodynamic interactions, nonlinear effects, and
performance characteristics by considering the high degree of nonlinearity under 3D wave
conditions, effectively demonstrating the limitations of weakly nonlinear formulations.
The remainder of this paper is organized as follows. Section 2 describes the CFD and
numerical wave tank (NWT) setups. Section 3 presents the implementation of the PTO load
torque and nonlinear restoring moment. The numerical models are validated by comparing
their results with the experimental results in Section 4. Section 5 presents the results and
discussion, and the conclusions are presented in Section 6.

2. Numerical Modeling
2.1. Computational Fluid Dynamics
This study focuses on the hydrodynamics of a fully nonlinear model related to the
interaction between waves and a WEC. It employs the finite-volume-based commercial code
STAR-CCM+. The approach involves solving the Navier–Stokes equations to accurately
model the fluid dynamics. This is executed through unsteady incompressible Navier–Stokes
equations, where the mass and momentum conservation are expressed as

∇·u = 0 (1)
 
∂u 1
+ ∇·(uu) = − ∇ p + ν∇2 u + g (2)
∂t ρ
where u denotes the velocity vector. In Equation (2), the total term on the left side represents
acceleration, the first and second terms on the right side correspond to the surface forces
(pressure and viscosity, respectively), and the third term is the body force influenced purely
by gravity. To account for the turbulence, the viscosity tensor shown in Equation (2) must
include the turbulent viscosity, which necessitates the solution of additional transport
equations using the RANS method. The continuity and momentum equations in the
expanded Cartesian coordinate system in (x, y, z) are expressed as

∂u ∂v ∂w
+ + =0 (3)
∂x ∂y ∂z
Energies 2024, 17, 898 4 of 28

∂ ( u2 )
 2 
∂(ρu) ∂u2 ∂u2
∂t + ∂x + ∂uv
∂y + ∂uw
∂z = − 1ρ ∂p + υ ∂u
∂x2
+ +
∂y2 ∂z2
 ∂x 2 (4)
∂(u′ ) ∂u′ v′ ′ ′
− ∂x + ∂y + ∂u∂zw + g
 2 
∂(ρv) ∂(uv) ∂v2 ∂v2 2
∂t + ∂x + ∂y + ∂vw
∂z = − 1ρ ∂p
∂y + υ ∂x∂v
2 + ∂y2
+ ∂v 2
∂z
(5)

2
∂(u′ v′ ) ′ ∂v′ w′
− ∂x + ∂y +
∂v
∂z +g
 2 
∂(ρw) ∂(uw) ∂w2 ∂w2 ∂w2
∂t + ∂x + ∂vw
∂y + ∂z = − 1ρ ∂p + υ ∂w
2 + 2 + 2
 ∂z ∂x ∂y ∂z
∂(u′ w′ ) 2 (6)
∂v′ w′ ∂w′
− ∂x + ∂y + ∂z + g

In the equations above, the third term on the right-hand side refers to the turbu-
lent/Reynolds stresses, which can be modeled based on one- or two-equation turbulence
models. A two-equation model (low-Re, standard k-ω turbulent model), reported by
Peric [56], was used in this study.

2.2. Numerical Wave Tank


Simulating the interaction between the wave and WEC rotor requires accuracy and
long duration in the NWT. The NWT was solved in right-handed Cartesian coordinates
O(x, y, z), with x pointing to the direction of the incoming waves; y-axis is along the width
of the WEC rotor and z-axis points vertically upward from the still water level (Figure 1).
The origin was located at the center of rotation of the WEC rotor. The properties of the
WEC rotor model and the prototype-scale details are listed in Table 1. A schematic of the
key features of a WEC rotor in a wave tank is shown in Figure 1. A major problem in
CFD-based NWTs is the effective generation and absorption of waves from both sides of the
tank. Different methodologies are available for implementation; however, in this study, the
wave forcing method was employed because it exhibits good overall performance under
various wave conditions [48,56,57]. Gradual forcing was applied over a specified distance
between the Navier and Stokes equations and the theoretical solution. A source term was
added to the momentum equation, as shown in Equation (7).

f δ = −γ f ρ(δcomputed − δ∗f orced ) (7)

where γ f is the forcing coefficient, δ∗f orced is the value toward which the solution is forced,
and δcomputed is the actual momentum equation solution. A large value of the forcing
coefficient was used at the inlet boundary, which gradually varied with cos2 and became
negligible at the end of the forcing distance. A 3D NWT with (3λ m × 0.4 m × 0.6 m) and a
symmetric axis in y-direction was used, as shown in Figure 2a. A velocity inlet was specified
for the upstream face of the tank. Pressure outlets were located downstream and at the
top of the tank. All the other boundaries applied a no-slip wall boundary condition. The
volumetric mesh generated by the trimmed cell mesher that mostly contains hexahedral
elements is shown in Figure 2a. An overset mesh is used to accommodate the large motions
of the WEC rotor. Two independent meshes are required and superimposed on each other,
and the data can be transferred between the two grids in the interfacial region, as shown
in Figure 2b. To ensure minimal or no loss during this information exchange, the element
size of the background mesh must be similar to that of the overset mesh. Therefore, in
the background mesh, a zone called the overlapping zone is created over a region with an
element size matching that of the overset mesh (Figure 2b).
Energies 2024, 17, 898 5 of 28

Table 1. Physical properties of the model and prototype scale.

Scaling Factor
Description Prototype Model
(k = 11)
Submergence depth, h 3.6 m 1/k 0.3275 m
Beak angle 60◦ 1 60◦
WEC rotor half width, W 2.5 m 1/k 0.2275 m
Total mass 21,328 kg 1/k3 13.65 kg
Inertia about the center of rotation 117,132 kg·m2 1/k5 0.7479 kg·m2
Center of gravity with respect to the center of rotation
xg −0.8934 m 1/k −0.0931 m
zg 1.0189 m 1/k 0.0998 m

Each cell was controlled according to its base cell size (h = 0.5 m). With reference to
the base cell size, the generated volume mesh was divided into very fine, fine, and coarse
regions. The overset and free surface regions were considered very fine, and the mesh was
gradually scaled to fine and coarse, as shown in Figure 2. A prism-layer mesh was adopted
to capture the boundary-layer flow fields, and the thickness of the prism layer was divided
into four layers, yielding y+ < 3. Second-order upwind and unsteady implicit schemes
were used to compute the spatial components in the acceleration term in Equation (2). A
segregated flow solver was used to solve the pressure and each component of the velocity
in an uncoupled manner. The link between the continuity and momentum equations
was established and solved using the semi-implicit method for pressure-linked equations
(SIMPLE) with weak coupling via sublooping. Segregated solvers suffer significantly
because of errors associated with the initial solutions obtained during the iterations. To
overcome these issues, the under-relaxation factors for the velocity and pressure must be
controlled. An automatic convective time-step control method was utilized to maintain a
stable solution with appropriate time steps throughout the simulations. Target mean and
maximum Courant numbers of 0.3 and 0.5, respectively, were employed. In the multiphase
flow, the volume-of-fluid (VOF) method was used to capture the free-surface interface.
Instead of particle motion, the volume of each fluid was tracked using the VOF method and
all fluids were assumed to have the same velocity, pressure, and field. Volume fraction was
introduced into the VOF method. As a result, the governing equations for a fluid mixture
for a single fluid can be developed. The volume fraction of each cell is assumed to contain
only one phase (if α = 1, then the cell is full of water; if α = 0, then the cell is full of air),
or the cell contains an interface between fluids, which is given as 0 < α < 1. The physical
properties of a fluid mixture (ϕ, which can be density (ρ) and viscosity (µ)) are given by

ϕ = αϕwater + (1 − α)ϕair (8)

A common approach was used throughout this study, with α = 0.5 defining the location
of the free-surface interface. A relatively fine mesh was generated near the free-surface
interface region and a high-resolution interface-capturing scheme was used to accurately
trace the interface (Figure 2).
898 6 of 28
Energies 2024, 17, 898 6 of 28
Energies 2024, 17, 898 6 of 28

Figure 1. Schematic of 1.
Figure the
Figure
WEC rotor the
Schematic
1.
in WEC
Schematicofofthe
aWEC
wave tank.
rotor
rotor inaawave
in wavetank.
tank.

(a)
(a)

(b)
Figure
Figure2. Three-dimensional NWT
2. Three-dimensional NWTdetails.
details.(a)(a)
NWT with
NWT withboundary
boundaryconditions,
conditions, (b)
(b) Overset
Overset mesh
details.
mesh details. (b)
3. Implementation
Figure 2. Three-dimensional of PTO (a)
NWT details. Load
NWTTorque
with and Nonlinear
boundary Restoring
conditions, (b) Moment
Overset mesh
details. The equation of motion with a single degree of freedom, where the WEC rotor rotates
about a fixed position, is given by
3. Implementation of PTO Load Torque and Nonlinear Restoring Moment
Energies 2024, 17, 898 7 of 28

3. Implementation of PTO Load Torque and Nonlinear Restoring Moment


The equation of motion with a single degree of freedom, where the WEC rotor rotates
about a fixed position, is given by
.. Z
I ξ = rm × pdS + ρg∀zb − Iwec g − Text , (9)
|wec
{z } | {z } | {z } |{z}
External moment
| {z }
Rate o f change o f Buoyancy/Static moment Inertia moment
Hydrodynamic moment
angular moment o f the WEC
..
where the overhead dot denotes the time derivative and ξ and ξ are the position and
acceleration of the WEC rotor, respectively. rm is the distance vector from the center of
gravity of each face of the WEC rotor. The first two terms on the right side of Equation (9)
are determined by integrating the fluid pressure over the surface area of the WEC rotor.
p is the dynamic pressure acting on the wetted surface, ∀(ξ )zb is the submerged volume
of the WEC rotor, g is the acceleration due to gravity, and the last term is the external
moment considered to be applied with PTO damping (τ pto ) added externally to the system
of the equation of motion. The equation of motion of the WEC rotor with a fully nonlinear
behavior can be estimated by solving the first term on the right-hand side of Equation (9),
using the RANS equation given in Equations (3)–(6). This type of estimation provides an
accurate method for resolving WEC–wave interactions. A linear model can be generated
by linearizing the first term of the equation, which can then be solved in the time domain
to simplify simulations. The linear formulations are as follows (Cummins, [58]):
.. Z t .
( Iwec + I∞ )ξ = Xe − B(t1 )(t − t1 )ξ (t1 )dt1 − ( R(ξ ) + K )ξ (t) − τpto − τvis , (10)
0

where I ∞ is the added mass moment of the WEC inertial matrix, B is the. retardation matrix,
K is the hydrostatic stiffness matrix, t1 is the time-lag variable, and ξ is the WEC rotor
velocity. These linear time-domain formulations are made weakly nonlinear by introducing
the nonlinear restoring moment R(ξ) and the Coulomb-type PTO load applied externally
(τ pto ). The linear viscosity (τ vis ) was obtained using an experimental free-decay test by Kim
et al. [24]. The PTO implementation was conducted using Coulomb damping by providing
a torque in the direction opposite to the motion of the WEC rotor (Figure 3). This damping
mechanism replicated a hydraulic system. Standard Coulomb damping can be applied in
two ways. The first is to apply a constant load torque in the direction opposite to that of the
WEC rotor until the hydrodynamic force on the WEC rotor exceeds the resistive moment
(two-way PTO loading). The second is to apply the load in only one direction and allow
the WEC rotor to move freely without any applied load; this execution is called one-way
loading, as shown in Figure 3a. A standard type can be implemented in a straight manner
in a fully nonlinear model, but it exhibits discontinuous behavior in a weakly nonlinear
model with a shift in direction. To overcome this problem, modified Coulomb damping
can be considered for a weakly nonlinear model, as shown in Figure 3b. Within the closed
.
range of the discontinuous region, a linear variation of the ramp function from −ξ crit to
.
+ξ crit can be applied. This is implemented in the present study using Python code, which
is provided as a user-defined function externally.
The one- and two-way PTO loading systems are selected mainly due to the nonlinear
behavior of the WEC rotor. Figure 3c provides further insight into the two types of PTO
implementation. It shows the pitch response and rotational moment of the WEC rotor,
and demonstrates that the response of positive pitch motion is higher than the negative
response owing to the unique asymmetric geometric configuration of the WEC. In the
one-way PTO loading, during the upward direction of motion of the WEC rotor, which is
the time when the WEC has a positive angular velocity, torque is applied in the opposite
direction, and the WEC rotor is allowed to move freely in the other direction. (Figure 3c).
To examine the efficacy of the implemented PTO loading systems (one- and two-way), a
thorough discussion is provided in Section 5.
Energies 2024, 17, 898 8 of 28
Energies 2024, 17, 898 8 of 28

