You are on page 1of 7

4

th
International Conference on Ocean Energy, 17 October, Dublin


1

Testing of a small-scale floating OWC model in a wave flume


R. P. F. Gomes, J . C. C. Henriques, L. M. C. Gato and A. F. O. Falco

IDMEC, Instituto Superior Tcnico, Technical University of Lisbon, Avenida Rovisco Pais, 1049-001 Lisbon, Portugal
Emails: ruigomes@ist.utl.pt; joaochenriques@ist.utl.pt; luis.gato@ist.utl.pt; antonio.falcao@ist.utl.pt



Abstract
The oscillating water column (OWC) wave
energy converter is probably the most attractive
way of converting the energy from the waves into
electricity due to its simplicity and reliability.
Shoreline full-scale prototypes were built in the
1990s and have proved the concept. However, large-
scale exploitation of wave energy can only be
performed off the coast, where large arrays of
floating OWCs are to be deployed.
This paper describes the testing, in a wave flume,
of a 1:120
th
-scale model of an axisymmetric floating
OWC with non-uniform tail tube cross section. The
device consists in a floater pierced by a hollow
cylinder opened at the bottom to the sea water and
at the top to the pneumatic chamber. A turbine
simulator was placed at the top of the pneumatic
chamber. Several values of the turbine damping
coefficient were tested to assess damping influence
on the system dynamics when subject to regular
waves. Two mooring configurations are tested: one
constraining the floater motion to heave; the other
was a slack-mooring. The motion of the floater and
of the inner water column, the chamber air pressure
and the reflection of waves are analysed. A
comparison with numerical results based on linear
wave theory is presented.
Keywords: Floating oscillating water column; model testing;
spar-buoy; wave energy.

1. Introduction
The oscillating water column (OWC) is a type of
wave energy converter (WEC) that has been object of
study since the 1940s, when Yoshio Masuda started
developing a navigation buoy powered by this device.
It consists of a pneumatic chamber open at the bottom
to the sea water and at the top to the atmosphere
through an air turbine. The oscillating motion of the
water column acted upon by waves produces a
bidirectional flow through the turbine that drives an
electrical generator. The oscillatory characteristic of the
flow requires the installation of a turbine and rectifying
valves or, alternatively, a self-rectifying turbine. The
latter option was found to be more robust and cost
effective and is widely used.
The OWC concept has been extensively studied
over the last thirty years. Shoreline full-scale
prototypes were built in the 1980s and 1990s. These
were the cases of the Kvrner multiresonant OWC,
Norway [1], the Pico plant in Azores, Portugal [2] and
the LIMPET plant in Islay, Scotland [3]. The Pico and
LIMPET plants are still operational. Those particular
devices have proved the principle of operation and the
extraction of energy under real sea conditions.
However, the shoreline location means a lower level of
energy resource and introduces limitations into the
deployment of large numbers of devices, as compared
with offshore locations. More recently, floating OWC
devices have been studied and developed, namely the
OE Buoy and the Oceanlinx Mk3 (see [4]). A 1/4
th
-
scale model of the OE Buoy device has been tested in
the Galway Bay (Ireland) since 2006. The Oceanlinx
Mk3 prototype was tested at an offshore location in
2010. The self-rectifying Wells turbine has equipped
most devices (Pico, LIMPET and OE Buoy). More
recently, a self-rectifying impulse turbine was installed
in the OE Buoy. The Oceanlinx Mk3 device was
equipped with a Denniss-Auld turbine and a HydroAir
variable radius turbine. A new self-rectifying air
turbine, named biradial turbine that substantially differs
from previous conceptions, was presented in [5].
Laboratory experiments on an axisymmetric
floating OWC in a wave flume were first reported by
Whittaker and McPeake [6]. Their work was initially
based on the geometry of a navigation buoy, which was
optimized through the variation of several parameters.
Sarmento [7] carried out experiments on a two-
dimensional bottom-standing OWC to validate a
hydrodynamic model [8]. More recent studies on two-
dimensional fixed OWCs were focused on the effect of
the front wall configuration upon the hydrodynamic
efficiency [9] and on the study of the air flow in the
chamber, through the use of the particle image
velocimetry technique [10]. The fixed cylindrical OWC
is another type of device that has been object of some
experimental studies due to its simplicity and to the
availability of analytical solutions in regular waves.
This geometry was used to validate hydrodynamic
models [11] and to study discrete control strategies
[12]. The floating version of the cylindrical OWC,
moored to the tank floor, was the object of studies with
4
th
International Conference on Ocean Energy, 17 October, Dublin


