You are on page 1of 13

China Ocean Eng., Vol. 30, No. 6, pp.

954 – 966
© 2016 Chinese Ocean Engineering Society and Springer-Verlag Berlin Heidelberg
DOI 10.1007/s13344-016-0062-2, ISSN 0890-5487

Hydrodynamic Analysis and Shape Optimization for Vertical Axisymmetric


Wave Energy Converters*

ZHANG Wan-chao (张万超), LIU Heng-xu (刘恒序)1,


ZHANG Liang (张 亮) and ZHANG Xue-wei (张学伟)
College of Shipbuilding Engineering, Harbin Engineering University, Harbin 150001, China

(Received 14 December 2015; received revised form 13 April 2016; accepted 26 May 2016)

ABSTRACT
The absorber is known to be vertical axisymmetric for a single-point wave energy converter (WEC). The shape of
the wetted surface usually has a great influence on the absorber’s hydrodynamic characteristics which are closely linked
with the wave power conversion ability. For complex wetted surface, the hydrodynamic coefficients have been predicted
traditionally by hydrodynamic software based on the BEM. However, for a systematic study of various parameters and
geometries, they are too multifarious to generate so many models and data grids. This paper examines a semi-analytical
method of decomposing the complex axisymmetric boundary into several ring-shaped and stepped surfaces based on the
boundary discretization method (BDM) which overcomes the previous difficulties. In such case, by using the linear wave
theory based on eigenfunction expansion matching method, the expressions of velocity potential in each domain, the
added mass, radiation damping and wave excitation forces of the oscillating absorbers are obtained. The good
astringency of the hydrodynamic coefficients and wave forces are obtained for various geometries when the discrete
number reaches a certain value. The captured wave power for a same given draught and displacement for various
geometries are calculated and compared. Numerical results show that the geometrical shape has great effect on the wave
conversion performance of the absorber. For absorbers with the same outer radius and draught or displacement, the
cylindrical type shows fantastic wave energy conversion ability at some given frequencies, while in the random sea
wave, the parabolic and conical ones have better stabilization and applicability in wave power conversion.

Key words: vertical axisymmetric; complex wetted surface; semi-analytical method; astringency; geometrical shape

1. Introduction

The slow-speed periodic waves with large forces can cause periodic resonant motion of single or
multiple buoys relative to a reactant through which the wave energy can be extracted. The oscillating
bodies utilizing heave mode have small horizontal dimensions compared with the incident wave length
and are often used in arrays which have been defined as the simplest point-absorber WECs. In such
case, the geometrical configuration of the absorber and the physical parameters of the actuator between
the absorber and the reactant have been the main objectives of optimizing the WECs to obtain the
maximum efficiency over the past several years. This paper presented herein focuses on the former
one.

* This paper is financially supported by the National Natural Science Foundation of China (Grant Nos. 11572094, 51579055 and
51509048).
1 Corresponding author. E-mail: liuhengxu@hrbeu.edu.cn
ZHANG Wan-chao et al. / China Ocean Eng., 30(6), 2016, 954 – 966 955