One-way Two-way

τpto τpto

-τ0 -τ0
Linear variation
.
τ0= 0 τ0= 0 − ξ crit
.
.
+ ξ crit .
ξ ξ
τ0 τ0

(a) (b)

(c)
Figure
Figure3. Implementation
3. Implementation of of
(a)(a)
standard
standardPTOPTOload
loadtorque
torque and
and (b) modifiedPTO
(b) modified PTOload loadtorque.
torque. (c)
Demonstration of PTO load application [left] and the essential components [e.g.,
(c) Demonstration of PTO load application [left] and the essential components [e.g., pitch response, pitch response,
moment, and angular velocity] of the WEC rotor with one-way and two-load
moment, and angular velocity] of the WEC rotor with one-way and two-load PTO load torque [right].
PTO load torque
[right].
It is known that with the linear restoration moment, the area of the secondary water
It isdoes
plane known that with
not change withthe
thelinear
rotationrestoration
angle of themoment,
WEC rotor. theToarea of the the
implement secondary
nonlinearwater
restoring moment, the instantaneous moment resulting from
plane does not change with the rotation angle of the WEC rotor. To implement the the buoyancy and weight of non-
the WEC rotor at each degree of rotation was obtained, as shown
linear restoring moment, the instantaneous moment resulting from the buoyancy and in Figure 4 (top). The
resulting
weight of thenetWEC
moment wasatestimated
rotor each degreeand plotted. Clockwise
of rotation rotation is denoted
was obtained, as shown with
in aFigure
(+) 4
sign, whereas counterclockwise rotation is denoted with a (−) sign, as shown in Figure 4.
(top). The resulting net moment was estimated and plotted. Clockwise rotation is denoted
The estimated linear and nonlinear hydrostatic restoration moments are shown in Figure 4
with a (+) sign,
(bottom). The whereas counterclockwise
slope of the nonlinear restoring rotation
moment ischanged
denotedrapidly,
with awhereas
(−) sign,theasslope
shown in
Figure 4. The estimated linear and nonlinear hydrostatic restoration
of the linear moment remained constant within the range of the rotation angle of the WEC moments are shown
in Figure
rotor. The 4 (bottom).
effectivenessTheofslope of the nonlinear
the nonlinear hydrostaticrestoring
moment in moment
the weaklychanged rapidly,
nonlinear
whereas the slope
formulations wasofvalidated
the linear moment
through remainedand
experimental constant within
numerical the range of the rotation
comparisons.
angle of the WEC rotor. The effectiveness of the nonlinear hydrostatic moment in the
weakly nonlinear formulations was validated through experimental and numerical com-
parisons.
Energies 2024, 17, 898 9 of
Energies 2024, 17, 898 9 of 28

Linear Nonlinear
1200
(+)P4
Restoring moment [kNm]

800 (+)P5
(+)P6
400
(+)P3
P0
0 (+)P2
(-)P5 (+)P1
(-)P4 (-)P1
-400
(-)P3 (-)P2
-800

-150 -120 -90 -60 -30 0 30 60 90 120 150 180


WEC-rotor rotation angle [deg]

Forces and
Figure4.4.Forces
Figure andrestoring moments
restoring momentsacting on the
acting onWEC
the rotor
WECatrotor
instantaneous positions (Red
at instantaneous positions (R
Arrow—Weight of the WEC rotor; Green Arrow—Buoyancy of the WEC
Arrow—Weight of the WEC rotor; Green Arrow—Buoyancy of the WEC rotor)—Top rotor)—Top and Computa-and Comp
tion of
tation ofLinear
Linearand
andNonlinear Restoring
Nonlinear Moments
Restoring at Various
Moments Angles ofAngles
at Various the WECofRotor—Bottom.
the WEC Rotor—Bottom
4. Verification of Numerical Models
4. 4.1.
Verification
Convergenceof Numerical Models
Test
NWT was employed
4.1. Convergence Test to simulate the wave–WEC rotor interaction, with the tank
extending from 1.5 × λ in the positive and negative x-direction. In addition, y ranged from
zeroNWT
to twicewas theemployed
half widthto of simulate the wave–WEC
the WEC rotor, rotor interaction,
and z was maintained equal to thewith the tank e
water
tending
depth (d) from
(2 m1.5
for×the λ in the scale
model positive
and 40and m negative x-direction.
for the prototype). Insimulations,
In the addition, yonly ranged fro
zero to twice the half width of the WEC rotor, and z was maintained equal to the wa
half of the section in the y-direction was considered because of the symmetry of the x–z
plane.(d)
depth Figure
(2 m3 fordepicts
the the
modelboundary
scale conditions
and 40 m required for the NWT. In
for the prototype). Thethe
uncertainty
simulations, on
of the numerical computation was examined in terms of the mesh size (spatial), time step
half of the section in the y-direction was considered because of the symmetry of the x
(temporal) and domain size by calculating the absorbed average power (Pabs ) to verify the
plane. Figure 3 depicts the boundary conditions required for the NWT. The uncertain
convergence of the numerical CFD results. Pabs is the angular velocity multiplied by the
ofPTO
the numerical
load torque.computation
The mean of the was examined
steady-state in terms
cycles of the
was used to mesh
quantifysize (spatial),
average Pabs . time st
(temporal)
Testing wasand domain
performed bysize by calculating
generating a regular wavethe absorbed
interaction average
with a WEC power
rotor (P abs) to verify t
at model
scale (T = 1.43 s; H = 0.136 m corresponding to a wave steepness
convergence of the numerical CFD results. Pabs is the angular velocity multiplied of 0.04; one-way PTO by t
load torque = 5.56 Nm) and prototype scale (T = 4.75 s; H = 2.0 m
PTO load torque. The mean of the steady-state cycles was used to quantify average P corresponding to a wave
steepness of 0.06; one-way PTO load torque = 70 kNm). The computations for both the
Testing was performed by generating a regular wave interaction with a WEC rotor
weakly and fully nonlinear models were carried out using Intel® Xenon® Gold processors.
model scale nonlinear
The weakly (T = 1.43 model
s; H =had
0.136 m corresponding
a 16-core processor, whereasto a wave steepness
the fully nonlinearofmodel
0.04; one-w
PTO
hadload torque
a 44-core = 5.56 Nm)
processor, with and prototype
1- and scale (T = dedicated
10-core processors 4.75 s; H =to2.0themsimulations,
corresponding to
wave steepness
respectively. For of 0.06; one-way
a simulation time ofPTO load
1 s, the CPUtorque = 70was
clock time kNm). The computations
approximately 5 s for the for bo
weakly nonlinear model, and 0.8 h for the fully nonlinear model.
the weakly and fully nonlinear models were carried out using Intel Xenon® Gold proce ®
The instantaneous
sors. The weakly nonlinear Pabs computed
model had with athe fully nonlinear
16-core processor, solution was numerically
whereas the fully nonline
verified based on the mesh size, time step, and domain size. The numerical uncertainty
model had a 44-core processor, with 1- and 10-core processors dedicated to the simu
tions, respectively. For a simulation time of 1 s, the CPU clock time was approximately
s for the weakly nonlinear model, and 0.8 h for the fully nonlinear model.
The instantaneous Pabs computed with the fully nonlinear solution was numerica
verified based on the mesh size, time step, and domain size. The numerical uncertainty
Energies 2024, 17, 898 10 of 28

of Pabs (εnum ) is the sum of the variations due to the mesh size (εm ), time step (εt ), and
domain size (εCm ). The mesh size uncertainty was evaluated using ∆zi (i = 1, . . ., 4), with
the number of cells within the wave height considered as 4, 8, 16, and 20. The ∆x/∆y
and ∆x/∆z values were maintained at 4. When computing the uncertainty of the mesh
size ratio, the reference mesh size was considered as ∆zref = H/8. Additionally, other
parameters such as the reference time step of ∆tre f = T/2000, reference domain size of Cmref
(length, L = (0.5λ, 2λ), where the first value represents wave forcing on both sides of the
computational domain, and the second value corresponds to the actual computational zone;
breadth, B = 2W; height, Ht = dm) are taken into account. A fine mesh was maintained
within the overset region accommodating the WEC rotor. Moreover, a gradual increase
in mesh size was implemented, commencing from a fine mesh that was 10% denser and
progressing to a coarse mesh that was 15% denser than the fine mesh. This resulted in a
typical total mesh size of 1.3 million cells in the background mesh and 137,430 cells in the
overset mesh regions, corresponding to half of the computational domain with a single
WEC rotor. Time size uncertainty was investigated with different time steps (∆ti , i = 1,. . .,4)
of T/1000, T/1500, T/2000, and T/2500 with the reference mesh size (∆zre f = H/8) and
reference domain size (L = (0.5λ, 2λ); B = 2W; Ht = dm). Finally, the uncertainty in the
computational domain size was examined by considering three different sizes. The first
size was based on Cm1 (L = (0.25λ, 1λ); B = 2W; Ht = dm), the second size was based
on Cm2 (L = (0.5λ, 2λ); B = 2W; Ht = dm), and the third size was based on Cm3 ((1.0λ,
4λ); B = 4W; Ht = dm). Other reference parameters were the same as those described
previously. In the simulations, the maximum Courant number was less than 0.5. The results
are presented in a normalized form as ∆zre f /∆zi for the mesh size, and for the domain size,
length parameter is considered and presented in Cmi/ Cmref and ∆ti /∆tre f for the time step.
A quadratic curve was fitted, as given by the equations, and the total standard deviation
of the fitted curve was obtained because σfit is used to extrapolate Pabs for an infinitely
 
fine grid/small time step/small domain size Pabs,i /Pabs,re f , as shown in Figure 5. The
0
uncertainty 
of the mesh size,
 time step, and
 domain size can be represented by the equation
(δ = 100 × 1.25 × 1 − Pabs,i /Pabs,re f + σFit ), as suggested by Cummins et al. [59].
0
The uncertainty values obtained for the model scale were 0.09% (mesh size), 0.13% (time
step) and 1.64% (domain size). At the prototype scale, the uncertainty values were 0.30%
(mesh size) and 3.1% (domain size). Hence, the total uncertainty εnum is 0.22% and 3.53%
for the model and prototype scales, respectively. The convergence study revealed that
an increase in the wave steepness resulted in an increase in the percentage of uncertainty.
However, the overall uncertainty remained below 3.5%. A comparative study with the
experimental data was performed and is presented in the following section to validate the
numerical results.

