Вы находитесь на странице: 1из 19

Dynamics of dual pontoon oating structure for cage aquaculture

in a two-dimensional numerical wave tank


Hung-Jie Tang
a
, Chai-Cheng Huang
b,n
, Wei-Ming Chen
c
a
Tainan Hydraulics Laboratory, National Cheng Kung University, Tainan 70955, Taiwan
b
Department of Marine Environment and Engineering, National Sun Yat-sen University, Kaohsiung 80424, Taiwan
c
Ocean Technology Department, Industrial Technology Research Institute, Hsinchu 31040, Taiwan
a r t i c l e i n f o
Article history:
Received 21 February 2010
Accepted 14 June 2011
Available online 27 July 2011
Keywords:
Dynamic response
Dual pontoon oating structure (DPFS)
Numerical wave tank (NWT)
Boundary element method (BEM)
a b s t r a c t
The trend of using oating structures with cage aquaculture is becoming more popular
in the open sea. The purpose of this paper is to investigate the dynamic properties of a
dual pontoon oating structure (DPFS) when attached to a sh net by using physical
and numerical models. A two-dimensional (2-D) fully nonlinear numerical wave tank
(NWT), based on the boundary element method (BEM), is developed to calculate the
wave forces on the DPFS. The wave forces on a sh net system are then evaluated using
a modied Morison equation. The comparisons of dynamic behaviors between numer-
ical simulations and experimental measurements on the DPFS show good agreement.
Results also display that a sh net system causes the resonant response of body motions
and mooring forces to be slightly lower due to the nets damping effect. Finally, for
designing the rearing space of cage aquaculture, the inuences which net depth and net
width have on the DPFS dynamic responses are also presented in this paper.
Crown Copyright & 2011 Published by Elsevier Ltd. All rights reserved.
1. Introduction
In recent years, the development of offshore cage culture has attracted wide attention. The central spar sh cage
(Fredriksson et al., 2003), tension-leg type sh cage (Lee and Wang, 2005), and gravity-type cage (Huang et al., 2007) are
some examples. On the other hand, the dual pontoon oating structure (DPFS) has become popular due to its versatile
features, such as using wave barriers, a catamaran (twin hull boat), or a device for raising sh. The proposed DPFS for cage
aquaculture may consist of a dual pontoon with a large sh net hanging between them. The deck of each pontoon offers
space for a control room, automatic feeding machine, crane, rearing or harvest equipments, etc. All of these facilities are
essential for marine culture in the open sea.
The hydrodynamic properties of the DPFS were investigated by Williams and Abul-Azm (1997), using a boundary
integral equation method. They found that the wave reection properties of the structure depend strongly on the width,
draft and spacing of the pontoons and mooring line stiffness. The pre-tension in the mooring system is less important.
Weng and Chou (2007) applied a boundary element method (BEM) and a physical model to examine the dynamic
responses of a oating dual pontoon structure. They discovered that the clear space between pontoons had a signicant
inuence on the responses of the structure and created an extra peak response in heave motion at a high frequency.
Although the results offered by both articles give valuable insights into the dynamic responses of a dual pontoon structure,
Contents lists available at ScienceDirect
journal homepage: www.elsevier.com/locate/jfs
Journal of Fluids and Structures
0889-9746/$ - see front matter Crown Copyright & 2011 Published by Elsevier Ltd. All rights reserved.
doi:10.1016/j.juidstructs.2011.06.009
n
Corresponding author. Tel.: 886 7 525 5169; fax: 886 7 525 5060.
E-mail addresses: hjtang@thl.ncku.edu.tw (H.-J. Tang), cchuang@mail.nsysu.edu.tw (C.-C. Huang), zp7689@hotmail.com (W.-M. Chen).
Journal of Fluids and Structures 27 (2011) 918936
further application to cage aquaculture remains unknown and should be investigated. Besides, these two studies were
limited to linear wave theory.
To solve the nonlinear wave problem, Longuet-Higgins and Cokelet (1976) introduced the Mixed Eulerian and
Lagrangian method (MEL). By this method, the fully nonlinear boundary conditions on the free water surface can be
satised instantaneously. Hereafter, the time-domain approach of the nonlinear wavebody interaction using the BEM has
become an extremely successful scheme. For instance, Isaacson (1982) used this scheme to simulate the nonlinear wave
effects on xed and oating structures with three-dimensional (3-D) arbitrary shapes. After the numerical damping zone
had been proposed (Cointe, 1990; Tanizawa, 1996), researchers started to take this advantage to develop the numerical
