You are on page 1of 10

Ocean Engineering 111 (2016) 483492

Contents lists available at ScienceDirect

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

Numerical and experimental analysis of the power output of a point


absorber wave energy converter in irregular waves
M.T. Rahmati a,b,n, G.A. Aggidis b
a

Department of Mechanical, Aerospace and Civil Engineering, Brunel University London, Uxbridge, UB8 3PH, UK
Lancaster University, Engineering Department, Renewable Energy Group, Faculty of Science and Technology, Bailrigg, Lancaster LA1 4YR, UK

art ic l e i nf o

a b s t r a c t

Article history:
Received 11 June 2015
Received in revised form
4 November 2015
Accepted 15 November 2015

This paper examines the optimum power output of a pitching-surge point absorber wave energy converter in irregular wave climates. A mathematical model based on frequency domain is used as the rst
step to estimate the hydrodynamic parameters of the device and its potential power output in realistic
sea waves. The numerical results predict that the point absorber energy converter has the potential to
absorb more energy than what is contained in its own geometrical width. The optimum power of the
device is then obtained from wave tank experiments in irregular wave climates. The comparison of
numerical and experimental results demonstrates that the frequency domain method based on linear
theory will lead to an overestimation of the energy absorption for this device. The frequency domain
method provides an upper estimate for wave energy absorption due to the non-linear, viscous effects and
constrained amplitude of device oscillation. However, comparison of the performance of the device with
other point absorber wave energy converters shows that this wave energy converter is one of the most
efcient in terms of absorbing wave energy.
& 2015 Elsevier Ltd. All rights reserved.

Keywords:
Wave energy converter
Point absorber
Irregular waves
Capture width ratio

1. Introduction
The oil crisis of 1973 led to a period of high energy prices which
intensied interest in renewable energies such as large-scale
energy production from the sea waves. One of the earliest works
in the UK was a paper published in Nature by Salter (1974) which
brought the feasibility of wave energy conversion to the attention
of the international scientic community. Research into wave
energy started at Lancaster University in the mid-1970s, mainly
concentrating on the invention and development of wave energy
converters (WECs) at model scale, along with generic work on the
theory and systematic design of WECs. This has led to the great
understanding of the theory and systematic prototype design of
WECs (French and Bracewell, 1985; Chaplin and Folley, 1998). One
of the earliest WECs was Lancaster exible bag invented by
Michael French at Lancaster University which was subsequently
developed by Wave power limited in the UK (Platts, 1981). By the
early 1980s, the period of high energy prices was ended and
consequently the economic and political climate in which WECs
were being developed had changed drastically from that in the
mid to late 1970s. Nevertheless, research devoted to wave energy
converters for the production of electricity continued but a greater

emphasis was placed upon the economic analysis of wave energy


extraction and efciency of WECs. As a result, hundreds of WECs
with various technologies to efciently harvest this energy have
been proposed during the last decades, see for example Antonio
and Falcao (2010) and Lopez et al., (2013). A few of those technologies has reached the sea trial test status. Lindroth and Leijon
(2011) have provided a review of various WECs that have made
ocean trials, together with the details of the data measurement
during the trials. It was demonstrated that, very few of those
devices have reported detailed measurements of the performance
of the devices which could be due to the great difculties in
working with experiments in a harsh environment.
Babarit (2015) has provided a database for the hydrodynamic
performance of various WECs based on the collection and analysis
of data available in the literature. The study focused on the capture
width ratio of the different technologies. The terms capture width
and capture width ratio instead of efciency is normally used for
evaluating WECs in wave tanks. Capture width is a parameter
dened as the width of the wave front (assuming uni-directional
waves) that contains the same amount of power as that absorbed
by the WEC, see Price et al. (2009). Capture width (Cw) is dened
by:
Cw

Corresponding author.
E-mail address: mt.rahmati@brunel.ac.uk (M.T. Rahmati).

http://dx.doi.org/10.1016/j.oceaneng.2015.11.011
0029-8018/& 2015 Elsevier Ltd. All rights reserved.

Pd
Pw

where Pd is the device power and Pw is the wave power per meter.

484

M.T. Rahmati, G.A. Aggidis / Ocean Engineering 111 (2016) 483492

q
l
m
m0
F
S.W.L
WEC
CW
cw

Nomenclature
a55
b55
c55
F1
Hm0
hcm
IQ
G
P
p
Q

added moment of inertia, kg m2


radiation damping coefcient, N m s
restoring moment coefcient, N m/rad
excitation force, N
signicant wave height, m
the height of crest above mean water level, m
dry moment of inertia, kg m2
center of gravity
center of force on the collector
distance between P and still water level, m
pivot point

Im

The capture width ratio, cw, is a better parameter as it is a nondimensional parameter which best reects the hydrodynamic
performance of various WECs. Capture width ratio reects the
fraction of wave power owing through the device that is absorbed by the device and dened by:
cw