4.2. Numerical Validation


The validation experiments were conducted in a wave basin (28 m × 22 m × 2.5 m) at
the Research Institute of Medium and Small Shipbuilding in South Korea. A piston-type
wavemaker was operated at one end of the tank to generate monochromatic directional
waves. A porous plate was placed at the other end of the tank to absorb the reflected
waves. A PTO load torque was applied using a hysteresis braking system to evaluate the
power of the WEC rotor. Initially, tests were conducted to determine the controllability,
performance, and repeatability before integrating with the WEC rotor. This process was
accomplished by placing weights and measuring the torque, which was then compared
with the performance curve provided by the manufacturer. This procedure was repeated
several times. Once the calibration was verified, the braking mechanism and WEC rotor
were used for testing.
17, 898 Energies 2024, 17, 898 11 of 28
11 of 28

(Pabs,i /Pabs,ref)0 = 0.9991; coeff1 = 0.001016 (Pabs,i /Pabs,ref)0 = 0.9987; coeff2 = 0.001437
1.004
1.004
Pabs,i /Pabs,ref

Pabs,i /Pabs,ref
1.002
1.002
1 1
0.998
0 0.5 1 1.5 2 0 0.5 1 1.5 2
zref / zi ti / tref
(P ) = ( Pabs , i Pabs , ref ) + coeff1 × (  zref  zi ) (P Pabs , ref ) = ( Pabs , i Pabs , ref ) + coeff 2 × ( ti tref )
2 2
abs , i Pabs , ref abs , i
Fit 0 Fit 0

σ Fit = 1.1876 × 10−7 σ Fit = 1.19 × 10−7


(a) (b)

(Pabs,i /Pabs,ref)0 = 0.9836; coeff3 = 0.01737


1.06
Pabs,i /Pabs,ref

1.04
1.02
1
0.98
0 0.5 1 1.5 2
Cmi /Cmref

(P ) = ( Pabs , i Pabs , ref ) + coeff 3 × ( Cmi Cmref )


2
abs , i Pabs , ref
Fit 0

σ Fit = 2.504 ×10−5


(c)

(Pabs,i /Pabs,ref)0 = 0.997; coeff4 = 0.002972 (Pabs,i /Pabs,ref)0 = 0.9714; coeff5 = 0.03086
1.012 1.12
Pabs,i /Pabs,ref

1.08
Pabs,i /Pabs,ref

1.008
1.004 1.04
1 1
0.996 0.96
0 0.5 1 1.5 2 0 0.5 1 1.5 2
zref / zi Cmi /Cmref

(P ) = ( Pabs , i Pabs , ref ) + coeff 4 × (  zref  zi ) (P Pabs , ref ) = ( Pabs , i Pabs , ref ) + coeff5 × ( Cmi Cmref )
2 2
abs , i Pabs , ref abs , i
Fit 0 Fit 0

σ Fit = 6.644 ×10−7 σ Fit =1.485 ×10−4


(d) (e)

Quadratic fit Extrapolated (Minimum cell size ratio)


Fully nonlinear (CFD) Extrapolated (Minimum time step ratio)
Extrapolated (Minimum computational domain size ratio)

Figure 5. Numerical
Figureconvergence
5. Numericaluncertainty
convergence obtained
uncertaintyby the fully
obtained nonlinear
by the solution
fully nonlinear for model
solution for model
scale ((a)—mesh size, (b)—time step, and (c)—domain size) and for prototype scale ((d)—mesh size size
scale ((a)—mesh size, (b)—time step, and (c)—domain size) and for prototype scale ((d)—mesh
and (e)—domainandsize).
(e)—domain size).

Simulations based on the weakly and fully nonlinear model results were compared
4.2. Numerical Validation
with the experimental results by resolving the absorbed power in terms of the pitch response
The validation experiments
and angular velocity.were conducted
The wave in were
conditions a wave basin (28
maintained at Tm= ×1.58
22 smand
× 2.5 m) at
a wave height
the Research Institute
H = 0.136ofm.Medium and
Figure 6a,b Small
show Shipbuilding
the time histories of in
theSouth
angularKorea.
velocityAand
piston-type
pitch response
wavemaker wasin operated
the absenceatofone
the PTO
end load torque.
of the tankThe repeatability
to generate of the profiles and
monochromatic magnitudes of
directional
the time series for up to 20 s were examined for comparison.
waves. A porous plate was placed at the other end of the tank to absorb the reflectedThe integration of the viscous
damping into the weakly nonlinear model was approximated using an experimental free-
waves. A PTO load torque was applied using a hysteresis braking system to evaluate the
decay test performed by Kim et al. [24]. A damping ratio of 0.0849 and a corresponding
power of the WEC rotor.
viscous Initially,
damping tests
value werekg.m
of 0.4193 conducted
2 /s wereto determine
added theThe
externally. controllability,
initial transit was
performance, and repeatability before integrating with the WEC rotor. This process was
accomplished by placing weights and measuring the torque, which was then compared
with the performance curve provided by the manufacturer. This procedure was repeated
several times. Once the calibration was verified, the braking mechanism and WEC rotor
were used for testing.
wave height H = 0.136 m. Figure 6a,b show the time histories of the angular velocity and
pitch response in the absence of the PTO load torque. The repeatability of the profiles and
magnitudes of the time series for up to 20 s were examined for comparison. The integra-
tion of the viscous damping into the weakly nonlinear model was approximated using an
Energies 2024, 17, 898 experimental free-decay test performed by Kim et al. [24]. A damping ratio of 0.0849 12 of 28 and
a corresponding viscous damping value of 0.4193 kg.m2/s were added externally. The in-
itial transit was ignored because the initial ramp times for the weakly and fully nonlinear
ignored
models because
differed,the initial
and wereramp times forinthe
not present theweakly and fullyresults.
experimental nonlinear models minimal
Relatively differed, dif-
and werewere
ferences not present
observed in between
the experimental results.
the positive and Relatively minimal
negative peaks, differences
indicating thatwere
the over-
observed between the positive and negative peaks, indicating that the overall
all comparison of the weakly and fully nonlinear solutions and the experimental results comparison
of
wasthegood.
weakly and fully
A percent nonlinear
variation of solutions
1.54 and 5.6andhasthebeen
experimental
observedresults
in the wasfullygood. A
and weekly
percent variation of 1.54 and 5.6 has been observed in the fully and weekly nonlinear
nonlinear models, respectively, when compared to the experimental results. Figure 6 also
models, respectively, when compared to the experimental results. Figure 6 also indicates
indicates that the WEC response trend is nonlinear relative to the incoming wave, and
that the WEC response trend is nonlinear relative to the incoming wave, and that the
that the weakly nonlinear model is capable of effective prediction. The experimental and
weakly nonlinear model is capable of effective prediction. The experimental and fully
fully nonlinear
nonlinear model model
results results
were inwere in agreement.
agreement. The investigation
The investigation was extended was extended
to a wide to a
wide range of wave periods from 0.8 to 2 s with a wave height of
range of wave periods from 0.8 to 2 s with a wave height of 0.03, and the results 0.03, and the results
werewere
compared (Figure 7). Thus, wave conditions encompassing a range
compared (Figure 7). Thus, wave conditions encompassing a range of low to moderately of low to moderately
steepwaves
steep waveswere
wereanalyzed
analyzed (see
(see Table
Table 2).2). Except
Except forfor
thethe region
region close
close to resonance
to the the resonancepeakpeak
pitchRAO,
pitch RAO,thetheoverall
overallcomparison
comparison between
between thethe experimental
experimental andand numerical
numerical results
results was was
satisfactory. Moreover, when wave periods exceeding 1.5 s were taken into account, the the
satisfactory. Moreover, when wave periods exceeding 1.5 s were taken into account,
CFD
CFD results
resultsshowed
showeda apartial
partialdeviation
deviation fromfromthethe
experimental
experimental data, unlike
data, the weakly
unlike the weakly
nonlinear
nonlinearmodel.
model.These
Thesedifferences
differences can bebe
can attributed, at least
attributed, partially,
at least to reflections
partially, fromfrom
to reflections
sidewalls that occur during long-wave
sidewalls that occur during long-wave periods.periods.

10
(a)
Angular velocity
[rad/s]

-10
0 2 4 6 8 10 12 14 16 18 20
Time (s)

100
(b)
Pitch response
[deg]

-100
0 2 4 6 8 10 12 14 16 18 20
Time (s)
Experiment Fully nonlinear (CFD) Weakly nonlinear (Potential)

Energies 2024, 17, 898 Figure6.6.Time


Figure Timehistory
historyofofthe (a)(a)
the angular
angularvelocity andand
velocity (b) (b)
pitch response
pitch of the
response of WEC rotorrotor
the WEC without
without
13 of 28
PTO load torque for wave parameters of H = 0.13 m and T = 1.58 s (Red dotted line—Experiment,
PTO load torque for wave parameters of H = 0.13 m and T = 1.58 s (Red dotted line—Experiment,
Solidline—Fully
Solid line—Fullynonlinear
nonlinear (CFD),
(CFD), Blue
Blue dotted
dotted line—Weakly
line—Weakly nonlinear
nonlinear (Potential)).
(Potential)).

Table Fully nonlinear


2. Wave (CFD) details for
steepness Weakly
the nonlinear (Potential)
results shown Experiment
in Figure 7.

S.No 60Wave Period (T) Wave Length (λ) Height (H) Wave Steepness (H/λ)
1 0.8 1.148 0.026
Pitch RAO [rad/m]

2 1.00 1.668 0.0180


40
3 1.28 2.627 0.0114
4 1.43 3.246 0.0092
5 20 1.58 3.931 0.03 0.0076
6 1.73 4.670 0.0064
7 1.88 5.445 0.0055
8 0 2.04 6.294 0.0048
0 1 2 3
Wave period (s)

Figure
Figure 7.
7. Comparison
Comparison and variation of
and variation of pitch
pitchRAO
RAOas
asaafunction
functionofofwave
waveperiod
periodtorque
torquefor
forHH= =0.03
0.03m.m.

The validation test was extended to examine the implementation of the PTO loading
system with the experimental generation; the incoming wave parameters were the same
as those in Figure 6. Typical time series are shown and compared in Figure 8. An assess-
ment was conducted by confirming the profiles and trends based on the expected time-
series behavior. When the applied load was zero during the negative angular velocity
40

Pitch RAO [rad/


Energies 2024, 17, 898 20 13 of 28

Table 2. Wave steepness


0 details for the results shown in Figure 7.

