You are on page 1of 26

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/323212609

Estimation of numerical uncertainty in computational fluid dynamics


simulations of a passively controlled wave energy converter

Article  in  Proceedings of the Institution of Mechanical Engineers Part M Journal of Engineering for the Maritime Environment · February 2018
DOI: 10.1177/1475090217726884

CITATIONS READS

4 293

4 authors:

Weizhi Wang Minghao Wu


Norwegian University of Science and Technology Ghent University
6 PUBLICATIONS   8 CITATIONS    4 PUBLICATIONS   4 CITATIONS   

SEE PROFILE SEE PROFILE

Johannes Palm Claes Eskilsson


Chalmers University of Technology Aalborg University
27 PUBLICATIONS   151 CITATIONS    70 PUBLICATIONS   564 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Prediction and Mitigation of snap loads in mooring cables View project

Method of moving frames for scientific computing View project

All content following this page was uploaded by Weizhi Wang on 02 March 2018.

The user has requested enhancement of the downloaded file.


Estimation of numerical uncertainty in CFD simulations of a
passively controlled wave energy converter
Weizhi Wang, Minghao Wu, Johannes Palm and Claes Eskilsson
Department of Shipping and Marine Technology, Chalmers University of Technology, Sweden

Journal of Engineering for the Maritime Environment, 2018, 232 1, pp. 71-84.
DOI: http://dx.doi.org/10.1177/1475090217726884

Abstract
The wave loads and the resulting motions of floating wave energy converters (WECs)
are traditionally computed using linear radiation-diffraction methods. Yet for certain
cases such as survival conditions, phase control and WECs operating in the resonance
region, more complete mathematical models such as CFD are preferred and over the last
five years CFD has become more frequently used in the wave energy field. However, the
issue of numerical errors and uncertainties associated with CFD simulations have largely
been overlooked in the wave energy sector. In this paper we apply formal verification and
validation (V&V) techniques to CFD simulations of a passively controlled point absorber.
The phase control causes the motion response to be highly nonlinear even for almost linear
incident waves. First, we show that the CFD simulations have acceptable agreement to
experimental data. We then present a rigorous V&V study focusing on the solution
verification covering spatial and temporal discretization, iterative and domain modeling
errors. It is shown that the dominating source of errors are, as expected, the spatial
discretization, but temporal and iterative errors cannot be neglected. Using hexahedral
cells with an aspect ratio of unity and 30 cells per wave height we obtain results with
less than 5% uncertainty in motion response (except for surge) and restraining forces for
the buoy without phase control. The enhanced nonlinear response due to phase control
caused a large increase in numerical uncertainty, illustrating the difficulty to obtain reliable
solutions for highly nonlinear responses and that much denser meshes are required for such
cases.

Keywords: Wave energy converter, passive control, CFD modelling, numerical uncertainty,
verification and validation

1 Introduction
The wave loads and the motion response of floating wave energy converters (WECs) are
typically computed in the time-domain using linear radiation-diffraction methods. The hy-
drodynamic coefficients are computed by boundary element methods such as Nemoh (Babarit
and Delhommeau, 2015) and WAMIT(WAMIT Inc., 2016) and are then used in solving the
Cummins equations – see e.g. the open-source model WEC-Sim.(Yu et al., 2014) This type
of models is based on small-amplitude and small-motion assumptions. If applied within the

1
underlying assumptions, time-domain linear radiation-diffraction models are accurate and
highly efficient. However, there are several important cases where the assumptions of small
amplitude and small motion do not hold and where computational fluid dynamics (CFD)
models based on more complete model equations, such as fully nonlinear potential flow or
Navier-Stokes equations, should be used:

Survival A key issue for the wave energy sector is survival of the devices at extreme events. A
WEC operating in survival condition is subjected to steep and breaking waves so a model
coping with highly nonlinear waves is needed. How to accurately generate and simulate
extreme waves is still an open question,Tiron et al. (2015) but clearly the mathematical
model needs to handle nonlinearity and wave-wave interaction. Additionally, inherently
nonlinear events such as slamming and overtopping will often take place during storms.
An example of CFD simulations of slamming on a oscillating surge wave converter is
found in Henry et al.Henry et al. (2014) Chen et al.Chen et al. (2017) compared CFD
simulations with experimental tests of a point absorber in extreme waves. Frequent
overtopping was observed and it was reported that the difference between CFD and
linear radiation-diffraction solutions could be significant.

Resonance It is well-known that in the resonance region the motion response is not a one-
to-one mapping with regard to the wave height and that linear radiation-diffraction
models overestimate the motion response unless the drag coefficients are calibrated. Yu
and LiYu and Li (2013) compared CFD simulations of a self-reacting point absorber
in heave to linear simulations and illustrated the nonlinear response. Palm et al.Palm
et al. (2016) coupled a dynamic mooring solver with CFD and simulated a generic WEC
in six-degrees-of-freedom and captured the same nonlinear wave height dependence as
in corresponding physical experiments.

Control Application of control strategies are often put forward as the most promising way to
increase the power extraction and lower the cost of energy.Bull et al. (2013) Typically,
control will enhance the nonlinear response of the devices. In a recent study Giorgio
and RingwoodGiorgi and Ringwood (2016) implemented latching control for a heaving
sphere in CFD. It was shown that the use of CFD was required in order to get the
optimal latching duration correct.