The geometrical optimization in wave energy utilization originates from the hydrodynamic
characteristics analysis of the floating structures. This pioneer work can be traced to Yeung (1980) who
examined a truncated cylinder with analytical method based on the traditional potential flow theory in
frequency domain in finite water depth and found the geometrical parameters’ effect on the cylinder’s
hydrodynamic performance. Recently, with the increasing interest in the research on wave energy, the
designers (Mavrakos and Katsaounis, 2009; Mavrakos et al., 2009; McCabe et al., 2010; McCabe,
2013; Colby et al., 2011; Gomes et al., 2012; Goggins and Finnegan, 2014; Bachynski et al., 2012)
have devoted their efforts to improving the wave power conversion efficiency by optimizing the
geometrical shape or parameters of their designs. For example, Mavrakos and Katsaounis (2009) and
Mavrakos et al. (2009) investigated the effect of floaters’ geometries on the power conversion
performance of tightly moored vertical axisymmetric wave energy converters. In their analysis, the
traditional cylindrical or conical absorbers attached with some special structures, such as bottom-
mounted vertical and horizontal skirts, two piston-like arranged internal floater and exterior torus, were
examined and comparatively assessed. McCabe et al. (2010) and McCabe (2013) examined a
surge-pitch wave energy converter by using a genetic algorithm. The unsymmetrical and bisymmetric
shape with parametric description, as well as the elementary cost function based on a first-order model
of the system are poured attention during their optimization. Gomes et al. (2012) presented a method
to optimize the geometry of a floating oscillating water column considering the floater’s diameter and
total submerged length and to obtain a perspective on how the dimensions of the device influence the
annual average power of their design. Very recently, Goggins and Finnegan (2014) introduced a
methodology to optimize the hydrodynamic performance of the floating oscillating absorber through
changing the geometric configuration such as the shape and radius. And then they studied in detail the
unstrained system based upon the WEC deployed at the Atlantic marine off the west coast of Ireland.
Usually, the oscillating absorber with simple configuration can be examined using the traditional
analytical method by which the fluid domain is encircled and the device can be decomposed into
several subdomains according to the classification of the boundaries. In such case, the expressions of
velocity potentials, added mass, radiation damping and wave forcing can be analytically presented as
in Yeung (1980) and Mavrakos (2004) where the hydrodynamic characteristics corresponding to the
geometrical parameters can be systematically described. However, for a buoy with complex wetted
surface, the numerical method should be used, and based on this method so much work (Alves et al.,
2007; Falcão et al., 2012; Babarit et al., 2012; Koh et al., 2014; Oskamp and Ozkan, 2012; Kramer and
Frigaard, 2002) has been conducted to improve the wave power conversion performance of the device
though it needs so many gird data and an ocean of calculations. For example, in the research process of
Falcão et al. (2012), to obtain the added mass and the radiation damping coefficient of the buoy-tube
pair, the commercial boundary-element software WAMIT was employed. Koh et al. (2014) proposed a
method to conduct the multi-objective optimum design of a buoy for the resonant-type wave energy
converter, in which the ANALYSIS AQWA was used for the calculations of hydrodynamic
characteristics. In the multi-objective optimization process to scientifically analyze the influence of the
geometrical parameters, sufficient relevant data corresponding to the optimization objective should be
obtained, which results in the establishment of an ocean of gird data and models.
956 ZHANG Wan-chao et al. / China Ocean Eng., 30(6), 2016, 954 – 966

In this paper, the main objective is to focus on the optimization of the geometrical configuration
of the vertical axisymmetric absorber scientifically by introducing a semi-analytical method which is
rarely or never researched before. This was first put forward by Kokkinowrachos et al. (1986) based on
the discretization of the flow field around the structure using coaxial ring elements, which are
generated from the approximation of the body’s meridian line by a stepped curve as shown in Fig. 1. In
such case, the vertical axisymmetric fluid domain can be decomposed into several subdomains which
can be described analytically. By using the linear wave theory based upon the eigenfunction expansion
matching method, the analytical expressions for velocity potentials in each domain, and for the
diffracted and radiated problems are first obtained. Further, the added mass, radiation damping, wave
excitation forces and the corresponding captured power of the oscillating absorbers are calculated
conveniently. Lately, numerical results for the optimization process are presented and the results of the
geometries’ effects on the wave energy conversion are analyzed systematically which can be referred
to in improving the wave energy conversion performance in the future.

Fig. 1. Sketch of the device and the definition of fluid


subdomains.