Cw
W

where W is a characteristic parameter usually considered as the


width of a WEC.
WEC technologies can be broadly classied based on their
operational principle to three groups: oscillating water columns
(OWCs), overtopping devices and oscillating bodies. Oscillating
bodies WECs are usually referred to as Oscillating Wave Surge
Converters (OWSCs). Since OWSCs covers a broad range of devices,
this category can be divided further into oating (heaving device)
and bottom xed devices (Babarit, 2015). Power can be absorbed
from heave, surge, pitch, or some combination of these modes. The
majority of existing point absorber devices operates in heave
mode alone. However, the potential power capture in OWSC which
operate in surge and pitch is twice that of the device working in
heave mode, as shown by Falnes (2002). A typical OWSC is the
Frog, of which several offshore point-absorber versions have been
developed at Lancaster University, UK. The Frog consists of a large
buoyant paddle with an integral ballasted handle hanging below it
(McCabe et al., 2003, 2006). The waves act on the blade of the
paddle and the ballast beneath provides the necessary reaction.
When the wave energy converter is pitching, power is extracted by
partially resisting the sliding of a power-take-off mass, which
moves in guides above sea level. It is interesting to note that the
rst two categories (i.e. OWCs and overtopping devices) have been
studied for more than 30 years, whereas most xed OWSCs have
only started developing over the last decade. However, as
demonstrated by Babarit (2015), the most efcient categories of
WECs, in terms of absorbing wave energy, are xed OWSCs.
OWSCs can be further classied according to their horizontal
size. If the extension is comparable to or larger than the typical
wavelength, the WEC is called a line absorber. On the contrary, if
the size is very small compared to the typical wavelength, the WEC
is called a point absorber (Budal and Falnes, 1975; Evans, 1980).
Whilst some designs are quite large, there is much interest in
point absorber devices which are relatively small compared to the
most common sea wavelengths, and have the potential for more
efcient power conversion, in terms of output per unit volume.
This is because during the initial research into the hydrodynamics
of wave energy converters it was discovered that bodies with
dimensions small relative to the incident wavelengths could act in
a similar way to an aerial and capture energy from a width greater

distance between the pivot and SWL, m


distance between G and SWL, m
mass, kg
total variance of the wave system
force acting normal to the arm, N
still water level
wave energy converter
capture width
capture width ratio
wave amplitude, m
width
pitch motion, radian
immersion (mm)

than their own width. This became known as the point-absorber


theory (Evans, 1980; Falnes and Budal, 1978). At an early stage the
theory proposed by Budal and Falnes (1975) that the most promising and economic form of WEC is a point absorber tuned to the
frequency of the waves was followed at Lancaster. This theory,
however, predicts that such a performance can only be achieved if
the device undergoes oscillations whose magnitude may be many
times that of the incident wave amplitude. This behavior is not
permissible in practice and has led to the development of theories
to predict the maximum capture width when the amplitude of the
device oscillation is constrained in magnitude but permitted to
maintain the frequency of oscillation (Falnes, 2002)
Oyster is a typical example of point absorber OWEC. The concept for this device originated from research at Queen's University,
Belfast, and is consisted of a buoyant ap which is hinged at the
seabed and moves back and forth with movement of the waves.
This back and forth oscillations operate hydraulic pistons that feed
high-pressured water to a hydro-electric turbine, which generate
electric power. The rst full-scale prototype of the device was
installed at the European Marine Energy Center (EMEC) in Orkney
when it began producing power to the grid in 2009. Recently Renzi
et. al. (2014) has presented a mathematical model, called ap-type
absorber which describes the dynamics of Oyster. Another point
absorber OWEC that has reached the stage of offshore device
testing is a heaving point absorber WEC which was developed at
Uppsala University. In the device the collector body which is a
oating buoy at the sea surface directly drives a linear generator
for power production. The buoy at the sea surface is connected
through a wire to a linear generator placed in a capsule on the
seabed. The translator inside the linear generator follows the
heaving motion of oating buoys, thus inducing a varying magnetic ux in the stationary stator windings, Leijon et al. (2005).
Recently Bozzi et. al. (2013) used the boundary element package,
ANSYS AQWA, to investigate the feasibility of this heaving point
absorber WEC for wave energy exploitation under Mediterranean
Sea conditions. Their numerical results show that the estimated
power captures are in good agreement with those obtained by
Babarit et. al. (2012) for the same device. ANSYS AQWA (2010) was
also used by Bhinder et al. (2009, 2015b) for modeling the motion
response of a pitching WEC,WRASPA. The capabilities of the
method are validated against both corresponding experimental
data and a time-domain CFD method. It is shown that the time
domain methods offer incorporation of viscous and vortex phenomenon and capturing ow details such as turbulence. These
methods however require signicant computer resources and long
runtimes. The BEM method on the other hand relies on the linear
frequency domain hydrodynamic calculation which provides an

M.T. Rahmati, G.A. Aggidis / Ocean Engineering 111 (2016) 483492

efcient numerical approach capable of capturing the major features of interest while reducing the solution time to an acceptable
level for use in routine design.
One of the widely used boundary-element method package has
been WAMIT. An assessment of this package for modeling a freeoating sloped wave energy device, developed at Edinburgh University, was reported by Payne et al. (2008a,b). Although there are
signicant discrepancies between numerical computations and
experimental measurements of the capture width, the agreement
is generally good enough, to provide a rst assessment of the
performance of the WEC. They suggested that the validity of linear
boundary-element method for this kind of wave energy converter
is case specic. Recently Zurkinden et al (2014) provided an
assessment of WAMIT for the prototype of a point absorber OWEC.
The laboratory device developed at Aalborg University is similar to
the Wavestar oat which is located in the Danish North Sea. It was
shown that the linear assumption is a good approximation of the
physical model of the device. In particular, the capture width
predicted using the linear numerical methods agrees well with the
experimental capture width.
This paper brings together the latest developments in analyzing numerical and experimental capture width of WRSPA (Rahmati
et al., 2008a,b) which is a xed OWEC operating in surge and pitch
mode. A 1:100 scale model of the device was developed. A brief
description of the model and its dimension is given in the next
section; this is followed by the numerical power estimation and
the results of wave tank tests on the scaled model in irregular
waves. The experimental results are provided to serve two purposes: the rst is to evaluate the actual optimum power of the
device in irregular waves; the second is to verify the capability of
the numerical method in predicting the optimum power of the
device.