CFD – and in this paper we by CFD mean the standard two-phase Reynolds-Averaged
Navier-Stokes (RANS) equations using the volume of fluid (VOF) approach for free-surface
capturing – is becoming increasingly popular for wave energy applications, see e.g. Davidson
et al.Davidson et al. (2015) and Wolgamot and FitzgeraldWolgamot and Fitzgerald (2015)
and the references therein. The rising trend is expected to continue as Kim et al.Kim et al.
(2016) recently presented a study comparing numerical wave tanks using CFD with physical
models for an offshore production unit. It was shown that the CFD-based numerical wave
tank had equal cost, but shorter project period and higher flexibility compared to traditional
test in wave basins.
CFD offers much in terms of completeness of the hydrodynamic model, but it is important
to keep in mind that the obtained solutions still contain errors and uncertainties. The question
of numerical errors and uncertainties has so far largely been overlooked in the CFD studies
concerning WECs. Some papers present basic grid convergence studies to illustrate that the

2
solutions are ‘grid-independent’.Bhinder et al. (2015); Eskilsson et al. (2015); Palm et al.
(2016); Vyzikas et al. (2017) However, that a solution is grid-independent neither implies that
the solution is properly converged nor that the numerical errors are insignificant and can be
disregarded. The lack of estimation of numerical errors among the wave energy community is
in stark contrast to the closely related field of ship hydrodynamics where standard verification
and validation (V&V) procedures (AIAA, 1998; Stern et al., 2001; Eça et al., 2010; Eça and
Hoekstra, 2014) are routinely used to estimate the numerical uncertainty associated with the
solutions. Wilson et al. (2001); Simonsen and Stern (2003); Guo et al. (2013); Pereira et al.
(2017) V&V procedures have also been applied to other fields of marine hydrodynamics such
as propulsion(Rijpkema and Vaz, 2011), tidal turbines(Gahree et al., 2016) and offshore wind
turbines.(Make and Vaz, 2015)
Needless to say, V&V involves two stages: (i) the verification step and (ii) the validation
step. The verification stage is commonly understood as making sure that the numerical
code is working correctly, so-called code verification. This is checked by comparing against
exact solutions, often relying on simplified equations having the same numerical operators.
In the validation stage we make sure that the underlying mathematical models approximate
the application problem under investigation, typically performed by comparing numerical
results to experimental test data. Yet, there is a second part of the verification stage that
is often omitted but should be performed ideally for every computational case: the solution
verification. Solution verification is described by Eça and HoekstraEça and Hoekstra (2014)
as how to “estimate the error/uncertainty of a given calculation, for which in general the exact
solution is not known”. It is here the numerical approaches of the V&V techniques come into
play.
Code verification can be considered upstream of the task of using CFD for WEC applica-
tions, although bespoke code development for WEC modules should go through this step (e.g.
PTO, control and mooring modules). Up to date almost all CFD studies concerning WECs
have been focused solely on the validation phase – the numerical solutions are compared to
experimental data.Yu and Li (2013); Schmitt and Elsaesser (2015); Palm et al. (2016); Chen
et al. (2017) To the authors best knowledge no study has yet done a solution verification for
CFD of WECs. So while there are validated CFD models for WEC applications, the numeri-
cal uncertainties related to the results obtained by these CFD models are still unknown. It is
generally accepted that the numerical uncertainties should be below 5% for the solutions to be
acceptable and reliable.(Eça et al., 2010) The main idea of using V&V techniques is to obtain
higher confidence in the numerical solutions. Thus facilitating numerical optimization of the
wave energy devices in as early development stages as possible, minimizing costly and time
consuming tank tests. It is also important to have knowledge of the numerical uncertainties
in order to compare with the uncertainties associated with experimental wave basin tests.
In the paper we will look into the solution verification stage and present estimates of
numerical uncertainty for VOF-RANS simulations of a passively controlled wave energy con-
verter, developed by CorPower Ocean AB.CorPower Ocean AB (2017) We will mainly follow
the V&V approach due to Eça et al.Eça et al. (2010); Eça and Hoekstra (2014) and focus
on the uncertainties associated with the numerical errors. We use a numerical wave tank
based on the incompressible VOF-RANS model in the open-source finite volume framework
OpenFOAM .(OpenFOAM,
R 2017; Weller. et al., 1998) OpenFOAM is perhaps the most
widely used numerical wave tank in the wave energy community,Davidson et al. (2015) and
thus the results obtained in this study have bearing outside the particular case of the CorPower

3
WEC but to the wider wave energy sector. The CorPower buoy is subjected to passive phase
control through a negative spring arrangement. In particular we investigate the influence of
phase control on the uncertainties of the simulations.
The paper is organized as follows. In section 2 we will outline the V&V procedureEça et al.
(2010) used to assess the numerical errors and how to estimate the uncertainties based on the
computed numerical errors. In section 3 we describe the CorPower buoy, the passive phase
control technology and the experimental tests used for validation. The numerical settings
used in the OpenFOAM solver and the series of grids used are presented in section 4. Section
5 presents the validation study, while section 6 deals with the solution verification stage of
the non-controlled device, divided into spatial and temporal errors, iterative errors and model
errors. Section 7 looks into the spatial uncertainties associated with the phase controlled
device. Finally, in section 8 we summaries the study and discuss the numerical settings
required in order to achieve results with acceptable numerical uncertainties.

2 Verification & validation


There are two main sources of errors and uncertainties in CFD simulations: modeling errors
and numerical errors. Modeling errors arise from simplifications and approximations of the real
problem in the numerical model. Typical examples are simplified body geometries, truncated
computational domains, boundary conditions and turbulence models. Numerical errors can
be further categorized into three parts:(Eça and Hoekstra, 2014)
(i) Discretization errors, including temporal and spatial error, arise from the neglecting of
high-order terms in the discretization process.
(ii) Iterative convergence errors, caused by the nonlinearity of the mathematical problems
and the truncation of the iterative process under a proper convergence criterion.
(iii) Round-off errors, due to the limit of machine precision. Round-off errors can largely be
neglected for double precision computations.
In this paper, dealing with modelling WEC motion response, we will focus on the numerical
errors (spatial and temporal discretization and iterative errors) and the modeling error will
be restricted to investigating the modelling error induced by truncating the computational
domain.
The V&V procedure used in this study is based on Richardson extrapolationEça et al.
(2010) and in order to obtain the numerical uncertainties a series of grids with different grid
density is used. Let hi denote the typical cell size of a grid and the subscript i = 1, 2, . . . , Ng
run over the Ng number of grids from the grid with highest density to the grid with lowest
density. The grid refinement ratio hi /h1 represents the ratio of cell size between grids with
different densities and is approximated as
r
hi N1
= 3 , (1)
h1 Ni
in which Ni is the total number of cells.
The error ε between the numerically obtained result φi using the ith grid and the exact
solution φ0 is defined as
ε = φi − φ0 = ahp , (2)

4
where p indicates the order of accuracy and a is a case specific constant. The true error is
approximated by the error obtained from the Richardson extrapolation δRE
φi − φ1
ε ≈ δRE = . (3)
(hi /h1 )p − 1