2. Hydrodynamic Formulations

The oscillating absorber considered in this study is shown in Fig. 1. We defined the cylindrically
coordinated system (r, θ, z) by its origin located at the center of the absorber and on the mean plane of
the free surface. The axis oz is vertically upward. The wetted surface of the absorber is assumed to
comprise a cylindrical surface and a vertical axisymmetric curved surface. The outer radii of the two
surfaces are the same and denoted as R. The cylinder height, the whole draught, and the water depth
are denoted as t, d and h, respectively.
For incompressible and inviscid fluid, and for small amplitude wave theory with irrotational
motion, we can introduce a velocity potential (r, , z, t) to describe the flow fluid. Assuming harmonic
motion in frequency, for the sake of simplicity, the time variation can be omitted and then the velocity
potential can be written as:
 ( r ,  , z , t )  Re[ ( r ,  , z )e  it ] , (1)
where  is the frequency of the incident wave. According to the line wave theory, the spatial velocity
potential  can be decomposed as the undisturbed incident wave velocity potential 0, the scattered
potential 7 for the fixed body and the radiation potential j (j=1, 3, 5) induced by the body motion
oscillation in otherwise calm water. The velocity potentials 0 and 7 comprise the diffracted potential
ZHANG Wan-chao et al. / China Ocean Eng., 30(6), 2016, 954 – 966 957

D around the structure with constraints. Then we have


 ( r, , z )  0 ( r,  , z )  7 ( r, , z )   j 1,3,5  j j ( r,  , z );
D ( r,  , z )  0 ( r,  , z )  7 ( r, , z ), (2)
where, j (j=1, 3, 5) is the j-th mode motion amplitude and j=1 for surge, j=3 for heave and j=5 for
pitch. The aforementioned velocity potentials should satisfy
Laplace’s equation:
 2  0 . (3)
Free surface condition:
 2  g  z  0 z  0, r  R. (4)
Seabed condition:
 z  0 z   h. (5)
Hull boundary condition:
 n j  Vn ,  nD  0. (6)
Radiation condition:
lim

kr ( r  ik )  0, (7)
r

where, g is the acceleration of gravity, i is the imaginary unit and k, the wave number comes from the
dispersion relation, and the symbol n() indicates that the derivative in the normal vector always points
outwards the wetted surface of the device.

3. Added Mass, Radiation Damping and Wave Forcing

Assuming an undisturbed incident sinusoidal wave of amplitude A and frequency  propagating


along the positive x-axis direction can be expressed in the cylindrical coordinates as:
Ag cosh[k ( z  h )] 
0 ( r,  , z )  
 cosh(kh )
l 0  l il J l (kr ) cos(l ) , (8)

where, Jl() is the first kind Bessel function of order l (l=0, 1, 2, ) and l is the Neumann’s symbol,
defined by 0=1 and l=2 for l1. In accordance with Eq. (8), the velocity potential of the diffraction
field caused by the fixed device can be written in the form:
Ag 
D ( r, , z )   l 0  l il D,l (r, z ) cos(l ) .

(9)
In the unbounded fluid domain with finite water depth, the radiation velocity potential caused by
the forced vertical axisymmetric body motion can be expressed as:
1 ( r ,  , z )  1,1 ( r , z ) cos 

3 ( r,  , z )   3,0 ( r , z ) (10)
 ( r,  , z )   ( r, z ) cos 
 5 5,1

The newly introduced velocity potential expressions D,l, 1,1, 3,0 and 5,1 should also satisfy the
former boundary conditions and their solution will be the key problems to the hydrodynamic analysis
of the oscillating buoy.
958 ZHANG Wan-chao et al. / China Ocean Eng., 30(6), 2016, 954 – 966

To obtain the hydrodynamic characteristics of the absorbers, the velocity potentials around the
vertical axisymmetric body should be confirmed. Before decomposing the Laplace’s equation with the
method of separation of variables, we should first divide the fluid domain around the device into
several subdomains. For the complex axisymmetric wetted surface, we decompose it into ring shaped
stepped surfaces by dividing the projection curve on the vertical plane averagely. Thus the velocity
potentials in each domain can be expressed in the form of Fourier series.
(a) Velocity potential in domain E (r  R, h z  0):
 Ej ,l ( r, z )   Ej ,l ,0  0 ( z ) H l (k0 r ) H l (k0 R )
  n 1  Ej ,l ,n  n ( z ) K l ( kn r ) Kl ( kn R )   jlE ( r, z )