2. The WRASPA device


WRASPA (Wave-driven, Resonant, Arcuate action, Surging
Point-Absorber) was developed at Lancaster University for moderate deep water with water depths of 2050 m. The dynamics of
this wave energy converter (Fig. 1) is similar to other bottomhinged WECs e.g. Oyster (Whittaker and Folley, 2012) and
WaveRoller (Lucas et al., 2012). In this device, wave forces act on
the face of a collector body carried on an arm that rotates about a
xed horizontal axis below sea level. Accordingly, the body

485

oscillates in pitch direction at about the frequency of the ocean


swell. For strong waves during storms, the arm below sea level
automatically moves to a position that minimizes the forces on the
collector to ensure its survival (Rahmati et al., 2008a,b; Chaplin et
al., 2009).
As shown in Fig. 2, the arm is connected to a rotating shaft
using wire and a crank. The rotation of the arm in the pitch
direction around the pivot point rotates a shaft which is placed
above the mean water level. The power is absorber by applying the
band brakes.
A diagram showing the collector body which rotates around a
specied pivot is shown in Fig. 3. The pivot axis depth, q, can be
varied between 200300 mm. The position of center of pressure, p
can be also varied by changing the vertical position of the collector
body on the arm. Experiments on several collector bodies with
various shapes showed that the one with a streamlined shape,
depicted in Fig. 4, give better performance and large amplitudes of
motion to extract energy from waves (Aggidis et al., 2009). The
height, width and length of the collector body are 150, 256 and,
100 mm respectively. The dry weight of the collector body is 140 g.
Discounting higher-order effects, the dynamics of WRASPA
device can be modeled as a rotational springdamper system
using a second-order differential equation in terms of the angle of
pitch, , thus
I Q a55 b55 _ c55 q  pF 1  F CON

in which I Q is the dry moment of inertia and a55 the added


moment of inertia due to immersion about the normal axis
through the pivot (in kg m2), b55 is the radiation damping coefcient (in N m s per radian), and c55 the restoring moment coefcient (in N m per radian).
The excitation moment, F 1 q p, consists of the product of a
force coefcient, F 1 (in N m  1), relative to the wave amplitude,
(in m), and a moment arm dened by the distances p and q (in m),
which are the perpendicular distances from the undisturbed free
surface to the center of wave pressure and the pivot respectively
(see Fig. 3), F CON is the applied control moment. The magnitude of
the parameters a55 , b55 , F 1 and F CON vary with the wave frequency,
their numerical subscripts conforming to the standard six-degreeof-freedom model notation. In this model notation 1 corresponds
to the surge mode and 5 to the pitch mode.
Using complex numbers Eq. (3) can be rewritten in frequency
domain. The linear frequency-domain model of the motion of the
controlled device in a regular wave, can be expressed as the

Fig. 1. Schematic diagram of the dynamics of WRASPA, the device moves back and forth with movement of waves. (a) Backward movement. (b) Forward movement.

486

M.T. Rahmati, G.A. Aggidis / Ocean Engineering 111 (2016) 483492

where the superscript * denotes complex conjugation. The performance of the model in regular wave has already been studied
both numerically and experimentally (Rahmati et al., 2008a,b).
Here, the performance of this WEC in irregular wave is
investigated.
As mentioned by (French, 2006) broadly speaking, the main
challenge in developing economical WECs is how to deal with
large forces moving slowly efciently. This suggests two criteria for
an economical WEC: a large working area relative to its size and a
high ratio of the speed of that surface to the particle speed of the
wave. Regarding the rst criterion Budal and Falnes (1978)
demonstrated the advantages of point absorbers in terms of ratio
of working surface to total area. While there are some very large
WECs, Point absorber WECs are relatively small compare to the
most common wavelengths and have the potential for more efcient power conversation in terms of output per unit volume
(McCabe et al., 2010). The second criterion indicates a resonant
system, and this paper is conned to WRASPA that have quasiresonant working surfaces. The system will be quasi-resonant
because the mechanics involved is not quite that of classical
resonance (French, 2006). However these devices tend to have
natural frequency responses of narrow bandwidth due to relatively small size, only achieving high efciency when excited by
waves with a frequency around their resonance point. In regular
(sinusoidal) waves optimum power output is obtained when the
motion of the device is controlled so that the phase and amplitude
of oscillation have specic optimum values, known as tuning the
device. As for any WEC, the requirement of keeping the device in
resonance, critical to economical power capture, involves continuous tuning. The optimization problem of maximizing the
absorbed wave power in sea waves is more complicated as they
are not sinusoidal. It is necessary to apply control-engineering
techniques in order to optimize the oscillation for maximizing the
absorbed power or the converted useful power. This requires
knowledge of the spectrum wave and of the optimum oscillation
(Falnes, 2002).

Fig. 2. Schematic diagram of the power take off WRASPA.

Fig. 3. Schematic diagram of WRASPA WEC.