Most CFD models today are designed to be second-order accurate in space and time yielding
02
δRE = φi − φ0 = a02 h2 . (4)

Assuming the error to be a mixture of both first order and second order, it can be expanded
as a sum of two components
12
δRE = φi − φ0 = a11 h + a12 h2 (5)

The order of accuracy p and the constants a01 , a11 and a12 can be obtained by evaluating
the errors of the different grids and then performing a least square fit. If the estimated order of
accuracy is larger than 0, the error is assumed to be monotonically converged. RoacheRoache
(1997) suggested using safety factors to convert the numerical errors ε into numerical uncer-
tainties Uφ . This is achieved by simply multiplying the error with the safety factor FS

Uφ = FS |ε| . (6)

The safety factor depends on the order of convergence.Roache (1997) For solutions in the
asymptotic range (0.95 ≤ p ≤ 2.05, for a standard second-order scheme), the safety factor
should be set to 1.25. Otherwise the safety factor should be set to 3. The uncertainty is then
evaluated as
Uφ = 1.25δRE + US if p ∈ [0.95, 2.05] , (7)
where US is the standard deviation obtained from the least square fits. For monotone con-
vergence outside the asymptotic range the uncertainty is given by
12
Uφ = min(1.25δRE + US , 3δRE + US12 ) if p < 0.95, (8)
02
Uφ = max(1.25δRE + US , 3δRE + US02 ) if p > 2.05, (9)

in which US02 and US12 are the standard deviations derived from the least square fits. In case
the convergence is not monotone but oscillatory the uncertainty is estimated as

Uφ = 3δ∆M , (10)

where the error between the maximum and minimum is obtained as


max |φi − φj |
δ∆M = 1 ≤ i, j ≤ Ng . (11)
(hNg /h1 ) − 1

Using the obtained numerical uncertainty, we can estimate the model validation uncer-
tainty as q
Uval = Unum2 2
+ Umod 2 ,
+ Uexp (12)

5
where Umod is the uncertainties arising from modeling errors and Uexp is the experimental
uncertainty. By comparing the model uncertainty to the validation comparison error Ei =
φi −D, where D is the experimental value, we can see if the model is validated, i.e. |E| < Uval .
In this work we will estimate the numerical uncertainty of the non-dimensional response
amplitudes and restraining forces. The translational movement (surge is denoted by η1
and heave by η3 ), the rotational movement (pitch is denoted by η5 ) and forces are non-
dimensionalized as
η1 η3 η5 F
η̄1 = , η¯3 b = , η¯5 = , F̄ = . (13)
H0 H0 kH0 ρgH0 Sw
Here the overbar denotes non-dimensionalized value. The wave number is denoted by k, H0
is the target wave height, ρ is the water density, g the acceleration of gravity and Sw is the
water-plane area at the equilibrium position. Since the dominant component of the forces are
in heave direction, the forces are scaled by the Froude-Krylov force in the heave direction.

3 Test case: CorPower buoy


In this paper we simulate the CorPower WEC, see Fig. 1. CorPower Ocean ABCorPower
Ocean AB (2017) is a Swedish company which focuses on the development of a bottom-
reacting point-absorber with a novel passive control system, called WaveSpring. The buoy
is very lightweight, to be manufactured by composite materials, as the mass of the buoy
corresponds only to some 30% of the displaced mass at equilibrium. The system thus requires
a relatively large pretension in the mooring line.
The WaveSpring technology is made up of pressurized pneumatic cylinders that provide
a negative spring force when the buoy is not in equilibrium position. The net vertical force,
FW S , along the body-fixed central axis of the buoy is given as(Todalshaug et al., 2016)
z
FW S (z) = Nc Ac P (z) , (14)
l02 − z 2

where Nc is the number of pneumatic cylinders, Ac is the cross-section area of a cylinder, l0


is the initial length of the cylinder in its horizontal position, P (z) is the cylinder pressure and
z is the relative WEC position along the central axis. The negative spring force amplifies the
power capture and widens the response bandwidth. Through experimental tests performed in
the wave basin at Ecole Centrale de Nantes during November 2014 the WaveSpring technology
has been shown to increase the average power generation with a factor 3.(Todalshaug et al.,
2016)
The 1:16 scale experiments performed in Nantes will be used as validation cases for the
numerical study. The experiments are described in detail in Todalshaug et al.Todalshaug
et al. (2016) The experiment consists of a buoy, connected to a motor rig through a mooring
line running through two pulleys, see Figure 2. The submerged pulley is kept as stationary
as possible through the bottom mooring lines. The mooring line connected to the buoy is
pre-tensioned while the motor rig is used to approximate the PTO system of the system. The
general experimental parameters used are found in Table 1. The natural period of the buoy in
heave is estimated to approximately 0.8 s, while the surge natural period was found through
decay tests with the mooring to be roughly 4.0 s.

6
Figure 1: Illustration of CorPower WEC. Courtesy of CorPower Ocean AB (2017).

The motion of the buoy was recorded by Qualisys optical position sensors and two load cells
were installed on the mooring line, one by the buoy and one at the motor rig. Unfortunately
there are no available uncertainty estimates regarding the experimental tests, so we cannot
relate the numerical uncertainties to the experimental uncertainties in order to obtain the
validation uncertainty. However, we note that there are uncertainties related to the friction
in the pulleys and in the pretension measurements. There were also problems with waves
reflecting from the basin walls and the used moment of inertia has been estimated based
on approximate gyration radius data. We should keep in mind that these tests were not
performed to give high-resolution validation data, but to illustrate the performance of the
WaveSpring technology.

Table 1: Experimental set-up.


Item Value
Buoy diameter 0.525 m
Buoy height 1.125 m
Buoy mass 19.75 kg
Equilibrium displacement 0.0708 m3
Equilibrium waterline 0.98 m
Moment of inertia 2.38 kgm2
Water depth 5.00 m
Water depth at mooring point 3.09 m
Rope length 15.26 m
Rope and sheave mass 1.85 kg
Mooring line spring stiffness 202 N/m
Pretension force 500 N

7
Figure 2: The model setup (left) and the model in the wave basin (right). Courtesy of
CorPower Ocean AB (2017).

We will use a regular wave case with a wave height of 0.156 m and a wave period of 2.25 s.
This is an almost linear incident wave with a wave steepness (H/L) of 0.02 (H being wave
height and L being the wavelength). Two set-ups are investigated:

• linear damper PTO system, referred to in the text as LD. The damping coefficient is
set to 628 kg/s which is the optimum value for unconstrained heave.