2
[13], and without [11], a turbine simulator, under
regular and irregular wave conditions. The performance
of the OE Buoy under regular tests has been reported in
[14]. That study presented results of small- and
medium-scale tests in a wave tank.
Theoretical and numerical models have been
important tools for the development of WECs. They are
used to assess the performance of WECs at an early
stage of development, to test different configurations
and, more recently, to optimize the shape of the
converter so as to maximize the power extraction (e.g.
[15]). The first known theoretical model of a floating
OWC was formulated by McCormick [16]. Whittaker
and McPeake [6] studied the effect of the floating
OWC dimensions and power take-off damping in the
power absorption spectrum, using a two degree of
freedom model. Recently, Alves et al. [17] developed
the numerical analysis of an axisymmetric floating
OWC by using a boundary element method to account
for the hydrodynamic interferences between the floater
and the OWC. The radiation capabilities of the floater
and the inertia effect of a deeply submerged mass
(rigidly connected to the floater) were found to be
important in the matching of the system dynamics to a
representative incident wave frequency.
This paper presents the results of experimental tests
on the spar-buoy OWC under regular waves at a wave
flume. The spar-buoy OWC geometry was based on the
numerical optimization presented in [15]. The air
chamber pressure, floater motion, interior water surface
displacement and wave reflections were measured. The
turbine was simulated using a porous membrane placed
at the top of the air chamber. The model was tested
with two mooring configurations: one constraining the
floater motion to heave; the other one was a
slack-mooring.

2. The spar-buoy OWC
The principle of operation of a floating OWC is
similar to that of a fixed one. The main difference is
that the structure oscillates, which consequently leads
to radiation of waves. In this case, the relative motion
between the device and the internal free surface
provides the air flow. From oscillating body theory it is
known that a system with two bodies is supposed to
have two resonance peaks, due to the dynamics of each
body (see [18]). If the system is tuned to have those
peaks placed close to the dominant wave frequency, it
is possible to widen the range of frequencies within
which the system performs well. This is one of the
main advantages of floating OWC devices. This work
deals with an axisymmetric floating OWC that consists
of a cylindrical floater pierced by a hollow cylinder
(tail tube) opened at the bottom to the sea water and at
the top to the pneumatic chamber. The pressure
variation inside the air chamber induces a bidirectional
air flow through a self-rectifying turbine. The inner
cross sectional area is larger at the tube lower part, as
proposed by [19] to enhance the hydrodynamic
performance. Fig. 1a presents a cross section view of
the spar-buoy OWC. The dimensions indicated are the
ones of the 1:120
th
-scale model. The geometry used in
this paper is based on the optimized geometry
presented in [15]. This geometry was the one that
maximized the annual average power extraction under
the wave climate at a location off the western coast of
Portugal, the diameter of the floater and the draft
having been fixed at 16 m and 48 m respectively.


(a) (b)
Figure 1: The 1:120
th
-scale model of the spar-buoy with non-
uniform tail tube cross section: (a) Cross section view with
dimensions in mm; (b) Cut view of the 3D CAD model.

3. Numerical model
The theoretical model used to compare with the
experimental results considers incompressible and
irrotational water flow. Linear water wave theory is
applied by assuming wave amplitude and body motions
much smaller than the wavelength. Two heave modes
are considered: one for the device; and the other for the
OWC. The top part of the OWC is modelled as a small
thickness piston with density equal to water density.
The equations of motion can be found in [18] for a two-
body oscillating system. We consider a turbine with a
linear characteristic curve, which is consistent with the
experiments.
The equations of motion of an isolated floating
OWC acted upon by sinusoidal waves of a frequency
( ) are given by the summation of inertia,
hydrodynamic and turbine induced forces. Since the
turbine has a linear pressure-versus-flow-rate
characteristic, the system is linear and time-invariant.
Then it is possible to apply a Fourier transform to the
time-dependent quantities involved in the equations and
directly extract information from the amplitudes of
motions and pressure, and average power captured
under regular waves. The details of this formulation are
presented in [15]. The hydrodynamic coefficients
(added mass and radiation damping) of the two heave
modes were computed using the boundary element
4
th
International Conference on Ocean Energy, 17 October, Dublin


3
method (BEM) WAMIT [20]. The higher-order
discretization method was applied.
A floating OWC in a narrow flume can be modelled
by an infinite array of devices, equally spaced,
positioned along a straight line perpendicular to the
wave propagation direction. The spacing between
devices is equal to the flume width. An array of five
devices was simulated since it is not practical to
discretize an infinite array, and the interference
between devices decreases with the distance between
them. The equations of motions were derived for each
device and the interaction between devices was
considered. Fig. 2 presents the surface discretization of
the devices. The results used for comparison with
experiments concerns the device in the middle position.