3. Numerical dynamic performance

Fig. 4. The streamlined shape collector geometry.

product of the complex impedance, Zf , and complex velocity


amplitude, Uf if , as
(
"
#)
c55
Zf Uf b55 f i2 f I Q a55 f 
Uf
2 f 2
q pf F 1 f  F CON f

in which F CON f is the applied control moment (Tucker and Pitt,


2001). The qualier in parentheses, f , denotes variation with the
wave frequency f. Optimum power absorption occurs (McCabe
et al, 2006) when
F CON f Z  f Uf

Frequency domain analyses are conventionally used as the rst


step in validation and analysis of the performance of wave energy
converters. For the evaluation of motions and forces the program
system WAMIT (Wave Analysis, developed at Massachusetts
Institute of Technology) for wave/structure interaction at zerospeed is applied (WAMIT, 2002). In the program a threedimensional panel method using potential theory is implemented (Lee and Newman, 2005). WAMIT solves the equation of
motion in frequency domain. The Green's theorem is applied to
derive integral equations for the radiation and diffraction velocity
potentials which are represented by the distribution of singularities over the body wetted surface. The body surface is discretized
so that a critical frequency can be determined.
The critical angular frequency is specied to satisfy the validity
criteria of a minimum wavelength greater than seven times the
maximum panel dimension. The critical frequency in this case, is
21.19 rad/s (3.37 Hz), well in excess of the highest required frequency. The frequency-dependent hydrodynamic parameters (a55,
b55 and F1) and the hydrostatic parameter, c55, are derived at
intervals of 0.5 rad/s up to 10 rad/s. It is shown that the excitation
moment magnitude has maximum value when the collector body
is completely submerged and the immersion value is zero. Similarly the values of added inertia and radiation damping have
maximum values when the immersion value is zero. It can be seen
in Fig. 5 that the value of parameter, a55 is changed by decreasing
immersion. Except for the case of zero immersion the values of

M.T. Rahmati, G.A. Aggidis / Ocean Engineering 111 (2016) 483492

Fig. 5. The added inertia for various immersions.

487

Fig. 8. The relative excitation moment phase for various immersion.

Fig. 6. The Radiation damping for various immersions.

Fig. 9. The pitch stiffness.

Fig. 7. The relative excitation moment magnitude for various immersions.

b55 and F1 are not signicantly changed by increasing or


decreasing the immersion, particularly when frequency is less
than 6 rad. (Figs. 69)
The focus of this study is on the performance of the device at a
resonance frequency of about 1 Hz. This is because the aim is to
place the full-scale device at North Atlantic spectrum (in a place
with a Bretschneider spectrum) with a wave period of around
10 s. The hydrodynamic parameters of the device at the angular
frequency of 6.28 rad/s are shown in Fig. 1013. These values are
used to estimate the power output of the device in
irregular waves.

Fig. 10. The added inertia at angular frequency 6.28 rad/s.

4. Numerical power estimation in irregular waves


Pizer (1993) showed that the mean power absorbed, P IN;opt , by
an optimally tuned unconstrained point absorber in a regular
wave of amplitude, , and frequency, f , is given by
P IN;opt

q  p2 jF 1 j2 2
8 b55

Fig. 11. The Radiation damping at angular frequency 6.28 rad/s.

Irregular waves are often dened by their energy period, T e ,


and signicant wave height (here using the spectral moment
denition), H m0 . The equivalent regular wave, in terms of energy
p
capacity, has frequency, f e ( 1=T e ), and amplitude H m0 =2 2.
If the incident wave comprises the superposition of more than one

488

M.T. Rahmati, G.A. Aggidis / Ocean Engineering 111 (2016) 483492

where Sf is the spectral density function of the wave, and df the


component frequency derivative. The optimum mean power is,
consequently, dependent on the wave spectrum, several wave
spectra have been specied for various types of irregular wave.
The basic energy spectrum used here is the Bretschneider spectrum (Tucker and Pitt, 2001). The energy density of each frequency
component in a Bretschneider spectrum is:
SB f

Fig. 12. The relative excitation moment magnitude at angular frequency 6.28 rad/s.

Q H 2m0 f
4

5

h
i
4
exp  Q f

where

4
Tp
Q
1:057

12

13

in which Tp is the peak period, the reciprocal of the peak frequency, fp, of the spectrum. A spectral moment is dened as
Z 1
n
f Sf df
14
mn
0

Fig. 13. The relative excitation moment phase at angular frequency 6.28 rad/s.

frequency component, thus

N
X

f k cos 2 f k t k

k1

in which t is the surface elevation, N the number of component


frequencies, f k , which are uniformly spaced, f k the component
wave amplitudes, and k the component wave phases. Assuming
simultaneous optimal tuning at every frequency component, the
mean power absorbed by an unconstrained point absorber, with
an incident wave described by Eq. (6) is given by
(
)
N
X
q  pk 2 jF 1 j2 f k 2
P IN;opt
8
8 b55
k1

and the corresponding energy and zero-crossing periods and signicant wave height can be expressed in terms of the spectral
p
moments as: T e m  1 =m0 , T z m2 =m0 , and H m0 4 m0 ,
respectively.
Irregular waves are often dened by their energy period, T e ,
and signicant wave height (here using the spectral moment
denition), H m0 . The equivalent regular wave, in terms of energy
p
capacity, has frequency, f e ( 1=T e ), and amplitude H m0 =2 2.
In sea-state tuning control, the control force and moment is
derived for a xed tuning frequency, f T , The mean power output in
an irregular wave with a spectral density function Sf , is