• WaveSpring+linear damper PTO system, referred to in the text as WS. The WaveSpring
system is made up of three cylinders with A0 = 5.82 × 10−4 m2 , l0 = 0.25 m and initial
volume V0 = 1.59 × 10−4 m3 . The damping coefficient used is set to 238 kg/s.

4 Numerical settings
4.1 VOF-RANS solver
We use here the OpenFOAM-4.0 two-phase VOF-RANS solver with moving mesh support,
interDyMFoam. This is a cell-centred finite volume segregated solver using a SIMPLE/PISO
(PIMPLE) approach. The two-phase interface is kept sharp by a pseudo-force compression
technique.Rusche (2002) Wave generation and absorption is handled by relaxation zones as
implemented in the waves2Foam package(Jacobsen et al., 2012) and the incoming waves are

8
given by 5th order Stokes theory. The shape of the relaxation function was calibrated be-
forehand using pure wave propagation. We note that the relaxation factor has been weighted
to the wave period to avoid any influence of the time step on the relaxation procedure. The
turbulence model used is the standard k − ω SST model, which is often used in marine ap-
plications. The boundary conditions are: total pressure condition to the atmosphere; no-slip
at the body; slip at the tank walls and pressure boundary at the downstream boundary. The
numerical schemes used are second-order van Leer scheme for convection terms, second-order
central differences for diffusion terms, while the turbulence equations are solved using the
first-order upwind method. The time-stepping is carried out using a blended second-order
Crank-Nicholson/first-order backward Euler scheme.
The only bespoke part of the numerical model used in this study is the implementation
of the pre-tension or the mooring line, simplified as a linear spring, and the WaveSpring
technology. The pre-tensioned spring and the WaveSpring force, as computed from eq. (14),
have been implemented as specialized restraints inside the 6 DoF body-motion solver.

4.2 Grids
The computational meshes used have been generated with the snappyHexMesh mesh generator.
snappyHexMesh builds a hexahedral dominant oct-tree mesh with extruded boundary layers
from a background hexahedral mesh and a triangulated surface file of the body geometry. The
meshes used in this study are roughly 48 m (6 wavelengths) long, 9 m (1 wavelength) wide
and 8 m (5 m still water depth) high. The relaxation zones for wave generation/absorption
are 2 wavelengths each. See Fig. 3a. We use four levels of grid refinement with the finest
resolution in boxes around the buoy and around the free surface interface. The background
mesh is cubic, giving a cell aspect ratio of unity. Figure 3b shows the grid refinement in
transverse and longitudinal center planes.
To test the effect of the grid density and analyze the spatial discretization error, four
grids with different cell numbers and resolutions are created. In each grid, the background
cell is scaled so that the grid topology is maintained. Figure 4 visualizes the mesh close to
the floating body. The properties of the four grids are listed in Table 2. For the baseline
mesh (G2) made up of 10M cells, the cell size ensures 30 cells within the wave height along
the vertical direction, which is sufficient to describe the interface. By adding five layers, the
near-wall cell has a y + value of roughly 50, and thus wall functions are used. Please note that
for oscillating flow cases preferably y + < 1 and no wall functions should be used, but in order
to keep the computational effort down we resort to using wall functions.

Table 2: Computational meshes used.


G4 G3 G2 G1
Number of cells 1.1E6 3.1E6 10.9E6 19.9E6
Refinement ratio 2.65 1.86 1.22 1.00
Base ∆x, ∆y [m] 0.48 0.33 0.215 0.175
Base ∆z [m] 0.33 0.25 0.167 0.136
Cells per wave height 15 20 30 37
Max y + 185 105 50 75

9
(a)

(b)

Figure 3: Computational meshes. (a) Geometry of the mesh and (b) grid refinement along
center plane.

10
(a) (b)

(c) (d)

Figure 4: Mesh resolution near the buoy for the three meshes: (a) 1 M cells, (b) 3 M cells,
(c) 10 M cells and (d) 20 M cells.

11
(a) (b) (c)

(d) (e) (f)

Figure 5: WEC motion in waves for the LD case at times (a) t=40.0 s, (b) t=41.5 s,
(c) t=43.0 s, (d) t=44.5 s, (e) t=44.5 s and (f) t=44.5 s.

12
(a) (b) (c)

(d) (e) (f)

Figure 6: WEC motion in waves for the WS case at times (a) t=30.25 s, (b) t=31.25 s,
(c) t=32.25 s, (d) t=33.25 s, (e) t=34.0 s and (f) t=35.0 s.

13
5 Validation
We initially use grid G2 (10M cells) to compare the numerical results using the linear damper
as well as the WaveSpring case to the experimental tank test data.
Prior to the simulations with the floating buoy, simulations covering only wave propaga-
tion were performed in order to calibrate the relaxation zones and to ensure that no waves
were reflected from the downstream boundary. Using the G2 mesh the waves were found to
loose roughly 1% in wave height when propagating through the computational domain. The
diffusion of the free surface thus has a negligible influence on the motion response.
Figures 5 and 6 show snapshots of the buoy in waves. The LD case moves smoothly and
there are hardly any diffracted waves. This is in sharp contrast to the WS case where the buoy
motion is amplified and the buoy almost gets over-topped even for this very linear incident
wave. The figures show implicitly that for the LD case there is no concern that the free surface
elevation will be outside the refinement region, but for the WS case this can be the case and
care should be taken to ensure that the free surface is always in the refinement region.
The enhancement of the nonlinear response of the WS is even more clearly seen in Figs. 7
and 8 showing the numerical and experimental results of motion and forces for the LD and
WS cases. In the figures, the restraint force FR is the sum of the mooring force and the PTO
force. The LD case shows clear harmonic behaviour for all four variables, the responses are
largely linear as can be seen by the lack of higher harmonics in the time series, and that the
positive and negative amplitudes around the equilibrium position are roughly equal. We note
the large surge displacement even for this linear wave excitation with a period well below the
surge natural frequency. This is caused by the small inertia of the CorPower buoy. The heave
response is slightly over predicted while the pitch is somewhat under predicted. These two
modes were found to be sensitive to the value of the pretension, which is a bit uncertain from
the experimental tests. In addition, there are uncertainties associated with the used moment
of inertia which also affect the pitch response.
For the WS case there is a significant increase in response – there is a threefold increase
in heave and a twofold increase in surge and pitch. The restraining/WaveSpring forces and
heave response have the same major period as the wave period, but with small but clearly
evident higher harmonics present. For the surge and pitch response the major period now is
2T which is close to the surge natural period. Further, the response is now non-symmetrical
with higher positive than negative amplitudes.
The simulations compare well to the experimental data for the LD case for all variables.
Table 3 presents the comparison error for all four grids. We see that the heave is over predicted
while the other quantities are under predicted. For the G2 baseline mesh the comparison errors
are roughly 5%.
The simulations compare reasonable well to the experimental data also for the WS case.