(11)
with j=D, 1, 3, 5, Hl() and Kl() are the first kind Hankel function and the modified second kind Bessel
function of order l, respectively. Here and hereafter, a prime denotes taking the differentiation of a
function with respect to its argument. The wave number k0, kn (n=1, 2, 3, …) are defined by 2=
gk0tanh(k0h)=gkntan(knh).
(b) Velocity potential in domain I1 (0  r  R1, h z  h1h) and Ip (2  p  N, Rp1  r  Rp, h z
 hph):
 Ij ,l ( r, z )   jI,l ,0G jlI ( r )
1 1 1

  n 1 [ lj ,l ,n1 I l (n1 r ) I l (n1 R1 ) cos n1 ( z  h )]   jlI1 ( r , z );



(12)
1

 j ,l ( r, z )   j ,l ,0G jl ( r )   n 1  j ,l ,n Qn l ( r ) cos  n ( z  h )    j ,l ,0G jl ( r )


Ip Ip Ip  Ip Ip Ip Ip

p p p p

  n 1  j ,pl ,n p Q n ppl ( r ) cos  n p ( z  h )    jlp ( r , z ). (13)


 I I I

In the eigenfunction expansion for the velocity potentials, the unknown function expressions of
(), G(), Q() and () are given in Appendix A for the convenience to readers. The sets of unknown
Fourier coefficients  are determined by taking advantage of orthogonality, in so-called Garrett’s
method, according to the matching of the potentials and its normal derivative on the juncture boundary
surface shared by subdomains. The flowing series of functions are expressed as:
(a) Coupled vertical plane Ip_Ip+1( 1  p  N1):
hp h hp h

  j ,l ( rp , z )dz    j ,l ( rp , z )dz;
Ip I p 1

h h
hp h hp h

  j ,l ( rp , z ) cos n ( z  h )  dz    j ,l ( rp , z ) cos n ( z  h )  dz;


Ip I p 1

h p h p

hp h h p 1  h

 r j ,l ( rp , z ) cosh  k0 ( z  h )  dz   r j ,l ( rp , z ) cosh  k0 ( z  h ) dz;


Ip I p 1

h h
hp h h p 1  h

 r j ,pl ( rp , z ) cos  kn ( z  h ) dz   r j ,pl1 ( rp , z ) cos  kn ( z  h )  dz.


I I
(14)
h h

(b) Coupled vertical plane Ip_E (p=N):


hp h hp h

  j ,l ( rp , z )dz    Ej ,l ( rp , z )dz;
Ip

h h
hp h hp h

  j ,l ( rp , z ) cos n ( z  h )  dz    Ej ,l ( rp , z ) cos n ( z  h )  dz;


Ip

h p h p
ZHANG Wan-chao et al. / China Ocean Eng., 30(6), 2016, 954 – 966 959

hp h
r j ,pl ( rp , z ) cosh  k0 ( z  h ) dz   r Ej ,l ( rp , z ) cosh  k0 ( z  h )  dz;
0


I

h h
hp h
r j ,pl ( rp , z ) cos  km ( z  h ) dz   r jE,l ( rp , z ) cos  km ( z  h )  dz.
0


I
(15)
h h

By solving the above linear equations, the Fourier coefficients can be confirmed. As the heave motion
mode is applied in the wave energy conversion, the wave forcing and radiation damping coefficients in
heave need to be obtained. Thus, by defining R0=0, the wave excitation force fd3 and hydrodynamic
coefficients (added mass 33 and radiation damping 33) of the absorber in heave can be calculated and
defined by
f d 3  i2π gA p 1 R p 
N Rp
 D,0 ( r, hp  h ) rdr ;
Ip
(16)
R p 1

33  i33   2π p 1 Rp  3,0 ( r, hp  h)rdr .


N Rp Ip
(17)
R p 1

4. Convergence and Verification

The numerical work in this section is involved in the choice of the number of terms used in
infinite simulations. The former 30 terms are adopted in the infinite summations to compute numerical
results because the infinite summations have excellent truncation characteristics as in Yeung (1980).
Two quintessential absorbers with the same radius of R/h=1.0 and draught of d/h=0.2 are adopted here
to verify the feasibility and validity of this semi-analytical method.
Figs. 2 and 3 show the examinations on the convergence of wave forces and hydrodynamic
coefficients of the above mentioned cone and semi-sphere in the presence of regular waves with
different frequencies in water depth of h=50 m, respectively. The convergent results can be achieved
when the discrete number is larger than 15.