2
Z 1
b55 f q  pf F 1 f 2 Sf
P IN
df
15

2
0
Zf Z  f T
The estimated rst-order power spectra depicted in Fig. 14 are
for the tank model in a Bretschneider spectrum wave with signicant wave height, H m0 0.033 m, zero-crossing period, T z .96
second. There are three tuning frequencies, corresponding to the
reciprocals of the zero-crossing period, T z , the energy period,
T e 1.02 seconds, and the spectrum peak period, T p 1.0 seconds.
The enveloping curve is the optimum mean power spectrum.
The values of mean power absorbed, P IN , are given in Table 1,
along with the proportion of optimum mean absorbed power
these represent. The optimum mean absorbed power is
438.38 mW. Not unexpectedly, the mean power absorbed by the

The energy per unit area of the irregular wave, Eq. (6), is given
by
E g

N
X

Sf k f

k1

g t2
N
X
f k 2
g
2
k1

where is the density of the uid, g is the acceleration due to


gravity, Sf k is a component of the spectral density function of the
wave, Sf , and f the component frequency interval. Hence, the
component wave amplitudes are related to the spectral density
components by

f k 2 2Sf k f

10

As N-1, and f -0, Eq. (8) for the optimum mean power in
irregular becomes:
)
2
Z 1 (
q  pk jF 1 j2 Sf
P IN;opt
df
11
4 b55
0

Fig. 14. Absorbed power against frequency with different f T .

M.T. Rahmati, G.A. Aggidis / Ocean Engineering 111 (2016) 483492

sea-state-tuned system is less than the optimum. It can be seen


that the choice of tuning period has a major effect on the power
absorbed, hence on the power output. With this type of irregular
wave input, the greatest amount of absorbed power is obtained by
tuning to the peak period of the wave spectrum. It is important to
note that this results relay on simplied mathematical models
which is based on several assumptions and approximations. The
model is based on the assumptions of inviscid and irrotational
ow, small amplitudes of motion of the device relative to its
dimension and small wave amplitudes relative to the wavelength.
So because of non-linear and viscous effects, it is expected that the
estimates of mean annual power absorption of actual device could
deviate from what is predicted by the numerical model. In the next
section the actual absorbed power of WRASPA is measured in a
wave tank and the results are used to assess the validity of the
numerical models.

5. Experiment in irregular waves

of the exciting force acting on the device. For maximum power


absorption, a WEC should respond in resonance with the waves
such that the exciting force and the response velocity are in phase,
Falnes (1995). While, with appropriate design, a WEC can possess
inherent dynamics such that its natural response frequency is in
the correct range to match the central excitation frequency of most
sea-states, active control is necessary to maximize power capture
across a range of sea-states and in irregular (realistic) seas where a
range of component frequencies are present.
A simplied owchart of the control system used to efciently
extract power from irregular waves where amplitudes vary from
wave to wave is shown in Fig. 16. The input to the control system is
the ocean wave conditions such as wave elevations which are
measured using wave gages placed in front of the device. Six wave
gauges were arranged to measure the oncoming local waves 1.2 m
ahead of device (see Fig. 15). A LabVIEW program is used to
evaluate the wave condition. The wave climate is compared with a
set of thresholds and the system applies braking torques accordingly. This system can have either four (e.g. the model shown in
Fig. 2) or eight steps. Each step represents a hydraulic brake closely

The experimental studies of controlled motion of the device, in


irregular waves were performed at University of Manchester wave
tank. The tank length, width and depth are 18 m, 5 m and 0.6 m
respectively, see Fig. 15. Waves were generated by eight Edinburgh
Designs piston-type wave paddles which are capable of generating
Bretschneider, Pierson Moskowitz, and Jonswap spectra with
spectra frequency in range of 0.52 Hz. The Bretschneider spectrum wave was generated in the tank for this experiment and the
device is tuned to the peak period of the sea state.
A latching control system is used in this study to capture the
wave power efciently. Latching control, which was rst proposed
by Falnes and Budal (1978) and French (1979) is based on the
observation of the dynamic hydrodynamic force. In this method
the hydrodynamic force is predicted for, at least, a semi-wave
period ahead of the present time. Falnes (1995) suggested that the
control signal for latching a device could be inferred via a twostage prediction process. In this method, the wave-eld would be
measured sufciently ahead of the device as to allow rst, an
accurate wave-elevation prediction at a point closer to the device;
and next, a causalized determination based on this information,
Table 1
Mean absorbed power
Tuning period

Zero-crossing period, Tz
Energy period, Te
Spectrum peak period, Tp

Mean absorbed power Proportion of opti-

Cw (%)

(P IN , mW)

mum P IN (%)

292.25

46

140

355.72
438.38

56
69

170
210

489

Fig. 16. The control system owchart.

Fig. 15. The diagram of the experimental set up (side view).

490

M.T. Rahmati, G.A. Aggidis / Ocean Engineering 111 (2016) 483492

modeling the load that would be applied by a hydraulic system.


Fig. 17 shows how the control system acts based on the wave
condition. The power take off system matches the wave power at
various conditions. For example, when the wave power is negligible, no brake is applied and when wave power is very signicant
both brake 1 and brake 2 are applied.
The motion of the arm was geared up (using thin high strength
wire and pulleys) to drive a 3 mm diameter shaft above the water.
A precision servo-potentiometer is used to measure the motion of
the arm. The two brakes used shown schematically in Fig. 2 exert a
Coulomb-type friction load, closely modeling the load that would
be applied at full scale by a xed pressure, xed area hydraulic
system. To measure power, the geared-up shaft was acted on by
3 electrically operated clutches each connected to a small bandbrake. These were loaded with hanging weights of various weights
from 100 to 500 g such that switching the brakes on in different
combinations gave 3 levels (or steps) of torque.