Table 3: Comparison error for the LD case.


Grid G4 G3 G2 G1
Eη̄1 -11.6% -6.5% -3.1% -1.8%
Eη̄3 -3.7% 1.2% 4.4% 6.0%
Eη̄5 -15.3% -9.8% -5.9% -4.7%
EF¯R -13.5% -8.9% -6.2% -4.4%

14
(a) (b)
0.05
Experimental Experimental
0.2 Numerical Numerical
0.025
0.1
η1 [m]

η3 [m]
0
0

-0.1 -0.025

-0.2 -0.05
0 10 20 30 40 0 10 20 30 40
Time [s] Time [s]
(c) (d)
4 -400
Experimental Experimental
3 Numerical Numerical

-450
η5 [degrees]

2
FR [N]

1
0 -500

-1
-2 -550

-3
0 10 20 30 40 0 10 20 30 40
Time [s] Time [s]

Figure 7: Motion response and force with linear damper PTO: (a) surge; (b) heave; (c) pitch
and (d) restraint force.

15
(a) (b)
0.6 0.2
Experimental Experimental
Numerical Numerical
0.4 0.1

[m]
[m]

0.2
0
1

3
0
-0.1

-0.2
-0.2
0 10 20 30 40 50 60 70 0 10 20 30 40 50 60 70
Time [s] Time [s]
(c) (d)
10 -350
Experimental Experimental
Numerical Numerical
-400
[degrees]

5 -450
FR [N]

-500
0
5

-550

-600
-5
-650
0 10 20 30 40 50 60 70 0 10 20 30 40 50 60 70
Time [s] Time [s]

(e)
300
Experimental
Numerical
200

100
FWS [N]

-100

-200

0 10 20 30 40 50 60 70
Time [s]

Figure 8: Motion response and force with the WaveSpring system: (a) surge; (b) heave; (c)
pitch; (d) restraint force and (e) WaveSpring force.

16
(a) (b)

Figure 9: Illustration of vortical structures near the buoy visualised by the Q = 1 iso-surface
(a) LD case and (b) WS case.

We note that the experimental test was stopped after some 50 s in order to avoid pollution
by reflected waves. Thus the experimental run never reached a steady periodic motion in
surge and pitch, the initial start-up transient is apparent in the time series. In the numerical
simulation the incident waves were not stopped and this is the reason why the numerical and
experimental results differ for t > 50 s. Generally the motion responses are slightly under
predicted, higher resolution is needed to fully capture the nonlinear response, but the results
support that the CFD simulation reflects the real physical characteristics of the WaveSpring
system.
The WaveSpring technology greatly affects the response and also the viscous drag. The
LD case has a Keulegan-Carpenter (KC) value of 1.8, giving that the flow is inertia dominated,
but the viscous effect should not be neglected. The WS case has a KC value of 2.9, indicating
that there exist considerable viscous effects which can lead to vortex shedding in horizontal
direction. This is illustrated in Fig. 9 which shows the vortical structures appearing near the
buoy for the LD and WS cases. Although the structures are weak in general, it is clear that
the WS case has much more vortical structures. Clearly neither case creates any downstream
wake, but for the WS case the buoy might move through its own wake.

6 Solution verification: LD case


6.1 Spatial discretization error
The motion amplitude in surge, heave and pitch and the restraint forces given by the four
grids are extracted from the last five periods, averaged and then non-dimensionalized. Fig. 10
show the Richardson extrapolations of the convergences for the LD case. It can be seen that
all variables converge monotonically and are in the “asymptotic range” with fitted p:s ranging
from 1.21 to 1.49. The estimated uncertainties are presented in Table 4. For the baseline
mesh G2 we have generally spatial uncertainties less than 5% with the exception of surge
that has an uncertainty larger than 10%. Here we note that the CorPower buoy have very

17
(a) (b)
2.1 0.55

η¯3
η¯1

1.9 0.5

1.8 Computed Computed


Fitted curve (p = 1.42) Fitted curve (p = 1.29)
Fitted curve (p = 2) Fitted curve (p = 2)
1.7 0.45
0 1 2 3 0 1 2 3
hi /h1 hi /h1

(c) (d)

0.75
0.46

0.44
0.7
F¯R
η¯5

0.42
0.65
Computed 0.4 Computed
Fitted curve (p = 1.49) Fitted curve (p = 1.21)
Fitted curve (p = 2) Fitted curve (p = 2)
0.6 0.38
0 1 2 3 0 1 2 3
hi /h1 hi /h1
Figure 10: Uncertainty estimation of spatial discretization for the LD case: (a) surge response,
(b) heave response, (c) pitch response, (d) restraint force.

little inertia and small horizontal restraints yielding a large surge response and making surge
especially troublesome to resolve accurately.
We note that the standard way of estimating ”grid independent” solutions (|φi −φG1 |/φG1 )
for this case shows errors in the same magnitude as the Richardson extrapolated uncertainties.
This is due to the wide range of refinement ratios used. Often the smaller refinement ranges
are used and then the standard method often shows very small error, often factions of a
percent. It is worth noticing that the standard method shows the lowest error for surge while
the solution verification highlights surge as the most troublesome variable to resolve.