Figure 2: Grid of the five-device array used in the
computations of the hydrodynamic coefficients. The device
considered for comparison with experiments is in the middle.

4. Experimental setup
The spar-buoy OWC 1:120
th
-scale model was tested
at the wave flume of the Laboratory of Hydraulics and
Environment of Instituto Superior Tcnico (IST),
Lisbon, Portugal. The flume is 20 m long, 0.7 m wide
and its maximum depth is 0.5 m. It is equipped with a
HR Wallingford wave generation system with a piston-
type wavemaker with active absorption of reflected
waves. A constant-slope dissipative beach is located at
the end of the flume. A side view of the flume and the
model position is presented in Fig. 3.

Figure 3: Side view of the experimental set-up of the tests
carried out at the wave flume (dimensions in meters). Four
wave gauges were used for the measurement of reflected
waves.

The model scale was limited by the maximum water
depth of the flume (0.5 m). The floating OWC model
has a draft approximately equal to 0.4m. Two model
configurations were tested. In the first configuration
(config. A) the model is constrained to oscillate in
heave by using four horizontal 1.2 mm nylon lines,
approximately 300 mm long, connecting the floater to
the flume side walls in a cross-shape configuration
(Fig. 4a). The cables were kept just slightly stretched so
that the stiffness of the cables would provide the
restoring effect on surge motions. The relative large
length of the cables when compared with the model
heave motion amplitude ensures a negligible effect of
the cables on the heave motion. The good stability of
the model, due to the low position of the centre of
gravity, produced a very small pitch motion for all the
frequencies tested (except at the pitch resonance
frequency); this enhanced the comparison with the
numerical model, where the device motion was limited
to heave. In the second configuration (config. B), the
model was moored to the flume floor by a three-section
slack-mooring configuration (Fig. 4b). A 1.2 mm
diameter nylon line was used as mooring line. The two
mooring points on the flume floor (one upstream and
one downstream) were located 1.4 m away from the
model vertical axis. Each line was divided in three
parts, separated by a small weight and a small fully-
submerged buoy, which provided pre-tension on the
lines when the model was at rest. This type of mooring
is one of the most promising configurations for WEC
moorings since small dynamic tensions on the cables
are induced and very small effects on the heave motion
are observed [20].


(a)

(b)
Figure 4: The two model configurations tested at the wave
flume: (a) constrained to oscillate in heave by a nylon cable
cross-shape configuration; (b) moored to the flume bottom
with a three-section slack-mooring configuration separated by
a weight and a buoy (the mooring cables are not visible in the
image).
4
th
International Conference on Ocean Energy, 17 October, Dublin


4
The spar-buoy model was built with PVC and ABS
parts with 2 mm thickness. Ballast was placed at the
lower part of the model to simulate the full-scale
moment of inertia and center of gravity (see Fig. 1). A
3D CAD model was created using SolidWorks
software to compute the model mass, moment of
inertia, buoyancy center and gravity center at design
stage. After construction, the 3D CAD model was
calibrated using a precision weighing balance. The
scale model relevant parameters are presented in
Table 1.

Parameter Value
Floater diameter [m] 0.1333
OWC diameter [m] 0.0490
Floater draft [m] 0.0467
Device draft [m] 0.4050
Centre of gravity, z
g
[m] -0.2630
Centre of buoyancy, z
b
[m] -0.1927
Mass [kg] 1.4708
Mass of ballast [kg] 0.600
Moment of inertia [kg m
2
] 0.15050
Mooring weight mass [kg] 0.0102
Mooring buoy vol. [m
3
] 4.210
-6

Table 1: Physical characteristic of the 1:120
th
-scale model of
the spar buoy OWC.