6. Actual power output and comparisons with other WECs


In Table 2, capture width and wave power of the model at the
peak frequency of 1 Hz is shown. It can be seen that a maximum of
935 kW power from a full scale device in real sea climate can be
captured which is very signicant considering the small size of the
device. The range of the capture width ratio of the device is from
33% to 60%. This is signicantly less than the range 140210%
predicted by the frequency domain method.
The main reason for these discrepancies is that the commercial
code WAMIT is a low order frequency domain method. In this
method viscous effects, separation and ow rotation are ignored in
the numerical calculations. Some efforts have been made to
include the viscous effect by adding a damping term similar to the
drag term in Morison's equation by an estimation of the drag
coefcients (Bhinder et al., 2015a). For example, by investigating
various WECs, Babarit at al. (2012) showed that that the inuence
of the viscous losses on the energy absorption of several WECs is
up to 30%. However, these rough estimates can bring uncertainty
in the model. Furthermore in these methods Linear wave theory
applies, so any nonlinear interactions between the wave

Fig. 17. The torque vs. the brake steps.

Table 2
Capture width and wave power of the model at the peak frequency of 1 Hz.
Hm0 (mm) cw (%) Wave power (mW/
m)

Full scale equivalent power


(MW)

19.7
29.8
33.2
38
42.5

0.161
0.573
0.619
0.92
0.935

33
53
46
60
48

186
368
528
598
748

components are ignored. Also it is important to note that in


practice the optimum power of the point absorber is highly
dependent on the constrained amplitude of device oscillation and
the efciency of control system used. Such power capture predictions are typically made using frequency-domain models of the
oating body response (Falnes, 2002; Newman, 1977). Those
models have also been used to predict actual device performance
(Falcao and Rodrigues, 2002; Delaure and Lewis, 2003; Zurkinden
et al., 2014). Generally reasonable agreement has been observed
between frequency domain models and experimentally measured
response of a heaving buoy in the frequency domain (Payne et al,
2006) but, lower accuracy has been observed for the response of a
surging or pitching body (Meadowcroft et al., 2006).
Among the assumptions upon which the frequency-domain
model (and the associated linear time-domain model) is based is
that the incident wave is simple harmonic and of sufciently small
amplitude so that, if the body is stable, the resulting motions are
proportionally small. The body motions are regarded to be so small
that any change in the angular position of the body can be disregarded. This assumption is in agreement with the design requirements of large offshore oil and gas projects for which these
frequency-domain models were originally designed. However, for
many wave energy conversion devices such as WRASPA, it is desirable for the device response to be resonant with the incident wave in
order to maximize power capture. Resonant response is typically
associated with high-amplitude oscillations and so the assumption of
negligible motion is invalid. So these methods provide an upper
estimate for wave energy absorption (Babarit, 2012). To include the
nonlinear effects Folley and Whittaker (2010) develop a spectral
model which is an extension of a frequency-domain model. In this
model the non-linear force are included and thereby provides
improved estimates of wave energy converter performance, without
the high computational cost of a time domain model. The suitability
and accuracy of a spectral model representation is demonstrated for
a ap type wave energy converter, by modeling the effect of vortex
shedding and large amplitudes of motion. Despite the shortcomings
of the frequency domain models, they remain fundamental to the
understanding of the underlying hydrodynamics of WECs because of
the clarity with which they illustrate how WEC hydrodynamics
inuence power capture.
In Table 3 the mean capture width ratios of three categories of
WECs are presented. It can be seen that the most efcient categories of WECs are xed OWSCs with a mean CWR of 3 times that
of oating OWECS and two times that of heaving and overtopping
devices. The mean capture width ratio of WRSPA at 39% is slightly
more than the mean capture width ratio of other xed OWECs.
This highlights the fact that the most efcient categories of WECs,
in terms of absorbing wave energy, are xed OWSCs and WRASPA
have a similar performance.
In Table 4 the capture width ratio of WRASPA is compared with
other xed OWECs. The results show that the device has a very
large capture width compared with other OWECS. The capture
width ratio is signicantly higher than the Lancaster exible bag
which was previously developed at Lancaster University. The
performance of the device is very similar to Bristol Cylinder wave
energy converter and much higher than other devices such as
wave plunger and Lancaster Clam.
Table 3
Comparison of Energy performance of WRASPA with other xed OWECs
(Babarit, 2015).
OWEC

Overtopping
devices

Mean cw 17

Heaving
devices

Floating
OWSCs

Fixed
OWSCs

WRASPA

16

12

37

39

M.T. Rahmati, G.A. Aggidis / Ocean Engineering 111 (2016) 483492

Table 4
Comparison of energy performance (cw) of WRASPA with other xed OWECs
(Babarit, 2015).
OWEC

cw Resources (kW) Characteristic dimension


(width, m)

WRASPA
Biopower
Edinburgh duck
Bristol cylinder
Lancaster exible
bag
Lanchester clam
Wave plunger
Top-hinged aps
WEPTOS

46
45
47
46
9

53
38.5
54
48
51

25.6
6.6
37
75
20

23
16
25
10

51
16
25
16

27
15
25
2.9

491

useful advice, help and guidance given during the experimental


works at University of Manchester and Lancaster University.

References

It is also important to note that WRASPA is a relatively small


device that would be economically competitive with other technologies and would be relatively easy to install and maintain
because of the simplicity of its design and having very few moving
parts underwater. Moreover, the actual WRASPA is designed for
near shore so, it will be easily accessible and most of the electrical
components can be located onshore thus reducing cost of installation and maintenance.