S.No Wave Period (T) Wave Length (λ) Height (H) Wave Steepness (H/λ)
0 1 2 3
1 0.8 Wave1.148
period (s)
0.026
2 1.00 1.668 0.0180
Figure3 7. Comparison
1.28 and variation2.627
of pitch RAO as a function of wave period
0.0114 torque for H = 0.
4 1.43 3.246 0.0092
5 1.58 3.931 0.03 0.0076
The
6 validation
1.73 test was extended
4.670 to examine the implementation
0.0064 of the PTO l
system7 with the1.88experimental 5.445
generation; the incoming wave 0.0055 parameters were th
8 2.04 6.294 0.0048
as those in Figure 6. Typical time series are shown and compared in Figure 8. An
ment was conducted by confirming the profiles and trends based on the expecte
The validation test was extended to examine the implementation of the PTO loading
series behavior. When the applied load was zero during the negative angular v
system with the experimental generation; the incoming wave parameters were the same as
phase,
those in the weakly
Figure and
6. Typical fully
time seriesnonlinear numerical
are shown and comparedmodelsin Figureimplemented
8. An assessment an appr
zero load. However,
was conducted the experiment
by confirming the profiles revealed
and trendsthat
based slightly higher values
on the expected were main
time-series
behavior. When the applied load was zero during the negative angular
throughout the negative-phase cycles. When the sign of the load torque changes, th velocity phase,
thethe
of weakly and fully nonlinear
experimental numerical
load torque models linearly
changes implemented andanreaches
appropriate
the zero
inputload.
load tor
However, the experiment revealed that slightly higher values were maintained throughout
indicated by the numerical
the negative-phase cycles. When simulation
the sign ofresults.
the loadSome
torquekinks occurred
changes, the slope when
of themeasure
ing the direction
experimental change,
load torque andlinearly
changes these kinks werethe
and reaches not found
input load in the weakly
torque, andbyfully no
as indicated
numerical
the numericalmodels;
simulationsubsequently, theyoccurred
results. Some kinks disappeared (Figureduring
when measured 8, left).
theFigure 8 (right)
direction
change, and these kinks were not found in the weakly and fully nonlinear
the two-way PTO load torque; the profile and trend are similar to those of the on numerical models;
subsequently, they disappeared (Figure 8, left). Figure 8 (right) depicts the two-way PTO
PTO load torque but with a smooth transition during the direction shift. Overall, th
load torque; the profile and trend are similar to those of the one-way PTO load torque but
load
with atorque
smoothtrend matched
transition duringwell with the
the direction weakly
shift. Overall,nonlinear,
the PTO load fully nonlinear,
torque trend and
mental
matchedresults.
well with the weakly nonlinear, fully nonlinear, and experimental results.

4 6
2 4
Load torque [Nm]

Load torque [Nm]

0
2
-2
0
-4
-2
-6
-8 -4
-10 -6

17 17.5 18 18.5 17 17.5 18 18.5


Time (s) Time (s)

(a) (b)
Figure
Figure 8.8.Comparison
Comparison of instantaneous
of instantaneous (a)
(a) one-way (5.3one-way (5.3
Nm) and (b) Nm) (2
two-way and
Nm)(b)
PTOtwo-way (2 Nm) PT
load torques
torques (Black line—Experiment,
(Black line—Experiment, Red line—Fully
Red line—Fully nonlinear (CFD), Blue nonlinear
line—Weakly(CFD),
nonlinearBlue line—Weakly no
(Potential)).
(Potential)).
After proper verification and validation of the numerical models, the remainder of this
paper discusses the prototype WEC rotor (Table 1). Most numerical settings were similar to
thoseAfter properscale,
of the model verification and validation
with the appropriate of the
parameters numerical
scaled models,
remai
up to suit the the
prototype.
this paper discusses the prototype WEC rotor (Table 1). Most numerical setting
5. Results
similar toand Discussion
those of the model scale, with the appropriate parameters scaled up to s
5.1. Effect of PTO Load Torque
prototype.
The WEC rotor average Pabs for each PTO load torque was examined to determine
the optimal average Pabs . The numerical testing included one-way τpto and τpto two-way
systems for wide ranges of 10–140 kNm and 5–50 kNm, respectively, which are shown in
optimal average Pabs. The numerical testing included one-way τ pto and τ pto two-way
systems for wide ranges of 10–140 kNm and 5–50 kNm, respectively, which are shown in
Figure 9 as the absorbed power and efficiency (Pabs/Pw), respectively. The energy flux or
Energies 2024, 17, 898 wave power (Pw) of a regular wave is expressed as follows:
14 of 28

Pw = ρ gH b
2
kW,
(16π × ω ) (11)
Figure 9 as the absorbed power and efficiency (Pabs /Pw ), respectively. The energy flux or
where
waveωpower is the (Pwave
w ) offrequency
a regular wave and b is expressed
the width of as the crest, which is equal to the width
follows:
of the WEC rotor. The wave steepness (H/λ) was maintained constant at 0.03 throughout
2
the simulations, and the water depth Pw = was b/(16πat× 40
ρgHfixed ω )kW,
m. The average Pabs increased and (11)
reached its maximum value with an increase in the PTO load, and the average Pabs de-
where with
creased ω is the wave frequency
an further increase and in theb is the torque,
load width ofwhich the crest, which
is true is equal
for both types to the width of
of loading
the WEC rotor. The wave steepness (H/λ) was maintained
systems. Under fixed wave conditions, the range of the applied load torque was greater constant at 0.03 throughout the
simulations, and the water depth was fixed at 40 m. The average Pabs increased and reached
foritsthe one-wayvalue
maximum
τ ptowith
system than forinthe
an increase thetwo-way
PTO load, and
τ pto the
system. As shown
average Pabs decreasedby the Pwith abs

behavior,
an further although
increase theinpower
the load in the one-way
torque, whichsystem wasfor
is true susceptible
both types to the torque applied,
of loading systems.
it Under
was relatively
fixed wave strong in the two-way
conditions, the range system.
of the A applied
second-order polynomial
load torque was greaterwas fitted for tothe
one-way
a pair of points τ system than for the two-way
pto to observe the trend in Pabs with the τ pto applied load torque ( τ
system. As shown by the P abs behavior,
pto and average
although the power in the one-way system was susceptible to the torque applied, it was
Pabs ). The strong
relatively goodness in theoftwo-way
fit wassystem.determined by measuring
A second-order the Rwas
polynomial
2 value given by
fitted to a pair of
(
1−points
Pabs,resto observe )
Pabs,tot , the
andtrend
the Rinvalues
2 Pabs witharethe applied
shown load torque
in Figure (τpto and
9 (where is thePsum
average
Pa bs , res abs ). The
of
goodness of fit was determined by measuring the R2 value given by 1 − ( Pabs,res /Pabs,tot ),
the squares 2of the distances of Pabs from the fitted curve and Pa b s , to t is the sum of the
and the R values are shown in Figure 9 (where Pabs,res is the sum of the squares of the
squares
distances of theof Pdistances
abs from the from thecurve
fitted meanand of P Pabs ). Figure
abs,tot is the9sumalsoofdepicts the resulting
the squares quad-
of the distances
from the mean of P ). Figure 9 also depicts the resulting
ratic fit equations andabsdemonstrates that the average Pabs for the one-way and two-way quadratic fit equations and
demonstrates that the average
systems are of the second order (quadratic) P abs for the one-way and two-way systems
in terms of their behavior with the PTO load are of the second
order (quadratic)
torque. The curve fitting in terms forofefficiency
their behavior with the
is omitted PTO load
because torque.isThe
efficiency curve derived
directly fitting for
efficiency
from the Pabsis, where
omittedthe because
curve fit efficiency
remainsissamedirectly both. from the Pabs , where the curve fit
for derived
remains same for both.
Fully nonlinear (CFD) Weakly nonlinear (Potential)
Quadratic fit (CFD) Quadratic fit (Potential)
20 100

15 80
Efficiency (%)
Average Pabs

60
[kW]

10
40
5
20
0 0

0 30 60 90 120 150 0 30 60 90 120 150


pto (kNm) pto (kNm)

Energies 2024, 17, 898 Quadratic fit Average Pabs = − 0.0042 ×τ pto
2
+ 0.6232 ×τ pto − 4.3841; R 2 = 0.9617 (CFD) 15 of 28

equations Average Pabs = − 0.0043 ×τ pto + 0.5643 ×τ pto − 3.3265; R = 0.8579 (Potential)
2 2

(a)
20 100

15 80
Efficiency (%)
Average Pabs

60
[kW]

10
40
5
20
0 0

0 30 60 0 30 60
pto (kNm) (kNm)
pto

Average Pabs = − 0.0152 ×τ pto


2
+ 0.9655 ×τ pto − 4.0634; R 2 = 0.9866 (CFD)
Quadratic fit 

equations Average Pabs = Pabs = − 0.0149 ×τ pto2
+ 0.8663 ×τ pto − 2.4504; R 2 = 0.998(Potential)

(b)
Figure
Figure9.9.Variation of of
Variation average PabsPand
average efficiency
abs and withwith
efficiency applied PTOPTO
applied loadload
torque. (a) One-way,
torque. (b)
(a) One-way,
two-way.
(b) two-way.

Based on the theory discussed by Falnes and Kurniawan [13], the quadratic behavior
can be justified. For small motion amplitudes of the WEC rotor, the excitation power (Pext)
varies linearly with the angular velocity, and the WEC rotor radiates waves, where the
associated power is the radiated power (Prad), which varies quadratically with the angular
Based on the theory discussed by Falnes and Kurniawan [13], the qu
can be justified. For small motion amplitudes of the WEC rotor, the excita
Energies 2024, 17, 898 varies linearly with the angular velocity, and the WEC rotor15radiates of 28 w
associated power is the radiated power (Prad), which varies quadratically
velocity (Figure 10). The difference between the two powers (Pext and Pr
Based on the theory discussed by Falnes and Kurniawan [13], the quadratic behavior
be be
can represented by amotion
justified. For small quadratic parabola,
amplitudes of the WEC asrotor,
shown in Figure
the excitation power10.
(PextP)abs can
taking
varies the product
linearly of thevelocity,
with the angular angular andvelocity and radiates
the WEC rotor applied PTOwhere
waves, loadthetorque.
associated power is the radiated power (Prad ), which varies quadratically with the angular
load torque must also be quadratic because the angular velocity has be
velocity (Figure 10). The difference between the two powers (Pext and Prad ) is that Pabs can
to represented
be be quadratic. The overall
by a quadratic parabola,agreement between
as shown in Figure 10. Pabsthe
canweakly and
be calculated byfully n
taking the product
was good, of thethat
except angular
thevelocity
weakly and applied PTO load
nonlinear torque. The applied
formulations PTO whe
deviated
load torque must also be quadratic because the angular velocity has been demonstrated
was
to more than
be quadratic. 50 kNm.
The overall These
agreement distinctions
between the weaklyare
and discussed
fully nonlinearinmodels
detail in S
one-
was andexcept
good, two-way
that the PTO
weaklyload torque
nonlinear systems,
formulations the maximum
deviated average Pabs
when the τ pto loading
was more than 50 kNm. These distinctions are discussed in detail in Section 5.4. In the one-
to be 70 and 30 kNm, respectively, using the two numerical models. F
and two-way PTO load torque systems, the maximum average Pabs was determined to be
weakly
70 nonlinear
and 30 kNm, models,
respectively, thetwo
using the equivalent maximum
numerical models. efficiencies
For the fully and weaklywere 8
spectively,
nonlinear with
models, the one-way
the equivalent maximumsystem, andwere
efficiencies these
82%values
and 77%,were 48% and 4
respectively,
with the one-way system, and these values were 48% and 44%, respectively, when the
when the two-way system was used.
two-way system was used.