Fig. 2. Convergence of wave force and hydrodynamic coefficients in heave with the discrete number for a cone at different
frequencies (rad/s); N stands for the discrete number; (a) wave force; (b) added mass; (c) damping.

Fig. 3. Convergence of wave force and hydrodynamic coefficients in heave with the discrete number for a semi-sphere at
different frequencies (rad/s); N stands for the discrete number; (a) wave force; (b) added mass; (c) damping.

To verify the correctness of the present analytical solution, the analytical solution method (Hulme,
960 ZHANG Wan-chao et al. / China Ocean Eng., 30(6), 2016, 954 – 966

1982) of semi-sphere is applied. As an example, the semi-sphere with the above geometries and water
depth is considered here. Table 1 shows the comparison between the wave forces and hydrodynamic
coefficients in heave calculated by the present method and the traditional analytical method. The
numerical results show a good agreement which can be regarded as the validation of the present
method.
Table 1 Verification of the solution method
Frequency (rad/s) 1.0 1.5 2.0
Method Present Analytical Present Analytical Present Analytical
fd3/(gAR2) 0.9925 0.9895 0.3736 0.3745 0.1388 0.1367
33/(R3) 0.4880 0.4903 0.2311 0.2297 0.0756 0.0763
33/(R3) 0.9058 0.9042 0.5676 0.5667 0.4668 0.4701

5. Wave Energy Conversion Performance

The wave energy conversion ability of oscillating-body WEC is mainly decided by the floater’s
geometries and physical parameters such as the PTO damping coefficients. To explore the geometries’
effect on the wave energy conversion performance, two groups of vertical axisymmetric floaters are
examined based upon the former verified hydrodynamic formulations, corresponding to draught of 5 m
and displacement of 392.5 m3 with the same radius of 5 m in the water depth of 50 m. The descriptions
of the geometric shapes for the two kinds are presented in Table 2.

Table 2 Description of the bottom shape of the floaters


Case No. 1 2 3 4
Bottom shape diagram
Bottom shape description Cylindrical Hemispherical Paraboloidal Conical
Draught d=5 m t=d=5 m, R=5 m t=0 m, R=5 m t=0 m, R=5 m t=0 m, R=5 m

Displacement 392.5 m3 t=1.67 m, R=5 m, t=0 m, R=5 m, t=0 m, R=5 m,


t=d=5 m, R=5 m
d=6.67 m d=7.07 m d=15 m

5.1 Power Estimation for Absorbers with the Same Draught


By considering the geometrical parameters for the first kind listed in Table 2, the absorbers are of
basic configuration. The hydrodynamic characteristics of the former two figures can be evaluated with
the analytical method, while for the latter two the introduced semi-analytical solution can be employed.
According to Falnes (2002), for an oscillating-body WEC, the captured wave power can be defined by
P=0.5Cpto223 . Here, Cpto denotes the damping coefficient in the power take-off mechanism and 3
represents the motion amplitude in heave in the incident wave with frequency .
Figs. 4 and 5 give a clear presentation on the heave RAO and captured wave power of the
absorbers for varying PTO damping coefficients. It can be seen that the heave RAO is reduced at the
peak frequency by increasing PTO damping. Just as the conclusion given by Koh et al. (2014), the
damped heave peak frequencies are hardly affected by the PTO damping. Further, the peak frequency
is the same as that of the heave RAO, at which, the captured wave power is promoted with the increase
of PTO damping. What is more important is that, for a given lower PTO damping, the cylindrical
ZHANG Wan-chao et al. / China Ocean Eng., 30(6), 2016, 954 – 966 961

absorber shows better wave energy conversion ability at the peak frequency than the others do. However,
the frequency bandwidth, in which the wave captured performance is considerable, of the cylinder is
far less than that of the parabolic and conical absorbers. In addition, to a great extent, the PTO
damping affects the frequency bandwidth. In such case, to synthetically evaluate the wave energy
conversion ability of each absorber with the same draught, the captured wave power for each absorber
in the frequencies ranging of 1.02.0 rad/s with their own optimum frequency-dependent PTO
damping coefficients should be considered.