Regular wave tests were carried out for frequencies
between 0.4 and 1.7 Hz and wave heights within the
range of 0.008 to 0.030 m. The motion of the floater
was captured by a motion tracking system with infrared
cameras developed by Qualisys. The cameras emit
infrared light that is reflected by low-mass targets fixed
on the model. A two-dimensional image of the targets
is detected by the three cameras and a triangulation
technique is used to detect the six degree-of-freedom
motion of the targets (surge, sway, heave, roll, pitch
and yaw). During the configuration A tests (floater
constrained to oscillate in heave), the vertical motion of
the floater was measured by a laser positioning system,
which used the top of the floater as the reference target.
The pressure inside the chamber was measured by a
high-accuracy low-pressure sensor placed in a fixed
position above the model. Standard resistive wave
gauges were used to measure the water elevation at
different positions of the wave flume; one was placed
inside the floater to measure the relative motion
between the floater and water column. Apart from the
motion tracking system, all instrumentation data was
captured at 100 Hz by a PIC microcontroller. The
floater motion was captured at 50 Hz.
The turbine effect on the hydrodynamics was tested
by introducing a porous filter material between the air
chamber and the atmosphere. The laminar flow ensures
an approximately linear relation between the pressure
difference and the air flow rate. A 4 mm thickness
synthetic carpet material was used, similar to the one
used in [7]. Different turbine damping magnitudes were
tested by changing the number of layers of the filters
and the air flow passage area. The instantaneous power
was computed using the pressure signal and the flow
rate; the latter was calculated through a pressure drop
calibration curve or, alternatively, using the water
elevation signal inside the floater (and assuming the
inner free surface to be flat). The three different turbine
damping levels used in the model are presented in
Table 2. The instantaneous power absorbed by the
model is given by P = p , where is the volume
flow rate passing through the turbine simulator and p is
the pressure difference between the air chamber and the
atmosphere.

Filter ref.
(p)
m
10
6

[m
4
s/kg]
(p)
p
10
3

[m
4
s/kg]
M1K 18.18 23.90
M2K 9.79 12.87
M4K 7.16 9.41
Table 2: Linear turbine characteristic curve coefficient p.
The subscript m stands for model scale and p for full scale.

5. Results
Regular wave tests were carried out for both
configurations. The numerical results presented in this
section were obtained from the model presented in
section 3. The time series extracted from the data
acquisition system were analysed using a code
developed for that purpose. The data was filtered to
remove high frequency noise and low frequency
electronic instrumentation variations. Reflections were
computed using the data from four wave gauges placed
between the model and the wavemaker. In the absence
of the model, the reflection coefficient was less than
10%, for the range of frequencies and wave heights
tested. The average incident wave power per unit of
wave crest (E) was computed through a frequency
domain analysis. A time domain analysis was used to
determine the average wave height (E

), average period
(I

), and average model amplitudes of motion (X

1
-
surge, X

2
-sway, X

3
-heave,

1
-roll,

2
-pitch and

3
-
yaw) and pressure (p ).
Decay tests have been carried out to assess the
natural periods of each oscillation mode. Table 3
presents the natural periods of the model in each
configuration.

Mode Surge Sway Heave Roll Pitch Yaw
config. A - - 0.91 - 2.08 -
config. B 16.89 21.92 0.88 1.99 1.97 4.02
Table 3: Natural periods (in seconds) of the model.

Tests have been carried out for 300 s. Results with
configuration A (model constrained to oscillate in
heave) are shown in Fig. 5. The effect of the different
turbine damping magnitudes is presented. The average
amplitudes of model heave motion (X

3
) and relative
heave motion between the model and the OWC
(measured by the wave gauge fixed inside the OWC
tube) X

are nondimensionalized by the incident wave


4
th
International Conference on Ocean Energy, 17 October, Dublin


5
amplitude A

. The nondimensional pressure is p


-
=
p pgA

, where p is the water density and g the


acceleration of gravity. The nondimensional capture
length is I
c
-
= P(EJ), where J is the floater
diameter. The numerical model presents a good
correlation outside the large motion amplitude zones.
The viscous effects, which were not accounted for, are
the major reason for the mismatches between numerical
and experimental results. As a consequence, the
pressure difference and the capture length are also
smaller than what was predicted. The major source of
viscous effect is related to the scale effect, which is a
consequence of the inability to replicate the full-scale
Reynolds number. The increase in the model scale is
expected to reduce this effect. The shape of the curves
is consistent with the results from the computations, as
shown by the two peaks (present in all graphs) that are
characteristic of two-body devices (used to expand the
power absorption spectrum of the device). Also
observed in the graphs is the effect of transversal waves
on the results, occurring at frequencies around 1.5 Hz.
The device heave amplitudes (X