1
Pext = Fext ,5 ω5 cos(γ 5 )
2
Pmax Prad Quadratic dependent
curve

Pabs

0 ωopt ,5 ω5
Figure 10. Quadratic variation of absorbed Pabs with angular velocity (ω), where the subscript 5
Figure 10.
represents the Quadratic variation
pitch mode (Falnes of absorbed
and Kurniawan [13]). Pabs with angular velocity (ω), where t
resents the pitch mode (Falnes and Kurniawan [13]).
5.2. Effect of the Wave Period
The influence of the wave period on pitch RAO and average Pabs was investigated
5.2.solving
by Effecttheoftwo
thenonlinear
Wave Period
numerical models (weakly and fully) at τ pto = 20 kNm for
various wave periods ranging from 4.25 to 6.25 s and by using one- and two-way loading
The
systems. Theinfluence
wave heightofwas
thekept
wave period
constant onm.pitch
at 1.0 RAO
Within and of
the range the wavePabs wa
average
solving
periods the two nonlinear
investigated, the pitch RAOnumerical
increased withmodels (weakly
the wave andit fully)
period until reachedat τpto = 20
the
maximum value and then decreased for the two types of loading systems
wave periods ranging from 4.25 to 6.25 s and by using one- and two-way (Figure 11, left). A
comparison of the weakly and fully nonlinear models revealed a difference in the resonance
The wave
region height
for one- and was
two-way kept with
loadings, constant at 1.0 m.
a higher magnitude Within
with the
the weakly range of th
nonlinear
model. Figure 11 (right) shows the matching average Pabs . The two models exhibited good
agreement with each other for all wave periods in both types of loading systems. When a
fixed τ pto value was used, the average Pabs was higher in the two-way loading system than
in the one-way loading system.
gion for one- and two-way loadings, with a higher magnitude with the weakly nonlin
model. Figure 11 (right) shows the matching average Pabs. The two models exhibited go
agreement with each other for all wave periods in both types of loading systems. Whe
fixed τpto value was used, the average Pabs was higher in the two-way loading system t
Energies 2024, 17, 898 16 of 28
in the one-way loading system.

Fully nonlinear (CFD) Weakly nonlinear (Potential)


3 16

12

Average Pabs
2

Pitch RAO
[rad/m]

[kW]
8
1
4

0 0

4 5 6 7 4 5 6 7
Wave period (s) Wave period (s)
(a)
3 16

12

Average Pabs
2
Pitch RAO
[rad/m]

[kW]
8
1
4

0 0

4 5 6 7 4 5 6 7
Wave period (s) Wave period (s)
(b)
Figure 11.Variation
Figure 11. Variationof of pitch
pitch RAO RAO and average
and average Pabswave
Pabs with withperiod
wavefor
period
a fixedfor
PTOa fixed PTO load
load torque of torqu
20
20kNm. (a)One-way
kNm. (a) One-way τpto
τ pto = 20
= 20 kNm,
kNm, (b) Two-way
(b) Two-way τ pto = τ20 = 20 kNm.
ptokNm.

5.3. Effect of the Number of WEC Rotors


5.3. Effect of the Number of WEC Rotors
The developed weakly nonlinear formulas were extended to multibody WEC rotors in
The
regular developed
waves, along withweakly
a fullynonlinear formulas
nonlinear analysis. Thewere extended to multibody
multiple-WEC-rotor response wasWEC rot
in regularbywaves,
captured along the
implementing with a fullyoverset
multiple nonlinear analysis.The
grid technique. Theanalysis
multiple-WEC-rotor
was performed respo
was captured by implementing the multiple overset grid technique. The12).
with three WEC rotors, and the influence of rotor spacing was included (Figure The was p
analysis
wave conditions were set as follows: H/λ = 0.03, T
formed with three WEC rotors, and the influence of rotor spacing = 4.75 s, and τ pto = 50 kNm using
was included a (Fig
one-way PTO load torque mechanism. Table 3 shows the pitch RAO, average Pabs , and
12). The wave conditions were set as follows: H/λ = 0.03, T = 4.75 s, and τpto = 50 kNm us
efficiency of multiple WEC rotors with different spacings (0.8 × W, 1.0 × W, and 1.2 × W)
aalong
one-way PTOofload
with that torque
a single WECmechanism. Table 3The
rotor for comparison. shows
effectthe pitch
of the RAO,ofaverage
number rotors Pabs, a
efficiency
with spacing ofwas
multiple
evaluatedWECusingrotors with different
the percentage spacings
ratio (PR) (0.8individual
of multiple × W, 1.0 rotors
× W, and
to 1.2 ×
along
a singlewith
WECthat ofasa listed
rotor, singleinWEC
Table 3.rotor for comparison.
In accordance The nonlinear
with the fully effect of simulations
the number of rot
for varying
with spacing spaces, the pitch RAO
was evaluated usingandthe
average Pabs of the
percentage PR (PR)
ratio decreased by 3%, and
of multiple the
individual rot
weakly nonlinear model for the side rotors retained a similar trend (Rotors 1 and 3) with
to a single WEC rotor, as listed in Table 3. In accordance with the fully nonlinear simu
the increase in spacing. In the case of the center rotor, the pitch RAO and absorbed Pabs of
tions for varying spaces, the pitch RAO and average Pabs of the PR decreased by 3%, a
the PR increased by 3% with a fully nonlinear model; however, the trend reversed when a
the weakly
weakly nonlinear
nonlinear model model
was used. forOverall,
the side therotors retained
pitch RAO a similar
and average Pabstrend
of the (Rotors
center 1 and
with
rotor the increase
PR and those ofin the
spacing. In the
side rotors werecase of the
always ≤5,center rotor,
regardless thenumerical
of the pitch RAO and absor
model
Pused.
abs of In Figure
the 13, a clear velocity
PR increased by 3% withflow pattern
a fullyisnonlinear
observed around
model; thehowever,
center rotor,
thewhich
trend rever
differsafrom
when the side
weakly rotors owing
nonlinear model to was
the cross-flow across the
used. Overall, the cylindrical
pitch RAO hole.
andThis flow Pabs of
average
pattern is characterized by high velocities and low wave elevations at the hole, resulting
center rotor PR and those of the side rotors were always ≤5, regardless of the numer
in a slamming phenomenon that is more prominent at the center rotor than at the side
rotors. Hence, the center rotor has a lower pitch RAO and lower average Pabs values than
the side rotors.
which
model used. In Figure 13, adiffers from the
clear velocity flowside rotors
pattern owing toaround
is observed the cross-flow
the centeracross
rotor, the cylind
which differs from the side rotors owing to the cross-flow across the cylindrical
flow pattern is characterized by high velocities and low wave hole. Thiselevations a
flow pattern is characterized
sulting in by high velocities
a slamming and low wave
phenomenon elevations
that is at the hole,
more prominent at re-
the center ro
sulting in a slamming phenomenon that is more prominent at the center rotor than
side rotors. Hence, the center rotor has a lower pitch RAO and 17lower at the aver
Energies 2024, 17, 898 of 28
side rotors. Hence, than
the center rotor has
the side rotors.a lower pitch RAO and lower average P abs values

than the side rotors.

Figure 12. Generated mesh around the multiple WEC rotors.


Figure
Figure 12.12. Generated
Generated meshthe
mesh around around
multiplethe
WECmultiple
rotors. WEC rotors.

Figure 13. Wave interaction


Figure 13.with
Wavethe multiple
interaction WEC
with rotors for
the multiple a spacing
WEC rotors for of 0.8 × W
a spacing 0.8 × W showing
of showing the free
the free
surface elevation andsurface elevation
velocity and velocity
magnitude magnitude
during duringand
the trough the crest
troughof
and crest
the waveof the wave period.
period.

Figure 13. Wave interaction with the multiple WEC rotors for a spacing of 0.8 × W s
surface elevation and velocity magnitude during the trough and crest of the wave p
Energies 2024, 17, 898 18 of 28

Table 3. Computed average pitch RAO, Pabs , and efficiency for the multiple WEC rotors.

Side (Rotors 1 and 3) Center (Rotor 2)


Spacing Pitch RAO Average Pabs Efficiency Pitch RAO Average Pabs Efficiency
(rad/m) (kW) (%) (rad/m) (kW) (%)

One-Way
Single 0.780 15.238 65.39 0.780 15.238 65.39
WEC
rotor 0.802 14.045 60.27 0.802 14.045 60.27

0.785 (101) 14.096 (93) 60.49 0.751 (96) 13.308 (87) 57.11
0.8 × W
0.825 (103) 12.376 (88) 53.11 0.796 (99) 12.004 (85) 51.51
Multiple
WEC 0.777 (100) 13.692 (90) 58.76 0.770 (99) 13.096 (86) 56.20
1.0 × W
rotors 0.827 (103) 12.405 (88) 53.23 0.810 (101) 12.188 (87) 52.30

0.768 (98) 13.703 (90) 58.80 0.770 (99) 13.184 (87) 56.57
1.2 × W
0.826 (103) 12.398 (88) 53.20 0.820 (102) 12.336 (88) 52.30
Note: First row: Fully nonlinear (CFD simulations); Second row: Weakly nonlinear (Potential); Values in parentheses
represent the percentage ratio of a single WEC rotor to each rotor of a multi-WEC-rotor system.

5.4. Effect of Wave Steepness


For a WEC device, the response motion and absorbed power are hindered by the
incident wave height. The wave height effect was analyzed to evaluate the accuracy of the
present weakly nonlinear formulations with wave steepness compared with fully nonlinear
solutions. Wave steepness values of H/λ = 0.04 and 0.06 with a wide range of the one-way
PTO load torque system were evaluated, with all other parameters similar to the values
in Figure 9. When the maximum average Pabs and efficiency were achieved using the
two nonlinear models, an exponential increase and decrease was observed, respectively,
with an increase in the wave steepness from 0.03 to 0.06 (Figures 9 and 14). The maximum
average Pabs was determined to be 46 kW, and the associated efficiency was 49% for the fully
nonlinear model and 1% lower for the weakly nonlinear model. As shown in Figure 14,
the trends and behaviors are similar to those for H/λ = 0.03 (Figure 9). The nature of
the quadratic behavior with τ pto was preserved, but the statistical measure of the data
with the fitted regression R2 value variance increased with increasing wave steepness
(Figure 14). Better R2 variance can be achieved by fitting higher-order interpolation curves
for cases with higher wave steepness. One apparent variation from H/λ = 0.03 was that
the susceptibility of Pabs decreased with the PTO load as the wave steepness increased. At
H/λ = 0.03, deviations were observed between the weakly and fully nonlinear solutions in
the long-range PTO load torque and increase with the increase in wave steepness. Possible
reasons were explored to understand the disparities between the numerical solutions. One
reason for this could be the use of a modified PTO load mechanism in weakly nonlinear
.
formulations. For several seconds, the large PTO load torque tended to be in the ξ crit zone,
resulting in the application of a linear PTO load torque, as shown in Figures 3b and 15 (top).
The PTO load torque system was applied well in the fully nonlinear solution (Figure 15,
bottom). Another possible reason could be the increasing prevalence of the nonlinear
behavior. To obtain a thorough understanding of the physical flow evolution, the wave
elevation around the WEC rotor is shown in Figure 16 for different wave steepness values
at a fixed τ pto value. For H/λ = 0.06, the wave elevation exhibits a significant increase in the
nonlinear flow interaction in the vicinity of the WEC rotor, particularly during the extreme
downward and backward positions, compared with H/λ = 0.03 (Figure 16). Near-field
phenomena, such as flow separation, capsizing, slamming, and turbulent waves, are clearly
observed for H/λ = 0.06. Such a violent flow around the WEC rotor is associated with drag
losses, where the pressure and frictional drag are anticipated to dominate in H/λ = 0.06
nergies 2024, 17, 898

Energies 2024, 17, 898 19 of 28

formulations could not have captured all the complex flow field interactions ar
and to
WEC a lesser
rotor, andextent
the in H/λ = 0.03. Hence,
implemented the present
modified PTOweakly nonlinear
load torque formulations
system may have re
could not have captured all the complex flow field interactions around the WEC rotor, and
the
thedeviation
implementedobserved
modifiedwith
PTO increasing wave may
load torque system steepness in theinlong-range
have resulted the deviationPTO loa
application
observed with(Figure 14).wave steepness in the long-range PTO load torque application
increasing
(Figure 14).