Fig. 4. Heave RAOs with the frequency for absorbers with different PTO damping coefficients; Cpto-1 Cpto-5 indicate
Cpto=0.1, 0.5, 1.0, 2.0, and 3.0 (104 kg/s) respectively; (a) Cylindrical; (b) Hemispherical; (c) Paraboloidal; (d) Conical.

Fig. 5. Captured wave power with the frequency for absorbers with different PTO damping coefficients; Cpto-1Cpto-5 indicate
Cpto=0.1, 0.5, 1.0, 2.0, and 3.0 (104 kg/s) respectively; (a) Cylindrical; (b) Hemispherical; (c) Paraboloidal; (d) Conical.

Fig. 6 shows the variation of the captured energy with the damping coefficients for different
absorbers. As mentioned before, the optimum PTO damping coefficients highly depend on the
frequency and show high mobility for different absorbers. Further, it is found that the parabolic and
962 ZHANG Wan-chao et al. / China Ocean Eng., 30(6), 2016, 954 – 966

conical absorbers also show good wave energy performance for a given frequency when the PTO
damping is optimal. Au contraire, at a given higher frequency, the cylindrical absorber shows poor
performance.
Fig. 7 depicts the captured energy for absorbers with their respective optimal frequency-
dependent PTO damping coefficients. Here we can easily obtain the optimum figure of the absorbers
with the same draught. As shown in Fig. 7, the cylindrical absorber shows excellent performance at the
frequencies ranging of 0.91.2 rad/s, while in the range of 1.2 to 1.5 rad/s, the semispherical absorber
will be better and at higher frequencies, and the paraboloidal and conical absorbers will be fine.
However, in general, the conical absorber shows better applicable performance in wave energy
conversion at various frequencies.

Fig. 6. Captured wave power with PTO damping coefficients for absorbers at different wave frequencies (rad/s); (a) Cylindrical;
(b) Hemispherical; (c) Paraboloidal; (d) Conical.

Fig. 7. Captured wave power with frequency for absorbers with


their optimum frequency- dependent PTO damping
coefficients.

5.2 Power Estimation for Absorbers with the Same Displacement


To estimate the wave energy conversion ability of the absorbers when the outer radius and
displacement are given, we set the four cases of absorbers with the same displacement of 392.5 m3 as
listed in Table 2, in which the absorber of Case 3 contains a cylinder and a semispherical cap at the
bottom. Similarly, the peak frequency performance of the heave RAO and captured power are the same
as before, which can be obtained from Figs. 8 and 9. At higher frequencies, the heave motions of the
cylindrical and semispherical absorbers approach zero quickly. Compared with the condition of the
ZHANG Wan-chao et al. / China Ocean Eng., 30(6), 2016, 954 – 966 963

same draught, the wave energy conversion abilities of the latter three absorbers with the same
displacement are better and the peak frequencies are smaller.

Fig. 8. Heave RAOs with the frequency for absorbers with different PTO damping coefficients; Cpto-1Cpto-5 indicate Cpto=0.1,
0.5, 1.0, 2.0, and 3.0 (104 kg/s) respectively; (a) Cylindrical; (b) Hemispherical; (c) Paraboloidal; (d) Conical.

Fig. 9. Captured wave power with frequency for absorbers with different PTO damping coefficients; Cpto-1Cpto-5 indicate
Cpto=0.1, 0.5, 1.0, 2.0, and 3.0 (104kg/s) respectively; (a) Cylindrical; (b) Hemispherical; (c) Paraboloidal; (d) Conical.