3
) can be higher than
three times the incident wave amplitude for frequencies
close to 1.2 Hz. A smaller peak at 0.8 Hz is also
observed. The OWC relative heave amplitude (X

) is
more than three times the incident wave amplitude over
a wide range of frequencies, between 0.8 and 1.3 Hz.
The tests with filter M2K and M4K provided higher
energy extraction than the ones with M1K.
Figs 6 and 7 presents results for the model with
configuration B. Fig. 6 shows the average amplitudes
of surge, heave and pitch modes. The pitch amplitude
graph exhibits two peaks: the first one, around 0.5 Hz,
corresponds to the natural frequency of oscillation of
the buoy-tube set, as shown in Table 3; the second one
occurs at wave frequencies around 1 Hz (twice the
natural pitch frequency). The roll mode is also excited
at the same frequency presenting similar amplitudes of
motion. This is probably caused by a nonlinear effect,
resulting from the coupled effect between heave and
pitch modes, named in the literature as Mathieu-type
instability [22]. This instability is caused by the
dynamic variation in the device metacentric height
(consequence of the heave motion), which during some
part of the wave cycle becomes negative and
consequently makes the device unstable, causing the
high amplitudes in the pitch and roll modes. This effect
has been reported by other authors [23, 13] and is
responsible for the nonlinearities observed in Figs. 6
and 7, at frequencies around 1 Hz. The nondimensional
amplitudes of the heave motion, OWC relative heave
motion and pressure difference decrease with the
increase in the wave height. This was not observed in
the tests with configuration A.
In terms of power extraction, the tests with
configuration A produced a slightly higher power
extraction. The major differences are suspected to be a
consequence of the Mathieu-type instability, which
removes energy from the desirable heave motion to an
undesirable pitch motion, rather than due to the
mooring configuration.


Figure 5: Experimental results of configuration A, for different turbine damping (see Table 2). Each dot corresponds to a regular
wave test. The numerical results are represented by the solid lines.
4
th
International Conference on Ocean Energy, 17 October, Dublin


6

Figure 6: Model surge, heave and pitch amplitudes in configuration B, for turbine simulator M2K.


Figure 7: Regular wave results of configuration B, for turbine simulator M2K.


6. Conclusions
A small scale spar-buoy OWC model has been
tested in a flume under regular waves. Its geometry is
based on the results of optimization work reported in a
previous paper. A linear turbine was simulated for
comparison with a frequency domain model. The
results presented the trend of the numerical results
when the model was constrained to oscillate in heave.
Viscous effects ignored in the mathematical modeling
were found to be important at this small scale and
prevented a better agreement with the experiments. The
tests with the more realistic slack-mooring
configuration showed the occurrence of a dynamic
instability for a wave frequency near the heave
resonance. This effect has decreased the power
extraction in the vicinity of that frequency, when
compared with the first tested configuration. Further
studies will be devoted to the detection and prevention
of this effect.

Acknowledgements
This work was funded by the Portuguese
Foundation for Science and Technology to IDMEC
through LAETA, and by contract PTDC/EME-
MFE/103524/2008 and contract PTDC/EME-
MFE/111763/2009. The first author was supported by
grant SFRH/BD/35295/2007 from the MIT-Portugal
Program. The second author was supported through
Cincia 2007 initiative.

References
[1] O. Malmo and A. Reitan (1986). Development of the
Kvrner multiresonant OWC. In: D. V. Evans and A. F.
de O. Falco, editors, Hydrodynamics of Ocean Wave
Energy Utilization. Berlin: Springer-Verlag.
[2] A. F. de O. Falco (2000). The shoreline OWC wave
power plant at the Azores. In: Proc. 4th European Wave
Energy Conference. Aalborg, Denmark.
[3] T. Heath, T. J . T. Whittaker and C. B. Boake (2000). The
design, construction and operation of the LIMPET wave
energy converter (Islay, Scotland). In: Proc. 4th
European Wave Energy Conference. Aalborg, Denmark.
[4] A. F. O. Falco (2010). Wave energy utilization: a
review of the technologies. Renewable & Sustainable
Energy Reviews; 14(3):899-918.
[5] A. F. O. Falco, L. M. C. Gato and E. P. A. S. Nunes
(2013). A novel radial self-rectifying air turbine for use
in wave energy converters. Renewable Energy,
50:289-298.
[6] T. J . T. Whittaker and F. A. McPeake (1986). Design
optimization of axi-symmetric tail tube buoys. In: D. V.
Evans and A. F. de O. Falco, editors. Hydrodynamics of
Ocean Wave Energy Utilization. Berlin: Springer-
Verlag.
[7] A. J . N. A. Sarmento (1992). Wave flume experiments
on two-dimensional oscillating water column wave
energy devices. Experiments in Fluids; 12:286-292.
[8] A. J . N. A. Sarmento and A. F. de O. Falco (1985).
Wave generation by an oscillating surface-pressure and
its application in wave-energy extraction. J ournal of
Fluid Mechanics; 150:467-485.
4
th
International Conference on Ocean Energy, 17 October, Dublin