Fully nonlinear (CFD) Weakly nonlinear (Potential)


Quadratic fit (CFD) Quadratic fit (Potential)
80 100

60 80

Efficiency (%)
Average Pabs

60
[kW]

40
40
20
20
0 0

0 50 100 150 200 250 300 0 50 100 150 200 250 300
pto (kNm) pto (kNm)

Average Pabs = − 0.002 ×τ pto


2
+ 0.5193 ×τ pto − 0.9212; R 2 = 0.9876 (CFD)
Quadratic fit

equations Average Pabs = − 0.0032 ×τ pto + 0.5421×τ pto − 1.8825; R = 0.9103(Potential)
2 2

(a)

80 100

60 80
Efficiency (%)
Average Pabs

60
[kW]

40
40
20
20
0 0

0 100 200 300 400 500 0 100 200 300 400 500
pto (kNm) pto (kNm)

Quadratic fit 
Average Pabs = − 0.0009 ×τ pto + 0.3904 ×τ pto − 6.936; R = 0.8747 (CFD)
2 2

equations Average P = − 0.0025 ×τ 2 + 0.6195 ×τ − 2.2508; R 2 = 0.9029 (Potential)


 abs pto pto

(b)
Figure
Figure14. Variation
14. Variation of average
of average Pabs
Pabs and and efficiency
efficiency with
with applied applied
one-way one-way
PTO load PTO
torque for load torque
different
wave steepness values. (a) Wave steepness H/λ = 0.04, (b) Wave steepness H/λ = 0.06.
ent wave steepness values. (a) Wave steepness H/λ = 0.04, (b) Wave steepness H/λ = 0.06.

120 kNm 150 kNm 200 kNm 250 kNm


Weakly nonlinear (Potential)
0
PTO load torque

-100
[kNm]

-200

-300
0 Fully nonlinear (CFD)
oad torque

-100
Nm]
equations 
Average Pabs = − 0.0025 ×τ pto + 0.6195 ×τ pto − 2.2508; R = 0.9029 (Potential)
2 2

(b)

Energies 2024, 17, 898 Figure 14. Variation of average Pabs and efficiency with applied one-way PTO load torque for d
20 of 28
ent wave steepness values. (a) Wave steepness H/λ = 0.04, (b) Wave steepness H/λ = 0.06.

120 kNm 150 kNm 200 kNm 250 kNm


Weakly nonlinear (Potential)
0

PTO load torque


-100

[kNm]
-200

-300
0 Fully nonlinear (CFD)
PTO load torque

-100
[kNm]

-200

-300

20 25 30 35 40
Time (s)
rgies 2024, 17, 898 20 of
Figure15.
Figure 15. Time
Time history
history of
of the
the instantaneous variationofofPTO
instantaneous variation PTOload
loadtorque
torquewith
withdifferent
different nume
models. models.
numerical

Figure 16. Free


Figure surface
16. Free augmentation
surface duetoto
augmentation due nonlinear
nonlinear flowflow interaction
interaction in theofvicinity
in the vicinity the WECof the W
rotor rotor
during the extreme downward and backward positions for different wave steepness value
during the extreme downward and backward positions for different wave steepness values.

5.5. Effect of Irregular Wave Conditions


The operation and high-sea conditions were selected based on the testing field site
explore the performance and interaction of the WEC rotor in irregular waves. The testi
field site was located on the west coast of Jeju Island, South Korea, as shown in Table
Energies 2024, 17, 898 21 of 28

5.5. Effect of Irregular Wave Conditions


The operation and high-sea conditions were selected based on the testing field site to
explore the performance and interaction of the WEC rotor in irregular waves. The testing
field site was located on the west coast of Jeju Island, South Korea, as shown in Table 4. The
best wave to express irregular waves is a superposition of several cosine waves, which are
characterized by frequency and wave height. In this study, the superposition of 200 linear
cosine waves with random phase values was investigated, and the analytical form, given
as the Joint North Sea Wave Project (JONSWAP), is expressed as

 −4 ! ( ω − ω p )2
exp (− )

−5 5 ω 2σ2 ω 2p
S J (ω ) = βω exp − .γ , (12)
4 ωp

where Hs is the significant wave height, σ is the variance of the wave record, ω p is the peak
frequency, γ is the peak enhancement factor, and β is the scale factor; β is expressed as
Energies 2024, 17, 898 21 of 28
  −1
5 2 4 0.925
β= H ω 1.15 + 0.168γ − (13)
16 s p 1.909 + γ
Table 4. Tested irregular wave conditions.
Table 4. Tested irregular wave conditions.
Spectral Peak Pa- Wave Power (kN/m)
Sea Wave
Sea Wave
Spectral Peak
rameter ρWave
g 2 Power (kN/m) 2
Conditions Parameters Parameter ρg2 × (0.865) × H2 s××T
Tp
Conditions Parameters (γ)
(γ) 64π
64π × (0.865) × Hs p
Operational Hs = 2 m; Tp = 6.65 s 11.31
Operational Hs = 2 m; Tp = 6.65 s 1.2 11.31
High 1.2
High Hs s==4.75
H 4.75m;m;
TpT=p 8.62
= 8.62
s s 82.66
82.66

Figure 17 shows a comparison of the wave elevations generated by employing both


Figure 17 shows a comparison of the wave elevations generated by employing both
weakly nonlinear and fully nonlinear models, in addition to the target wave spectrum
weakly nonlinear and fully nonlinear models, in addition to the target wave spectrum
under the design wave conditions. The numerical spectra generated by both models were
under the design wave conditions. The numerical spectra generated by both models were
in close agreement with the target spectrum, except for a slight disparity after the peak
in close agreement with the target spectrum, except for a slight disparity after the peak
value. Notably, the fully nonlinear model produced slightly higher predictions than both
value. Notably, the fully nonlinear model produced slightly higher predictions than both
the weakly nonlinear model and the target spectrum beyond the peak value. Nevertheless,
the weakly nonlinear model and the target spectrum beyond the peak value. Nevertheless,
the overall comparison between the numerical models and target spectrum demonstrates
the overall comparison between the numerical models and target spectrum demonstrates a
a satisfactory level of agreement.
satisfactory level of agreement.

Target Fully nonlinear (CFD) Weakly nonlinear (Potential)


Wave elevation (m) Wave elevation (m)

1.5
0 0.5

-1.5 0.4
PSD (m2.s)

0.3
0 200 400 600 800
1.5 0.2

0 0.1
-1.5 0

0 200 400 600 800 0.4 0.8 1.2 1.6 2


(rad/s)

Figure 17.
Figure 17. Comparison
Comparison of
of wave
wave elevations:
elevations: weakly
weakly nonlinear
nonlinear and
and fully
fully nonlinear
nonlinear models
modelsvs.
vs. target
target
spectrum under designed wave conditions.
spectrum under designed wave conditions.

Further testing
Further testing was
was carried
carried out
out with
with the
the application
application of
of one-way
one-way PTO
PTO load
load torque
torque in
in
irregular waves.
irregular waves. The PTO load
load torque
torque was
was kept
kept constant
constant at
at 50
50 kNm
kNm for
for the
the operational
operational
settings and at 100 kNm for the high-sea wave conditions. Figures 18 and 19 compare the
instantaneous pitch response, angular velocity, PTO load torque, and power absorption
obtained using the two numerical models. The amplitudes of the pitch response and an-
gular velocity (Figures 18 and 19, top and middle) are notably different and comparatively
high in the case of the weakly nonlinear model compared to that with the fully nonlinear
Energies 2024, 17, 898 22 of 28

settings and at 100 kNm for the high-sea wave conditions. Figures 18 and 19 compare the
instantaneous pitch response, angular velocity, PTO load torque, and power absorption
obtained using the two numerical models. The amplitudes of the pitch response and angular
velocity (Figures 18 and 19, top and middle) are notably different and comparatively high
in the case of the weakly nonlinear model compared to that with the fully nonlinear model.
The subsequent estimation of Pabs showed a similar pattern. To estimate the average Pabs
under irregular wave conditions, the time history without the ramp time was considered,
and the mean instantaneous estimated power for the overall simulation time was computed.
This is represented by a red line in each figure, with the corresponding value highlighted.
Furthermore, the power statistics are estimated and tabulated in Table 5, based on the
cases presented in Figures 18 and 19. The predicted average Pabs under high sea conditions
obtained using the fully nonlinear model was approximately two times greater than that
under operational sea conditions, with a three-to four-fold reduction in efficiency. To
Energies 2024, 17, 898 22 of 28
ensure a direct comparison between the numerical models, the power spectral density
(PSD) is plotted in Figures 18c and 19c. The figures show that the PSD comparison under
operational wave conditions is better than high-sea wave conditions. The deterioration in
the results can be attributed to the following reasons. First, the deviation observed in the
could be attributed to the observed differences. Second, the transition from operational to
fully nonlinear model after the peak PSD value, as shown in Figure 17, could be attributed
high-sea
to the conditions introduces
observed differences. an increased
Second, strong
the transition from nonlinear
operational interaction between the ir-
to high-sea conditions
regular waves an
introduces and the WEC
increased rotor.
strong This strong
nonlinear nonlinear
interaction interaction
between can result
the irregular waves in
andcomplex
flowthe
behaviors,
WEC rotor. including wave
This strong run-up,
nonlinear slamming,
interaction and breaking
can result in complex offlow
waves, and in some
behaviors,
including
cases, wave run-up,
overtopping of theslamming,
WEC rotor andasbreaking of waves,
manifested and
and in some
shown incases, overtopping
Figure 16. Although a
of the WEC rotor as manifested and shown in Figure 16. Although
fully nonlinear model may capture the complex flow physics around the WEC a fully nonlinear modelrotor, a
may capture the complex flow physics around the WEC rotor, a weakly nonlinear model
weakly nonlinear model lacks these complex effects, leading to increased discrepancies as
lacks these complex effects, leading to increased discrepancies as the WEC rotor interaction
the WEC
changesrotor
frominteraction
operational changes from
to high-sea waveoperational
conditions. to high-sea wave conditions.
Pitch response

Pitch response
[deg]

[deg]
Angular velocity

Angular velocity
PTO load torque

PTO load torque


[kNm]

[kNm]
[rad/s]

[rad/s]
[kW]

[kW]
Pabs

Pabs

(a) (b)

Figure 18. Cont. Fully nonlinear (CFD)


Weakly nonlinear (Potential)
500

400
(kW2.s)

300
Pabs

Pabs
[kW

[kW
Energies 2024, 17, 898 23 of 28

(a) (b)

Fully nonlinear (CFD)


Weakly nonlinear (Potential)
500

400
PSD (kW2.s)
300

200

100

0 2 4 6 8 10
(rad/s)
(c)
Energies 2024, 17, 898 Figure 18. Comparison
Figure 18. Comparison ofof(a)
(a)weakly and(b)
weakly and (b)fully
fully nonlinear
nonlinear modelsmodels
with with instantaneous
instantaneous variationvariation
of 23 of of
28
pitchpitch
response, angular velocity, PTO load torque, and P using one-way load torque
response, angular velocity, PTO load torque, and Pabs using one-way load torque (50 kNm), and
abs (50 kNm),
and (c) spectraldensity
(c) spectral density of the
of the extracted
extracted Pabs for for operational
Pabsoperational wave conditions
wave conditions of WEC rotor.
of WEC rotor.