As shown in Fig. 10, all the optimum PTO damping coefficients of the four absorbers decrease
first and then increase with the increase of the frequency. In addition, the parabolic and conical
absorbers seem to be better than the cylindrical absorber in wave energy conversion when their
damping coefficients are optimum. At relative high frequencies, the captured wave powers of the two
absorbers are really small, which means that the parabolic and conical absorbers have better
applicability.
Fig. 11 gives a comprehensive comparison of the wave energy conversion performance among the
four absorbers. At the frequencies of 1.051.17 rad/s, the cylindrical absorber is better than the other
three. At higher frequencies, the performances of the latter two are almost the same. However, at the
frequencies smaller than 1.05 rad/s, the parabolic absorber shows a good potential in wave energy
964 ZHANG Wan-chao et al. / China Ocean Eng., 30(6), 2016, 954 – 966

conversion. In such case, the parabolic absorber will be a good choice when the outer radius and
displacement are given in the whole frequency range.

Fig. 10. Captured wave power with PTO damping coefficients for absorbers at different wave frequencies (rad/s); (a) Cylindrical;
(b) Hemispherical; (c) Paraboloidal; (d) Conical.

Fig. 11. Captured wave power with frequency for absorbers with
their optimum frequency-dependent PTO damping
coefficients.

6. Conclusions

Four types of absorbers with the same radius and same draught or displacement are examined to
analyze the geometries’ effect on the wave energy conversion ability, in which a semi-analytical
method of decomposing the vertical axisymmetric curved surface into several ring shaped step surfaces
is introduced and examined. Combined with the updated body boundary characteristics, by using
eigenfunction expansion matching method, the expressions of velocity potential in each domain, added
mass, radiation damping coefficients and wave exciting forces of the oscillating buoy were obtained.
The calculation results show that:
(1) The semi-analytical method by which the vertical axisymmetric curved surface is decomposed
into several ring shaped step surfaces can calculate accurately and investigate systemically the
geometries’ effect on the hydrodynamic characteristics.
(2) For a wave energy absorber with given outer radius and draught, at some given frequencies,
the cylindrical absorber has fantastic wave power capturing performance, while in the whole frequency
range, the conical absorber has better adaptability and flexibility. When the outer radius and
displacement of the absorber are determined, whether the frequency is low or high, the parabolic
absorber shows better wave energy conversion ability than the other three types of absorbers.
ZHANG Wan-chao et al. / China Ocean Eng., 30(6), 2016, 954 – 966 965

(3) When the frequency of the working sea area is determined, by accommodating the geometrical
parameters and the resonance frequency, the cylindrical absorber will output wave energy efficiently.
However, in the random wave sea, the parabolic and conical absorbers have better stabilization in wave
power conversion.

Appendix A
Γ 0 ( z )  cos(k0 h ) cosh  k0 ( z  h ) [2k0 h  sinh(2k0 h )] , Γ n ( z )  cos  kn ( z  h ) [2kn h  sin(2kn h )],
 DlE ( r, z )  cosh  k0 ( z  h ) cosh(k0 h)  l i l J l (k0 r ) , 11E ( r, z )  30E ( r, z )  51E ( r, z )  0,
 Dl ( r, z )  11 ( r, z )  0 , 30 ( r, z )  [2( z  h )2  r 2 ] (4hp ) , 51 ( r, z )   r[( z  h )2  r 2 4] (2hp h ),
Ip Ip Ip Ip

GDI1l ( r )  ( r R1 )l , G11I1 ( r )  r r1 , G30I1 ( r )  1 , G51I1 ( r )  r r1 ,


ln( r Rp 1 ) ln( Rp Rp 1 )
G jlp ( r )  
I

[( r Rp 1 )  ( Rp 1 r ) ] ( Rp Rp 1 )  ( Rp 1 Rp )
l l l l

ln( Rp r ) ln( Rp Rp 1 )
G jlp ( r )  
I

[( Rp r )  ( r Rp ) ] [( Rp Rp 1 )  ( Rp 1 Rp ) ]
l l l l

Qn ppl ( r )  [K l (n p , p Rp 1 )I l (n p , p r )  K l (n p , p r )I l (n p , p Rp 1 )] Π ,


I

Q n ppl ( r )  [K l (n p , p r )I l (n p , p Rp )  K l (n p , p Rp )I l (n p , p r )] Π ,


I

and
n  n p π h p , Π  K l (n , p R p 1 )I l (n , p R p )  K l (n , p R p )I l (n , p R p 1 ).
p p p p p