6.2 Iterative error


As mentioned previously, a segregated solution algorithm is used in the numerical model. In
the segregated approach the number of iterations performed at each time step is an important

18
Table 4: Spatial discretization estimation for the LD case.
Grid G4 G3 G2 G1
η̄1 1.816 1.920 1.984 2.018
|η̄1 − η̄1G1 |/η̄1G1 10.0% 4.8% 1.7% –
Uη̄1 33.9% 20.9% 11.6% 8.8%
η̄3 0.476 0.500 0.517 0.524
|η̄3 − η̄3G1 |/η̄3G1 9.2% 4.5% 1.3% –
Uη̄3 8.5% 5.5% 3.3% 2.6%
η̄5 0.648 0.690 0.715 0.729
|η̄5 − η̄5G1 |/η̄5G1 11.1% 5.3% 1.6% –
Uη̄5 13.3% 8.0% 4.3% 3.2%
F¯R 0.401 0.422 0.436 0.443
|F̄R − F̄RG1 |/F̄RG1 9.5% 4.7% 1.9% –
UF̄R 7.6% 5.2% 3.2% 2.6%

-3
Log ( presidual )

-4

-5
2 pimple loops
3 pimple loops
-6 4 pimple loops
5 pimple loops
10 pimple loops
-7 15 pimple loops
20 21 22 23 24 25
Time [s]

Figure 11: Initial pressure residual using different number of iterations in the PISO/SIMPLE
algorithm for the LD case.

parameter as the residual is typically decreasing with increasing iteration numbers. In order to
investigate the iterative error we use the baseline G2 mesh and apply Niter = [2, 3, 4, 5, 10, 15]
outer iterations.
The initial pressure residual at every time step is shown in Fig. 11. As expected the
pressure residual decreases with increasing iteration numbers. However, the effect flattens
out after 10 iterations and there is little to gain to use more than 10 iterations (for the used
time step). Thus we assume that the solution using 15 iterations will be fully converged and
the relative iterative errors are calculated as

φi − φ15
Uφ = . (15)
φ15
The resulting iterative errors are presented in Table 5. The iterative error is not negligible
and for iterations less or equal to 5 the iterative error is smaller but of the same order of
magnitude as the spatial discretization error. Using 10 iterations makes the iterative error an

19
Table 5: Iterative errors estimation for the LD case.
Niter 2 3 4 5 10
Uη̄1 0.64% 0.80% 0.84% 0.73% 0.16%
Uη̄3 2.00% 1.35% 1.03% 0.81% 0.38%
Uη̄5 1.56% 1.20% 1.10% 0.90% 0.17%
UF̄R 0.29% 0.09% 0.08% 0.16% 0.14%

order of magnitude smaller than the spatial discretization uncertainty. In addition, one can
also observe considerable accumulative errors and phase difference after several periods when
less iterations are applied.
It should be noted that for unsteady simulations like the present, iterative errors and
temporal resolutions are linked. Typically a smaller time step would require less iterations for
the same pressure residual. We also note that the specific case of the CorPower buoy might
be especially sensitive to iterative errors. The CorPower buoy has a very small inertia and
large pretension, giving that the restraint forces are the dominating forces in the six-degrees-
of-freedom body motion solver. It is believed that this might call for additional iterations for
the body motion solver to converge.

6.3 Temporal discretization error


As the Crank-Nicholson time stepping scheme is used in the computations, second order
convergence in time is expected. The simulations are carried out on the G2 base mesh with
three different time steps, ∆t = [0.005, 0.01, 0.02] s denoted by T1, T2 and T3. As performed
for the spatial discretization error the motion amplitudes of the last five periods are used for
the uncertainty analysis.
The fitted p-values are in the range 0.96 to 1.62, showing monotone convergence in the
asymptotic range. We have two clusters of values; surge/pitch with mixed convergence at
roughly 1.5 and heave/force with approximately first order convergence.
The estimated uncertainties associated with the temporal discretization are presented in
Table 6. It can be seen that the uncertainties associated with time are in general smaller than
for the spatial discretization and are typically a couple of percent for ∆t = 0.01 s and below
one percent for a time step of 0.005 s.

6.4 Domain error


The domain error possibly introduced by using a too narrow numerical wave tank is investi-
gated by testing a wave tank with a width of 16.4 m, corresponding to two wave lengths and
twice the width of the original wave tank. The obtained values of η̄1 and η̄3 were virtually
identical. The errors in pitch (η̄5 ) and restraint force (F̄R ) are also very small, in the order of
0.1%.

7 Solution verification: WS case


As identified from the LD case the spatial discretization error is the dominant part. We thus
restrict the solution verification of the WS case to cover only the spatial uncertainties.

20
Table 6: Temporal error estimation for the LD case.
∆t T3 T2 T1
η̄1 1.948 1.984 1.995
|η̄1 − η̄1T 1 |/η̄1T 1 2.3% 0.6% —
Uη̄1 6.5% 2.1% 0.6%
η̄3 0.509 0.517 0.521
|η̄3 − η̄3T 1 |/η̄3T 1 2.3% 0.8% —
Ux̄ 2.0% 1.0% 0.5%
η̄5 0.704 0.715 0.719
|η̄5 − η̄5T 1 |/η̄5T 1 2.1% 0.8% —
Uz̄ 2.2% 0.8% 0.2%
F̄R 0.424 0.436 0.441
|F̄R − F̄RT 1 |/F̄RT 1 3.9% 1.2% —
UF̄R 6.9% 3.1% 1.4%

7.1 Spatial discretization error


We compare the computed maximum amplitudes after approximate 35 s using the four differ-
ent meshes. It turns out that it is quite hard to get good convergence for this case: only the
restraining and WaveSpring forces show monotone convergence within the asymptotic range.
Consequently only these two variables have acceptable uncertainty, below 5%, as seen from
Table 7. For surge and heave we have monotonic convergence but well outside the asymptotic
range, while the pitch motion shows oscillatory convergence. The resulting uncertainties are
very large, up to almost 40% for mesh G2. There is a striking difference compared to using the
standard method for grid independence studies. The standard method gives that the WS case
is as well resolved as the LD case. The Richardson extrapolation method clearly illustrates
that this is not the case and that the WS case is computationally much more demanding,
requiring much more resolution in order to get acceptable numerical uncertainties.