100 100

0 0

-100 -100
0 200 400 600 800 0 200 400 600 800
Time (s) Time (s)
200 200
2 2

0 0 0 0

-2 -2
-200 -200
0 200 400 600 800 0 200 400 600 800
Time (s) Time (s)
200 200

100 100

0
20.03 kW 0
22.27 kW
0 200 400 600 800 0 200 400 600 800
Time (s) Time (s)

(a) (b)

Figure 19. Cont. Fully nonlinear (CFD)


Weakly nonlinear (Potential)
1200

800
PSD (kW2.s)

400
100 100

0
20.03 kW 0
22.27 kW
0 200 400 600 800 0 200 400 600 800
Energies 2024, 17, 898 Time (s) Time (s) 24 of 28

(a) (b)

Fully nonlinear (CFD)


Weakly nonlinear (Potential)
1200

PSD (kW2.s) 800

400

0 2 4 6 8 10
(rad/s)
(c)
Figure
Figure 19. Comparisonofof(a)
19. Comparison (a)weakly
weakly and
and(b)(b)fully
fullynonlinear
nonlinearmodels with with
models instantaneous variationvariation
instantaneous of of
pitch response, angular velocity, PTO load torque,
pitch response, angular velocity, PTO load torque, and and P using one-way load torque (100 kNm),
abs Pabs using one-way load torque (100 kNm)
and and (c) spectral
(c) spectral densityofofthe
density extracted PP
the extracted absfor
abs forhigh seasea
high wave conditions
wave of WEC
conditions of rotor.
WEC rotor.

Table
Table 5. Power
5. Power statisticsusing
statistics using different
different numerical
numericalmodels for different
models irregular
for different wave conditions.
irregular wave conditions.
Model
Model
One-WayOne-Way PTO Sea WeaklyNonlinear
Nonlinear Fully Nonlinear
Power Weakly Fully Nonlinear
Power PTO Load Sea Condition
Load (kNm) Condition Standard Standard
Max. Standard De- Avg. Standard
(kNm) Avg.Avg. Max. Deviation Avg. Max. Max.Deviation
viation Deviation
50 Operational 9.34 81.88 15.11 12.49 68.69 17.06
Pabs 50 100 Operational
High sea 9.3420.03 81.88
187.86 15.11
32.77 12.49 129.3 68.69 29.69 17.06
22.27
Pabs
100 High sea 20.03 187.86 32.77 22.27 129.3 29.69
6. Conclusions
6. Conclusions
In this study, a 3D WEC rotor was subjected to a nonlinear investigation using a
weakly
In thisnonlinear
study, amodel based rotor
3D WEC on thewas
linearsubjected
potential flow
to a theory and ainvestigation
nonlinear fully nonlinear using a
model
weakly based on CFD
nonlinear modelby solving
based the
on RANS equation.
the linear The nonlinear
potential restoring
flow theory andmoment
a fullywas
nonlinear
incorporated into the weakly nonlinear model by considering the instantaneous variable
model based on CFD by solving the RANS equation. The nonlinear restoring moment was
secondary moment of the water-plane area of the WEC rotor. A Coulomb-type PTO
load torque was incorporated by considering one- and two-way loading systems, and the
movement of the WEC was restricted during exertion until the external force acting on
it exceeded the applied load torque. Fully nonlinear simulations were performed and
compared to determine the limitations and degree of accuracy achieved by the weakly
nonlinear model. The experimental results were used for numerical validation. After
validating the experimental results, the designed sensitive parameters of the WEC rotor
were numerically studied, and the effects of the PTO load torque, wave period, number
of WEC rotors, and wave steepness in regular waves were examined. The performance of
the WEC rotor in irregular waves under operating and high sea wave conditions was also
studied. The conclusions are summarized below:
• With 3D nonlinear simulations, adopting a one-way PTO load torque in conjunction
with a WEC rotor can be highly efficient compared with adopting a two-way PTO
load torque.
Energies 2024, 17, 898 25 of 28

• The average Pabs varied quadratically with the PTO load torque, and the WEC rotor
exhibited higher sensitivity in the one-way PTO system, even in steeper waves, albeit
with a decrease in the maximum efficiency.
• The pitch RAO and absorbed Pabs were higher during the resonance period, with
weakly nonlinear models exhibiting lower values by the fully nonlinear model for
both PTO load–torque systems within the considered wave period range.
• For the evaluated range of spacing between multiple WEC rotors, the center rotor
consistently exhibits lower pitch RAO and average Pabs (PR ≤ 5) compared to the side
rotors, regardless of the numerical model.
• Using a weakly nonlinear model can maintain accuracy up to moderate wave steep-
ness; however, its applicability decreases with an increase in wave steepness. Fully
nonlinear simulations can address a high degree of nonlinearity.
• In irregular waves, the estimated average Pabs is twice that under high-sea conditions,
and the associated efficiency is three to four times lower than that under operational
sea conditions.
The present investigation acknowledges and highlights the influence of nonlinear
analysis, encompassing both weakly and fully nonlinear models, on power absorption and
efficiencies of WECs. This underscores the essential requirement for customized strategies
in practical applications, ensuring resilience and effectiveness in the face of diverse and
severe sea conditions. This study will be extended in the future by comparing 3D nonlinear
models with real-time testing data.

Author Contributions: Conceptualization, S.K.P. and Y.H.B.; methodology, S.K.P.; validation, S.K.P.
and D.K.; formal analysis, S.K.P. and D.K.; investigation, S.K.P.; resources, Y.H.B.; writing—original
draft preparation, S.K.P.; writing—review and editing, S.K.P. and Y.H.B.; supervision, Y.H.B.; funding
acquisition, Y.H.B. All authors have read and agreed to the published version of the manuscript.
Funding: This research was supported by the Basic Science Research Program through the National
Research Foundation of Korea (NRF), funded by the Ministry of Education (No. 2022R1I1A1A01069442).
This work was also supported by the National Research Foundation of Korea (NRF) grant funded by
the Korea government (MSIT). (No. 2021R1A2C1014600).
Data Availability Statement: Dataset available on request from the authors.
Conflicts of Interest: The authors declare no conflicts of interests.

Nomenclature

u Velocity vector
(x, y, z) Cartesian coordinate system
u, v, w Mean velocity components in the x-,y-and z-directions
u′ , v′ , w′ Fluctuating velocity components in the x-,y- and z-directions
m Viscosity
υ Kinematic viscosity
p, p Pressure and mean pressure
t time
t1 Time-lag
g Acceleration due to gravity
γf Forcing coefficient
r Mixture of fluid density
δ∗f orced Forced solution
δcomputed Actual solution
λ Wave length
K Hydrostatic linear stiffness
h Base cell size
a Volume fraction
Energies 2024, 17, 898 26 of 28

fwater Density of water


fair Density of air
k Scaling factor
W Half width of the WEC rotor
xg longitudinal center of gravity
zg Vertical center of gravity
. ..
ξ, ξ and ξ Displacement, velocity and acceleration of the WEC rotor
rm Distance vector
∀ Submerged volume of the WEC rotor
τ pto Applied PTO load
R(ξ) Nonlinear restoring moment
τ vis Linear viscosity
.
−ξ crit Negative critical displacement
.
+ξ crit Positive critical displacement
d Water depth
Pabs Absorbed power
T Time period
H Wave height
ε num Total uncertainty
εm Mesh size uncertainty
εt Time step uncertainty
ε Cm Domain size uncertainty
∆tre f Reference time step
∆zre f Reference mesh size
Cmref Reference domain size
σfit Total standard deviation
Pabs,ref Reference Absorbed power
Pw Wave power
H/λ Wave steepness
Pext Excitation power
Prad Radiated power
w Angular velocity
Hs Significant wave height
σ Variance of the wave record
ωp Peak frequency
γ Peak enhancement factor
β Scale factor
Tp Peak period
PR Percentage ratio of multiple individual rotors to a single WEC rotor

References
1. McCormick, M.E. (Ed.) ; Ocean Wave Energy Conversion; Dover Publications: New York, NY, USA, 2013.
2. Salter, S. Wave power. Nature 1974, 249, 720–724. [CrossRef]
3. Evans, D.V. A theory for wave-power absorption by oscillating bodies. J. Fluid Mech. 1976, 77, 1–25. [CrossRef]
4. Mei, C.C. Power extracted from water waves. J. Ship Res. 1976, 20, 63–66. [CrossRef]
5. Budal, K. Theory for absorption of wave power by a system of interacting bodies. J. Ship Res. 1977, 21, 248–253. [CrossRef]
6. Mynett, A.E.; Serman, D.D.; Mei, C.C. Characteristics of Salter’s cam for extracting energy from ocean waves. Appl. Ocean Res.
1979, 1, 13–20. [CrossRef]
7. Pizer, D. The Numerical Prediction of the Performance of a Solo Duck; European Wave Energy Symposium: Edinburgh, UK, 1993;
pp. 129–137.
8. Skyner, D. Solo Duck Linear Analysis; University of Edinburgh: Edinburgh, UK, 1987.
9. Ferri, F.; Pecher, A.F.S.; Kofoed, J.P. Numerical Analysis of a Large Floating Wave Energy Converter with Adjustable Structural
Geometry. In Proceedings of the ISOPE The 25th International Ocean and Polar Engineering Conference, Kona, HI, USA, 21–26
June 2015; pp. 1025–1033.
10. Rapuc, S. Numerical Study of the WEPTOS Single Rotor. Ph.D. Thesis, Aalborg University and Technical University of Denmark
(DK), Aalborg, Denmark, 2012.
11. Folley, M. Numerical Modelling of Wave Energy Converters: State-of-the-Art Techniques for Single Devices and Arrays; Academic Press:
Cambridge, MA, USA, 2016.
Energies 2024, 17, 898 27 of 28