References
Alves, M., Traylor, H. and Sarmento, A., 2007. Hydrodynamic optimization of a wave energy converter using a
heave motion buoy, Proceedings of the 7th European Wave and Tidal Energy Conference, Porto, Portugal.
Babarit, A., Hals, J., Muliawan, M. J., Kurniawan, A., Moan, T., and Krokstad, J., 2012. Numerical benchmarking
study of a selection of wave energy converters, Renew. Energy, 41, 4463.
Bachynski, E. E., Young, Y. L. and Yeung, R. W., 2012. Analysis and optimization of a tethered wave energy
converter in irregular waves, Renew. Energy, 48, 133145.
Colby, M., Nasroullahi, E. and Tumer, K., 2011. Optimizing ballast design of wave energy converters using
evolutionary algorithms, Proceedings of the 13th Annual Genetic and Evolutionary Computation Conference,
Dublin, 17391746.
Falcão, A. F. O., Henriques, J. C. C. and Cândido, J. J., 2012. Dynamics and optimization of the OWC spar buoy
wave energy converter, Renew. Energy, 48, 369381.
Falnes, J., 2002. Ocean Waves and Oscillating Systems: Linear Interactions Including Wave-Energy Extraction,
Cambridge University Press, Chapter V.
Goggins, J. and Finnegan, W., 2014. Shape optimization of floating wave energy converters for a specified wave
energy spectrum, Renew. Energy, 71, 208220.
Gomes, R. P. F., Henriques, J. C. C., Gato, L. M. C. and Falcão, A. F. O., 2012. Hydrodynamic optimization of an
axisymmetric floating oscillating water column for wave energy conversion, Renew. Energy, 44, 328339.
Hulme, A., 1982. The wave forces acting on a floating hemisphere undergoing forced periodic oscillations, J.
Fluid Mech., 121, 443463.
966 ZHANG Wan-chao et al. / China Ocean Eng., 30(6), 2016, 954 – 966

Koh, H. J., Ruy, W. S., Cho, I. H. and Kweon, H. M., 2014. Multi-objective optimum design of a buoy for the
resonant-type wave energy converter, J. Mar. Sci. Technol., 20(1): 5363.
Kokkinowrachos, K., Mavrakos, S. A. and Asorakos, S., 1986. Behavior of vertical bodies of revolution in waves,
Ocean Eng., 13(6): 505538.
Kramer, M. M. and Frigaard, P. B., 2002. Efficient wave energy amplification with wave reflectors, Proceedings
of the 12th International Offshore and Polar Engineering Conference, Kitakyushu, Japan.
Mavrakos, S. A., 2004. Hydrodynamic coefficients in heave of two concentric surface-piercing truncated circular
cylinders, Appl. Ocean Res., 26, 8497.
Mavrakos, S. A. and Katsaounis, G. M., 2009. Effects of floaters’ hydrodynamics on the performance of tightly
moored wave energy converters, IET Renew. Power Gen., 4(6): 531544.
Mavrakos, S. A., Katsaounis, G. M. and Apostolidis, M. S., 2009. Effects of floaters’ geometry on the performance
characteristics of tightly moored wave energy converters, Proceedings of the 28th International Conference
on Ocean Offshore Arctic Engineering, ASME, Honolulu, Hawaii, Paper No. OMAE 2009-80133.
McCabe, A. P., Aggidis, G. A. and Widden, M. B., 2010. Optimizing the shape of a surge-and-pitch wave energy
collector using a genetic algorithm, Renew. Energy, 35(12): 27672775.
McCabe, A. P., 2013. Constrained optimization of the shape of a wave energy collector by genetic algorithm,
Renew. Energy, 51, 274284.
Oskamp, J. A. and Ozkan-Haller, H. T., 2012. Power calculations for a passively tuned point absorber wave energy
converter on the Oregon coast, Renew. Energy, 45, 7277.
Yeung, R. W., 1980. Added mass and damping of a vertical cylinder in finite depth waters, Appl. Ocean Res., 3(3):
119133.

You might also like