Table 7: Spatial discretization estimation for the WS case.


Grid G4 G3 G2 G1
η̄1 2.488 2.529 2.690 2.640
|η̄1 − η̄1G1 |/η̄1G1 5.7% 4.2% 1.9% —
Uη̄1 58.9% 48.7% 38.8% 35.0%
η̄3 0.722 0.779 0.805 0.816
|η̄3 − η̄3G1 |/η̄3G1 11.5% 4.5% 1.3% —
Uη̄3 37.8% 35.6% 28.2% 24.4%
η̄5 0.959 0.987 1.061 0.958
|η̄5 − η̄5G1 |/η̄5G1 0.0% 3.0% 10.7% —
Uη̄5 10.7% 10.7% 10.7% 10.7%
F̄R 1.279 1.267 1.264 1.258
|F̄R − F̄RG1 |/F̄RG1 1.7% 0.7% 0.5% —
UF̄R 3.2% 1.9% 1.1% 0.8%
F̄W S 0.517 0.541 0.550 0.555
G1 G1
|F̄W S − F̄W S |/F̄W S 6.7% 2.5% 0.7% —
UF̄W S 13.3% 6.8% 3.1% 2.1%

21
8 Discussion and conclusions
We have presented CFD simulations of a point-absorber wave energy device with and without
passive phase control, developed by CorPower Ocean ABCorPower Ocean AB (2017). The
simulations showed an acceptable fit to experimental data from a 1:16 scale test performed
at Nantes.(Todalshaug et al., 2016)
More importantly we have presented a rigorous verification and validation (V&V) study
focusing on the solution verification stage. This stage has hitherto been overlooked in CFD
simulations of wave energy devices. Previous studies have focused on validation studies com-
paring to experimental data and possibly making sure the solutions are “grid independent”.
The solution verification procedure followed the technique of Eça et al.Eça et al. (2010)
using Richardson extrapolation and following Roache’sRoache (1997) approach to assign safety
factors based on convergence rates. We found that for the non-controlled LD case, using the
baseline 10M cell mesh with 30 cells per wave height and an aspect ratio of unity, the validation
uncertainty Uval can be below 5% if a time-step of 0.005 s and 5 iteration loops are employed,
giving an acceptable uncertainty for practical applications. The exception is in surge where
the uncertainty is above 10%. The CorPower buoy differs from traditional buoys in being
very lightweight with low inertia which could be a likely reason for the larger uncertainty in
surge.
No information regarding the experimental uncertainties are given and for the sake of
simplicity we assume them to be zero. The validation comparison error E for the LD case
was computed using 0.01 s time-step with 10 iterations. Although |E| < Uval for all quantities
except pitch (and also pitch being very close, 5.9% vs 5.3%) the model is not validated as we
are not inside the 95% confidence interval except for heave. The model should be refined as
mentioned above.
With regard to the WS case the numerical model was capable of replicating the key features
of the WaveSpring behaviour, but clearly the quality of the numerical solution needs to be
improved before any firm conclusions can be drawn. Indeed the spatial discretization error
is very large, up to almost 40% in surge, caused by convergence rates outside the asymptotic
range or even oscillatory convergence.
The results from the study highlights:
• that standard “grid independence” studies can grossly underestimate the uncertainties
involved in the CFD. This is most obvious in the surge response where the grid in-
dependence approach gives just a percent error, the same order of magnitude as the
other responses, while the solution verification underlines the larger error associated
with surge for the CorPower buoy;
• although the spatial discretization errors dominate, we found that all numerical errors
(except round-off) need to be included in the solution verification step;
• the difficulty in getting reliable results for highly nonlinear motion response. The imple-
mentation of passive phase control caused the convergence to be outside the asymptotic
range, or even becoming oscillatory, giving unacceptable large uncertainties. Clearly, for
highly nonlinear waves and responses much finer meshes in time and space are required.
The need for high fidelity experimental data for CFD validation purposes has been put
forward previusly.Wolgamot and Fitzgerald (2015) This study adds to that request the need

22
to also have reliably experimental uncertainty estimates in order to facilitate model validation
through V&V techniques.
Finally, the need to estimate the numerical uncertainties and errors is vital in order to
enhance the trust in numerical simulations, in order to perform more of the optimization as
early in the design process as possible, and to provide a more economical development path.

Acknowledgements. The experimental data was gratefully provided by CorPower Ocean


AB. We especially thank Dr. Jørgen Hals Todalshaug at CorPower for the invaluable assis-
tance with regard to the WaveSpring implementation. Support for this work was given by the
Swedish Energy Agency under grant no. P40428-1. The simulations were made on resources
from Chalmers Centre for Computational Science and Engineering, provided by the Swedish
National Infrastructure for Computing.

References
AIAA (1998). Guide for verification and validation of computational fluid dynamics simula-
tions. AIAA Standards, (AIAA G-077-1998).

Babarit, A. and Delhommeau, G. (2015). Theoretical and numerical aspects of the open
source BEM solver NEMOH. In: Proceedings of the 11th European Wave and Tidal Energy
Conference. 6-11 September Nantes, France.

Bhinder, M.A., Babarit, A., Gentaz, L. and Ferrant, P. (2015). Potential time domain model
with viscous correction and CFD analysis of a generic surging floating wave energy con-
verter. International Journal of Marine Energy, 10, 70–96.

Bull, D., Ochs, M., Laird, D., Boren, B. and Jepsen, R. (2013). Technological cost-reduction
pathways for point absorber wave energy converters in the marine hydrokinetic environment.
Technical report, Sandia National Laboratories.

Chen, W., Dolguntseva, I., Savin, A., Zhang, Y., Li, W., Svensson, O. and Leijon, M. (2017).
Numerical modelling of a point-absorbing wave energy converter inirregular and extreme
waves. Applied Ocean Research, 63, 90–105.

CorPower Ocean AB (2017). http://www.corpowerocean.com/.

Davidson, J., Cathelain, M., Guillemet, L., Huec, T.L. and Ringwood, J.V. (2015). Implemen-
tation of an OpenFOAM numerical wave tank for wave energy experiments. In: Proceedings
of the 11th European Wave and Tidal Energy Conference. 6-11 September Nantes, France.