12. Heo, K.; Choi, Y.R. Numerical Investigation of Multi-body Wave Energy Converters’ Configuration. J. Ocean Eng. Technol. 2022,
36, 132–142. [CrossRef]
13. Falnes, J.; Kurniawan, A. Ocean Waves and Oscillating Systems: Linear Interactions Including Wave-Energy Extraction; Cambridge
University Press: Cambridge, UK, 2020; Volume 8.
14. Penalba, M.; Giorgi, G.; Ringwood, J.V. Mathematical modelling of wave energy converters: A review of nonlinear approaches.
Renew. Sustain. Energy Rev. 2017, 78, 1188–1207. [CrossRef]
15. Gilloteaux JC, Bacelli G; Ringwood J. A non-linear potential model to predict large-amplitude-motions: Application to a multi-
body wave energy converter. In Proceedings of the 10th World Renewable Energy Conference, Glasgow, UK, 19–25 July 2008;
pp. 934–940.
16. Gilloteaux, J.C.; Ducrozet, G.; Babarit, A.; Clement, A.H. Non-linear model to simulate large amplitude motions: Application
to wave energy conversion. In Proceedings of the 22nd International Workshop on Water Waves and Floating Bodies, Plitvice,
Croatia, 15–18 April 2007; pp. 97–100.
17. Guérinel, M.; Alves, M.; Sarmento, A. Nonlinear modelling of the dynamics of a free floating body. In Proceedings of the 9th
European Wave and Tidal Energy Conference (EWTEC), Southampton, UK, 5–9 September 2011.
18. Merigaud, A.; Gilloteaux, J.C.; Ringwood, J.V. A nonlinear extension for linear boundary element methods in wave energy device
modelling. In Proceedings of the International Conference on Offshore Mechanics and Arctic Engineering, New York, NY, USA,
1–6 July 2012; American Society of Mechanical Engineers: New York, NY, USA, 2012; Volume 44915, pp. 615–621.
19. Viuff, T.H.; Andersen, M.T.; Kramer, M.; Jakobsen, M.M. Excitation forces on point absorbers exposed to high order non-linear
waves. In Proceedings of the European Wave and Tidal Energy Conference. Technical Committee of the European Wave and
Tidal Energy Conference, Aalborg, Denmark, 2–5 September 2013.
20. Lawson, M.; Yu, Y.H.; Nelessen, A.; Ruehl, K.; Michelen, C. Implementing nonlinear buoyancy and excitation forces in the
wec-sim wave energy converter modeling tool. In Proceedings of the International Conference on Offshore Mechanics and
Arctic Engineering, 8–13 June 2014; American Society of Mechanical Engineers: New York, NY, USA, 2014; Volume 45547,
p. V09BT09A043.
21. Zurkinden, A.S.; Ferri, F.; Beatty, S.; Kofoed, J.P.; Kramer, M.M. Non-linear numerical modeling and experimental testing of a
point absorber wave energy converter. Ocean Eng. 2014, 78, 11–21. [CrossRef]
22. Wolgamot, H.A.; Fitzgerald, C.J. Nonlinear hydrodynamic and real fluid effects on wave energy converters. Proc. Inst. Mech. Eng.
Part A J. Power Energy 2015, 229, 772–794. [CrossRef]
23. Poguluri, S.K.; Cho, I.H.; Bae, Y.H. A study of the hydrodynamic performance of a pitch-type wave energy converter–rotor.
Energies 2019, 12, 842. [CrossRef]
24. Kim, D.; Poguluri, S.K.; Ko, H.S.; Lee, H.; Bae, Y.H. Numerical and Experimental Study on Linear Behavior of Salter’s Duck Wave
Energy Converter. J. Ocean Eng. Technol. 2019, 33, 116–122. [CrossRef]
25. van Rij, J.; Yu, Y.H.; Guo, Y.; Coe, R.G. A wave energy converter design load case study. J. Mar. Sci. Eng. 2019, 7, 250. [CrossRef]
26. Katsidoniatski, E.; Yu, Y.H.; Goteman, M. Midfidelity Model Verification for a Point-Absorbing Wave Energy Converter with Linear
Power Takeoff (No. NREL/CP-5700-78569); National Renewable Energy Lab. (NREL): Golden, CO, USA, 2021.
27. Van’t Hoff, J. Hydrodynamic Modelling of the Oscillating Wave Surge Converter. Ph.D. Thesis, Queen’s University Belfast,
Belfast, UK, 2009.
28. Folley, M.; Whittaker, T.J.T.; Henry, A. The effect of water depth on the performance of a small surging wave energy converter.
Ocean Eng. 2007, 34, 1265–1274. [CrossRef]
29. Falnes, J. Optimum control of oscillation of wave-energy converters. Int. J. Offshore Polar Eng. 2002, 12, 147–155.
30. Bjarte-Larsson, T.; Falnes, J. Laboratory experiment on heaving body with hydraulic power take-off and latching control. Ocean
Eng. 2006, 33, 847–877. [CrossRef]
31. Babarit, A.; Clément, A.H. Optimal latching control of a wave energy device in regular and irregular waves. Appl. Ocean Res.
2006, 28, 77–91. [CrossRef]
32. Babarit, A.; Duclos, G.; Clément, A.H. Comparison of latching control strategies for a heaving wave energy device in random sea.
Appl. Ocean Res. 2004, 26, 227–238. [CrossRef]
33. Lopes, M.F.P.; Hals, J.; Gomes, R.P.F.; Moan, T.; Gato, L.M.C.; Falcão, A.D.O. Experimental and numerical investigation of
non-predictive phase-control strategies for a point-absorbing wave energy converter. Ocean Eng. 2009, 36, 386–402. [CrossRef]
34. Sheng, W.; Alcorn, R.; Lewis, A. A new latching control technology for improving wave energy conversion. In Proceedings of the
5th International Conference on Ocean Energy (ICOE), Halifax, NS, Canada, 4–6 November 2014.
35. Sheng, W.; Alcorn, R.; Lewis, A. On improving wave energy conversion, part II: Development of latching control technologies.
Renew. Energy 2015, 75, 935–944. [CrossRef]
36. António, F.D.O. Modelling and control of oscillating-body wave energy converters with hydraulic power take-off and gas
accumulator. Ocean Eng. 2007, 34, 2021–2032.
37. Costello, R.; Ringwood, J.; Weber, J. Comparison of two alternative hydraulic PTO concepts for wave energy conversion. In
Proceedings of the 9th European Wave and Tidal Energy Conference (EWTEC), Southampton, UK, 5–9 September 2011; School of
Civil Engineering and the Environment, University of Southampton: Southampton, UK, 2011.
38. Babarit, A.; Guglielmi, M.; Clément, A.H. Declutching control of a wave energy converter. Ocean Eng. 2009, 36, 1015–1024.
[CrossRef]
Energies 2024, 17, 898 28 of 28

39. António, F.D.O. Phase control through load control of oscillating-body wave energy converters with hydraulic PTO system. Ocean
Eng. 2008, 35, 358–366.
40. Wilkinson, L.; Doherty, K.; Nicholson, J.; Whittaker, T.; Sandy, D. Modelling the performance of a modular flap-type wave energy
converter. In Proceedings of the 11th European Wave and Tidal Energy Conference, Nantes, France, 6–11 September 2015.
41. Wilkinson, L.; Whittaker, T.J.T.; Thies, P.R.; Day, S.; Ingram, D. The power-capture of a nearshore, modular, flap-type wave energy
converter in regular waves. Ocean Eng. 2017, 137, 394–403. [CrossRef]
42. Wilkinson, L. An Assessment of the Nearshore Modular Flap-Type WEC. Ph.D. Thesis, University of Edinburgh: Edinburgh, UK,
2018.
43. Orszaghova, J.; Wolgamot, H.; Draper, S.; Rafiee, A. Non-linear behaviour of an idealised submerged WEC with Coulomb
damping. In Proceedings of the 12th European Wave and Tidal Energy Conference, Cork, UK, 27 August–1 September 2017;
Technical Committee of the European Wave and Tidal Energy Conference: Uppsala, Sweden, 2017; pp. 1088-1–1088-9.
44. Gong, S.K.; Gao, J.L.; Mao, H.F. Investigations on fluid resonance within a narrow gap formed by two fixed bodies with varying
breadth ratios. China Ocean Eng. 2023, 37, 962–974. [CrossRef]
45. Gao, J.; Chen, H.; Zang, J.; Chen, L.; Wang, G.; Zhu, Y. Numerical investigations of gap resonance excited by focused transient
wave groups. Ocean Eng. 2020, 212, 107628. [CrossRef]
46. Schmitt, P. Investigation of the Near Flow Field of Bottom Hinged Flap Type Wave Energy Converters. Ph.D. Thesis, Queen’s
University Belfast, Belfast, UK, 2013.
47. Sjökvist, L.; Wu, J.; Ransley, E.; Engström, J.; Eriksson, M.; Göteman, M. Numerical models for the motion and forces of
point-absorbing wave energy converters in extreme waves. Ocean Eng. 2017, 145, 1–14. [CrossRef]
48. Windt, C.; Davidson, J.; Ringwood, J.V. High-fidelity numerical modelling of ocean wave energy systems: A review of computa-
tional fluid dynamics-based numerical wave tanks. Renew. Sustain. Energy Rev. 2018, 93, 610–630. [CrossRef]
49. Devolder, B.; Stratigaki, V.; Troch, P.; Rauwoens, P. CFD simulations of floating point absorber wave energy converter arrays
subjected to regular waves. Energies 2018, 11, 641. [CrossRef]
50. Poguluri, S.K.; Ko, H.S.; Bae, Y.H. CFD investigation of pitch-type wave energy converter-rotor based on RANS simulations.
Ships Offshore Struct. 2020, 15, 1107–1119. [CrossRef]
51. Poguluri, S.K.; Kim, D.; Ko, H.S.; Bae, Y.H. Performance analysis of multiple wave energy converters due to rotor spacing. J.
Ocean Eng. Technol. 2021, 35, 229–237. [CrossRef]
52. Windt, C.; Davidson, J.; Ringwood, J.V. Investigation of turbulence modeling for point-absorber-type wave energy converters.
Energies 2020, 14, 26. [CrossRef]
53. Ransley, E.J.; Brown, S.A.; Hann, M.; Greaves, D.M.; Windt, C.; Ringwood, J.; Davidson, J.; Schmitt, P.; Yan, S.; Wang, J.X.; et al.
Focused wave interactions with floating structures: A blind comparative study. Proc. Inst. Civ. Eng.-Eng. Comput. Mech. 2021, 174,
46–61. [CrossRef]
54. Ko, H.S.; Kim, S.; Bae, Y.H. Study on Optimum Power Take-Off Torque of an Asymmetric Wave Energy Converter in Western Sea
of Jeju Island. Energies 2021, 14, 1449. [CrossRef]
55. Ha, Y.J.; Park, J.Y.; Shin, S.H. Numerical Study of Non-Linear Dynamic Behavior of an Asymmetric Rotor for Wave Energy
Converter in Regular Waves. Processes 2021, 9, 1477. [CrossRef]
56. Peric, M. Best Practices for Flow Simulations with Waves, Star Global Conference; Siemens AG: Berlin, Germany, 2017.
57. Enger, S.; Perić, M.; Monteiro, H. Coupling of 3D Numerical Solution Method Based on Navier-Stokes Equations with Solutions Based on
Simpler Theories; CILAMCE: Fortaleza, Brazil, 2014.
58. Cummins, W.E.; Iiuhl, W.; Uinm, A. The Impulse Response Function and Ship Motions; Department of the NAVY: Washington, DC,
USA, 1962.
59. Cummins, C.; Viola, I.M.; Mastropaolo, E.; Nakayama, N. The effect of permeability on the flow past permeable disks at low
Reynolds numbers. Phys. Fluids 2017, 29, 097103. [CrossRef]

Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual
author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to
people or property resulting from any ideas, methods, instructions or products referred to in the content.

You might also like