7
[9] M. T. Morris-Thomas, R. J . Irvin and K. P. Thiagarajan
(2007). An investigation into the hydrodynamic
efficiency of an oscillating water column. J ournal of
Offshore Mechanics and Arctic Engineering-
Transactions of the ASME; 129:273-278.
[10] K. Ram, M. Faizal, M. R. Ahmed and Y. H. Lee (2010).
Experimental studies on the flow characteristics in an
oscillating water column device. J ournal of Mechanical
Science and Technology; 24:2043-2050.
[11] R. K. Sykes, A. W. Lewis and G. P. Thomas (2011).
Predicting hydrodynamic pressure in fixed and floating
OWC using a piston model. In: Proc. 9th European
Wave Energy Conference. Southampton, UK.
[12] M. F. P. Lopes, J . Hals, R. P. F. Gomes, T. Moan, L. M.
C. Gato and A. F. de O Falco (2009). Experimental and
numerical investigation of non-predictive phase-control
strategies for a point-absorbing wave energy converter.
Ocean Engineering; 36(5): 386-402.
[13] W. Sheng, B. Flannery, A. Lewis and R. Alcorn (2012).
Experimental studies of a floating cylindrical OWC
WEC. In: Proc. 31st International Conference on Ocean,
Offshore and Arctic Engineering. Rio de J aneiro, Brazil.
[14] Y. Imai1, K. Toyota, S. Nagata, T. Setoguchi and M.
Takao (2011). An experimental study on generating
efficiency of a wave energy converter Backward Bent
Duct Buoy. In: Proc. 9th European Wave and Tidal
Energy Conference. Southampton, UK.
[15] R. P. F. Gomes, J . C. C. Henriques, L. M. C. Gato and A.
F. O. Falco (2012). Hydrodynamic optimization of an
axisymmetric floating oscillating water column for wave
energy conversion. Renewable Energy 44:328-339.
[16] M. E. McCormick (1974). Analysis of a wave energy
conversion buoy. J ournal of Hydronautics; 8(3):77-82.
[17] M. A. Alves, I. R. Costa, A. J . N. A. Sarmento and J . F.
Chozas (2010). Performance evaluation of an
axisymmetric floating OWC. In: Proc. 20th International
Offshore and Polar Engineering Conference. Beijing,
China.
[18] J . Falnes (1999). Wave-energy conversion through
relative motion between two single-mode oscillating
bodies. J ournal of Offshore Mechanics and Arctic
Engineering-Transactions of the ASME; 121(1):32-38.
[19] A. F. O. Falco, J . J . M. B. Cndido and L. M. C. Gato
(2010). Floating oscillating-water-column device for
wave energy conversion. Portuguese Patent Application
PT 105 171 [In Portuguese].
[20] C.-H. Lee and J . Newman (2005). Computation of wave
effects using the panel method. In: S. Chakrabarti, editor,
Numerical models in fluid-structure interaction.
Southampton, United Kingdom: WIT Press.
[21] P. C. Vicente, A. F. de O. Falco and P. A. P. J ustino
(2012). Slack-chain mooring configuration analysis of a
floating wave energy converter. In: Proc. 26th
International Workshop on Water Waves and Floating
Bodies, Athens, Greece.
[22] J . B. Rho, H. S. Choi, H. S. Shin and I. K. Park (2005).
A study on Mathieu-type instability of conventional spar
platform in regular waves. International J ournal of
Offshore and Polar Engineering; 15(2):104-108.
[23] G. S. Payne, J . R. M. Taylor, T. Bruce and P. Parkin
(2008). Assessment of boundary-element method for
modelling a free-floating sloped wave energy device.
Part 2: Experimental validation. Ocean Engineering;
35(3-4): 342-357.

You might also like