7. Conclusion
The optimum power of a point absorber wave energy converter
in irregular wave was estimated using a linear potential theory, a
method commonly used for evaluating the performance of WECs.
Numerical results show that the choice of tuning period has a
major effect on the power output of the device. The frequency
model predicts that when the device is tuned with the spectrum
peak period, it could absorb more than twice energy than was
contained in its own geometrical width. The wave tank experiments in a Bretschneider spectrum, however, show that the capture width ratio of the surging pitch point absorber in this study is
in the range of 3360%, considerably less than what is predicated
by the frequency domain method. This is because in frequency
domain models it is assumed that the incident wave is simple
harmonic and of sufciently small amplitude so that the resulting
motions are proportionally small. However, resonant response of
WRASPA and similar devices are associated with high-amplitude
oscillations which violates the assumption of small amplitude
motion. The results demonstrate that the linear frequency domain
methods numerical methods provide an upper estimate for wave
energy absorption. It is also highlighted that signicant discrepancies in comparison with experiments may be observed due
to the high-amplitude oscillations of device, the nonlinear interactions between the wave components and restricted efciency of
the power take off system in irregular waves. In comparison with
other xed OWECs, however, this point absorber has an excellent
performance in terms of absorbing wave energy.

Acknowledgments
The authors are indebted to late Dr. Andy McCabe for his help in
performing the numerical studies in this study. We would also like
to thank the Joule Center (Grant no. JIRP306/02) for providing
nancial support for this work. We thank two anonymous
reviewers for their constructive comments and highly valuable
suggestions, which helped us to improve the manuscript. We are
also extremely grateful to Bob Chaplin and Tim Stallard for their

Aggidis, G.A., Rahmati, M., Chaplin, R., McCabe, A.P., Mingham, C., Causon, D.,
Bhinder, M., 2009. A control system for a new wave energy converter in irregular wave climates. In: Proceedings of the 28th ASME International Conference on Ocean, Offshore and Arctic Engineering, Honolulu, HI.
ANSYS AQWA, User Manual v.13.0, 2010. Century Dynamics Ltd., Horsham, UK.
Antonio, F., Falcao de, O., 2010. Wave energy utilization: a review of the technologies. Renew. Sustain. Energy Rev. 14 (3), 899918.
Babarit, A., 2015. A database of capture width ratio of wave energy converters.
Renew. Energy 80, 610628.
Babarit, A., Hals, J., Muliawan, M.J., Kurniawan, A., Moan, T., Krokstad, J., 2012.
Numerical benchmarking study of a selection of wave energy converters.
Renew. Energy 41, 4463.
Bhinder, M.A., Babarit, A., Gentaz, L., Ferrant, P., 2015a. Potential time domain
model with viscous correction and CFD analysis of a generic surging oating
wave energy converter. Int. J. Mar. Energy 10, 7096.
Bhinder, M.A., Rahmati, M.T., Mingham, C.G., Aggidis, G.A., 2015b. Numerical
hydrodynamic modelling of a pitching wave energy converter. Eur. J. Comput.
Mech. 24 (4).
Bhinder, M.A., Mingham, C.G., Causon, D.M., Rahmati, M.T., Aggidis, G.A., Chaplin, R.
V., 2009. Numerical modelling of a surging point absorber wave energy converter. In: Proceedings of the 8th European Wave and Tidal Energy Conference
EWTEC, Uppsala, Sweden.
Bozzi, S., Moreno Miquel, A., Antonini, A., Passoni, G., Archetti, R., 2013. Modelling
of a point absorber for energy conversion in Italian seas. Energies 6 (6),
30333051. http://dx.doi.org/10.3390/en6063033.
Budal, K., Falnes, J., 1975. A resonant point absorber of ocean-wave power. Nature
256, 478479.
Budal, K., Falnes, J., 1978. Wave-power conversion by point absorbers. Nor. Marit.
Res. 6 (4), 211.
Chaplin, R.V., Folley, M.S., 1998. Sea-bed devices: technical comparisons of existing
and new types. In: Proceedings of the 4th European Wave Energy Conference,
Patras.
Chaplin, R.V., Rahmati M.T., Gunura K., Ma, X., Aggidis, G.A., 2009. Control systems
for Wraspa. In: Proceedings of the IEEE International Conference on Clean
Electrical Power.
Delaur, Y.M.C., Lewis, A., 2003. 3D hydrodynamic modelling of xed oscillating
water column wave power plants by a boundary element methods. Ocean Eng.
30, 309330.
Evans, D.V., 1980. Some analytic results for 2D and 3D wave energy absorbers. In:
Count, B.M. (Ed.), Power From Sea Waves. Ac. Press, Edinburgh, UK,
pp. 213249.
Falcao, A.F., Rodrigues, R.J.A., 2002. Stochastic modelling of OWC wave power plant
performance. Appl. Ocean. Res. 24, 5971.
Falnes, J., 1995. Principles for capture of energy from ocean waves, Phase control
and optimum oscillation. Institutt for fysikk, Norway, Technical report.
Falnes, J., 2002. Ocean Waves and Oscillating Systems. Cambridge University Press,
Cambridge, UK.
Folley, M., Whittaker, T., 2010. Spectral modelling of wave energy converters. Coast.
Eng. 57 (10), 892897.
French, M.J., 1979. A generalized view of resonant energy transfer. J. Mech. Eng. Sci.
21 (4), 299300.
French, M.J., 2006. On the difculty of inventing an economical sea wave energy
converter: a personal view. Proc. IMechE. M 220, 149155.
French, M.J., Bracewell, R.H., 1985. Heaving point absorbers reacting against an
internal mass, hydrodynamics of ocean wave energy utilization. In: Proceedings
of the UTAM symposium, Lisbon.
Lee, C.H., Newman, J., 2005. Computation of wave effects using the panel method.
In: Chakrabarti, S. (Ed.), Numerical Models in FluidStructure Interaction. WIT
Press, Southampton, United Kingdom.
Leijon, M., Bernhoff, H., Agren, O., Berg, M., Isberg, J., Sundberg, J., Karlsson, K.-E.,
Wolfbrandt, A., 2005. Multiphysics simulation of wave energy to electric energy
conversion by permanent magnet linear generator. IEEE Trans. Energy Convers.
20 (1), 219224.
Leijon, M., Lindroth, S., 2011. Offshore wave power measurements A review.
Renew. Sustain. Energy Rev. 15 (9), 42744285.
Lopez, I., Andreu, J., Ceballos, S., Martnez de Alegra, I., Kortabarria, I., 2013. Review
of wave energy technologies and the necessary power-equipment. Renew.
Sustain. Energy Rev. 27, 413434.
Lucas, J., Livingstone, M., Vuorinen, M., Cruz, J., 2012. Development of a wave energy
converter (WEC) design toolapplication to the WaveRoller WEC including
validation of numerical estimates. In: Proceedings of the Fourth International
Conference on Ocean Energy, Dublin, Ireland.
McCabe, A.P., Bradshaw, J.A., Meadowcroft, C., Aggidis, G.A., 2006. Developments in
the design of the PS Frog Mk 5 wave energy converter. Renew. Energy 31,
141151.
McCabe, A.P., Bradshaw, A., Widden, M.B., Chaplin, R.V., French, M.J., Meadowcroft,
J.A.C., 2003, PS Frog Mk5: an offshore point-absorber wave energy converter.