Eça, L. and Hoekstra, M. (2014). A procedure for the estimation of the numerical uncertainty
of CFD calculations based on grid refinement studies. Journal of Computational Physics,
262, 104 – 130. ISSN 0021-9991.

Eça, L., Vaz, G. and Hoekstra, M. (2010). A verification and validation exercise for the
flow over a backward facing step. In: Fifth European Conference on Computational Fluid
Dynamics. 14-17 June Lisbon, Portugal.

23
Eskilsson, C., Palm, J., Kofoed, J. and Friis-Madsen, E. (2015). CFD study of the over-
topping discharge of the Wave Dragon wave energy converter. In: Proceedings of the 1st
International Conference on Renewable Energies Offshore, 287–294. 24-26 November 2014,
Lisbon, Portugal.

Gahree, B., Eskilsson, C., Bensow, R. and Vaz, G. (2016). Simulation of cavitation on a
horizontal axis tidal turbine. In: Proceedings of the 26th International Symposium on
Ocean and Polar Engineering. 26 June 1 July, Rhodes, Greece.

Giorgi, G. and Ringwood, J. (2016). Implementation of latching control in a numerical wave


tank with regular waves. Journal of Ocean Engineering and Marine Energy, 2(2), 221–226.

Guo, B.J., Deng, G.B. and Steen, S. (2013). Verification and validation of numerical calcula-
tion of ship resistance and flow field of a large tanker. Ships and Offshore Structures, 8(1),
3–14.

Henry, A., Rafiee, A., Schmitt, P., Dias, F. and Whittaker, T. (2014). The characteristics of
wave impacts on an oscillating wave surge converter. Journal of Ocean and Wind Energy,
1(2), 101–110.

Jacobsen, N.G., Fuhrman, D.R. and Fredsøe, J. (2012). A wave generation toolbox for the
open-source CFD library: OpenFoam . R International Journal for Numerical Methods in
Fluids, 70(9), 1073–1088. ISSN 1097-0363.

Kim, J.W., Jang, H., Baquet, A., OSullivan, J., Lee, S., Kim, B., Read, A. and Jasak, H.
(2016). Technical and economical readiness review of CFD-based numerical wave tank
for offshore floater design. In: Proceedings of Offshore Technology Conference. 2-5 May
Houston, USA.

Make, M. and Vaz, G. (2015). Analyzing scaling effects on offshore wind turbines using CFD.
Renewable Energy, 83, 1326–1340.

OpenFOAM (2017). http://www.openfoam.com.

Palm, J., Eskilsson, C., Moara Paredes, G. and Bergdahl, L. (2016). Coupled mooring analysis
for floating wave energy converters using CFD: Formulation and validation. International
Journal of Marine Energy, 16, 83–99.

Pereira, F., Eça, L. and Vaz, G. (2017). Verification and validation exercises for the flow
around the KVLCC2 tanker at model and full-scale Reynolds numbers. Ocean Engineering,
129, 133–148.

Rijpkema, D. and Vaz, G. (2011). Viscous flow computations on propulsors: Verification,


validation and scale effects. In: Proceedings of the RINA Marine CFD Conference. 22 - 23
March, London, UK.

Roache, P.J. (1997). Quantification of uncertainty in computational fluid dynamics. Annual


Review of Fluid Mechanics, 29, 123–160.

Rusche, H. (2002). Computational Fluid Dynamics of Dispersed Two-Phase Flows at High


Phase Fractions. Ph.D. thesis, Imperial College.

24
Schmitt, P. and Elsaesser, B. (2015). On the use of OpenFOAM to model oscillating wave
surge converters. Ocean Engineering, 108, 98–104.

Simonsen, C.D. and Stern, F. (2003). Verification and validation of RANS maneuvering sim-
ulation of Esso Osaka: effects of drift and rudder angle on forces and moments. Computers
& Fluids, 32, 1325–1356.

Stern, F., Wilson, R.V., Coleman, H.W. and Paterson, E.G. (2001). Comprehensive approach
to verification and validation of CFD simulations – Part 1: Methodology and procedures.
Journal of Fluids Engineering, 123, 793–802.

Tiron, R., Mallon, F., Dias, F. and Reynaud, E. (2015). The challenging life of wave energy
devices at sea: A few points to consider. Renewable and Sustainable Energy Reviews, 43,
1263–1272.

Todalshaug, J.H., Ásgeirsson, G.S., Hjálmarsson, E., Maillet, J., Möller, P., Pires, P.,
Guérinel, M. and Lopes, M. (2016). Tank testing of an inherently phase-controlled wave
energy converter. International Journal of Marine Energy, 15, 68–84.

Vyzikas, T., Deshoulieres, S., Giroux, O., Barton, M. and Greaves, D. (2017). Numerical
study of fixed oscillating water column with RANS-type two-phase CFD model. Renewable
Energy, 102, 294–305.

WAMIT Inc. (2016). Wamit user manual version 7.2.

Weller., H., Tabor, G., Jasak, H. and Fureby, C. (1998). A tensorial approach to computational
continuum mechanics using object-oriented techniques. Computers In Physics, 12(6), 620–
631.

Wilson, R.V., Stern, F., Coleman, H.W. and Paterson, E.G. (2001). Comprehensive approach
to verification and validation of CFD simulations – Part 2: Application for Rans simulation
of a cargo/container ship. Journal of Fluids Engineering, 123, 803–810.

Wolgamot, H. and Fitzgerald, C. (2015). Nonlinear hydrodynamic and real fluid effects on
wave energy converters. Proceedings of the Institution of Mechanical Engineers, Part A:
Journal of Power and Energy, 229(7), 772794.

Yu, Y., Lawson, M., Ruehl, K. and Michelen, C. (2014). Development and demonstration of
the WEC-Sim wave energy converter simulation tool. In: Proceedings of the 2nd Marine
Energy Technology Symposium. 15-17 April, Seattle, US.

Yu, Y.H. and Li, Y. (2013). Reynolds-averaged NavierStokes simulation of the heave perfor-
mance of a two-body floating-point absorber wave energy system. Computers & Fluids, 73,
104–114.

25

View publication stats

You might also like