492

M.T. Rahmati, G.A. Aggidis / Ocean Engineering 111 (2016) 483492

In: Proceedings of the 5th European Wave Energy Conference, Cork, Ireland,
pp. 3137.
McCabe, A.P., Stables, M.A., Aggidis, G.A., 2010. A Frequency response design
method for the control of wave energy converters in irregular waves. In: Proceedings of the 3rd International Conference on Ocean Energy, 6 October, 2010,
Bilbao.
Meadowcroft, J.A.C., Stallard, T.J., Baker, N.J., Aggidis, G.A., 2006. Absorption of
energy from irregular waves by a buoyant, surging body. In: Proceedings of the
16th International Offshore and Polar Engineering Conference, p. 437.
Newman, J.N., 1977. Marine Hydrodynamics. The MIT Press, Cambridge (MA, USA),
pp. 307311 [9th reprint., 1999, Section 6.19].
Payne, G., Taylor, J., Bruce, T., Parkin, P., 2008a. Assessment of boundary-element
method for modelling a free-oating sloped wave energy device. Part 1:
numerical modelling. Ocean Eng. 35 (34), 333341.
Payne, G., Taylor, J., Bruce, T., Parkin, P., 2008b. Assessment of boundary-element
method for modelling a free-oating sloped wave energy device. Part 2:
experimental validation. Ocean Eng. 35 (34), 342357.
Payne, G.S., Taylor, J.R.M., Parkin, P., Salter, S.H., 2006. Numerical modelling of the
sloped IPS buoy wave energy converter. In: Proceedings of the 16th International Offshore and Polar Engineering Conference, p. 396.
Pizer, D.J., 1993. Maximum wave-power absorption of point absorbers under
motion constraints. Appl. Ocean. Res. 15 (4), 227234.
Platts, M.J., 1981, Engineering development of the Lancaster exible bag. In: Proceedings of the Second International Symposium on Wave and Tidal Energy,
Proceedings BHRA Fluid Engineering, Bedford.

Price, A., Dent, C.J., Wallace, A.R., 2009. On the capture width of wave energy
converters. Appl. Ocean. Res. 31 (4), 251259.
Rahmati, M.T., Aggidis G.A., Chaplin, R.V., 2008a. Investigating pitching-surge
power-absorber wave energy converters. In: Proceedings of the ASME
POWER, POWER2008-60119, Florida, USA, pp. 699705.
Rahmati, M., Aggidis, G.A., Chaplin, R., McCabe, A.P., 2008b. Test on a novel
pitching-surge wave energy converter. In: Proceedings of the World Renewable
Energy Conference, WREC 2008, Glasgow.
Renzi, E., Doherty, K., Henry, A., Dias, F., 2014. How does oyster work? the simple
interpretation of oyster mathematics. Eur. J. Mech. B/Fluids 47, 124131.
Salter, S.H., 1974. Wave power. Nature 249, 720724.
Tucker, M.J., Pitt, E.G., 2001. Waves in Ocean Engineering. Elsevier (Ocean Engineering Series). ISBN: 0-08-043566-1.
WAMIT, 2002. User Manual. WAMIT Inc., Chestnut Hill, Massachusetts, USA, 024672504, www.wamit.com.
Whittaker, T., Folley, M., 2012. Nearshore oscillating wave surge converters and the
development of Oyster. Philos. Trans. R. Soc. A 370, 345364.
Zurkinden, A.S., Ferri, F., Beatty, S., Kofoed, J.P., Kramer, M.M., 2014. Non-linear
numerical modeling and experimental testing of a point absorber wave energy
converter. Ocean Eng. 78, 1121.

You might also like