wave tank (NWT) for long period simulations. Over the last two decades, the BEM-based NWT has become one of the most
popular tools in ocean engineering research, due to its highly computational efciency, e.g. Cointe (1990, 2-D), Tanizawa
(1996, 2-D), Contento (2000, 2-D), Boo (2002, 3-D), Koo and Kim (2004, 2-D), and Bai and Eatock Taylor (2006, 2007, 3-D).
The review of 2-D and 3-D NWT developments and their applications can be found in detail in Kim et al. (1999) and
Tanizawa (2000). The descriptions about time-stepping technique, re-gridding method, numerical damping zone, and
wavebody interaction problems etc., were also presented in these works.
However, the fully nonlinear simulation for a freely oating body is still considered as a very challenging problem, since
it requires the treatment of the coupling between hydrodynamic forces and body motions. To precisely obtain wave forces
on a body, a highly accurate calculation of the time derivative of velocity potential f
t
is required for the unsteady Bernoulli
equation. Tanizawa (1995) introduced the acceleration potential and derived the exact body boundary condition for the
acceleration eld. This was known as the acceleration potential method. Koo and Kim (2004) used this method to
investigate the fully nonlinear wavebody interactions for a freely oating barge-type structure in a 2-D NWT. On the
other hand, Wu and Eatock Taylor (1996) proposed another technique to solve the velocity and acceleration eld
separately. The merit of this method is in computing faster than that of Tanizawa, while the disadvantage is that it cannot
compute the pressure distribution on the body surface unless it solves another boundary value problem for the auxiliary
functions, details referred to Kim et al. (1999). Bai and Eatock Taylor (2009) used this method to simulate the fully
nonlinear wave interactions with xed and freely oating ared structures in a 3-D NWT. The above BEM-based methods
were shown to be robust and stable, and thus are suitable to deal with oating body problems. Recently, Yan and Ma
(2007) and Ma and Yan (2009) developed a quasi arbitrary LagrangianEulerian nite element method (QALE-FEM) for 2-D
and 3-D oating body problems. In thus method, an improved iterative procedure, called iterative semi-implicit time
integration method for oating body (ISITIMFB) is incorporated to deal with the fully nonlinear interaction problems
between steep waves and oating bodies, from which the numerical results are identied in good agreement with other
published data.
In recent years, the fully nonlinear wavebody interactions for multiple oating structures (2-D) have been investigated
by many researchers. For example, Koo and Kim (2007) studied the shielding effect and the pumping/sloshing modes of a
water column with various gap distances for xed double barges in a 2-D NWT. Koo (2009) simulated wave blocking and
wave energy absorption of a pneumatic-type oating breakwater in a 2-D NWT. Huang and Tang (2009) developed a 2-D
NWT to investigate the dynamic response of body motion and tension force for a moored DPFS. Sun et al. (2010)
investigated the hydrodynamic interactions between waves and two parallel closely spaced rectangular barges, to
characterize the general problem of liquid natural gas (LNG) ofoading from a oating plant into a shuttle tanker.
In the present study, we applied the 2-D NWT to investigate the dynamic response of the DPFS for cage aquaculture,
based on our previous studies, Huang et al. (2008) and Huang and Tang (2009), while the calculation of wave loadings on
aquaculture cage nets could be found in Loland (1991), Huang et al. (2006), and Moe et al. (2010). In this paper, the
boundary integral equation (BIE) is formulated by using a linear element method. Subsequently, the nonlinear free surface
is traced by the MEL technique, and the 4th order RungeKutta method (RK4). Meanwhile a node-regridding and
smoothing technique is applied by using a cubic spline scheme to prevent free surface nodes from moving too close to each
other. Two numerical damping zones proposed by Cointe (1990) and Tanizawa (1996) are used at both ends of the
numerical wave tank to absorb or dissipate the reected and transmitted waves induced by wavebody interaction. The
immediate hydrodynamic force is calculated by an acceleration potential method (Tanizawa, 1995) and a modal
decomposition method (Vinje and Brevig, 1981). The wave forces on the sh net are evaluated by using a modied
Morison equation (Brebbia and Walker, 1979) in order to deal with uid and net relative motion problems.
In this paper, both numerical and physical models are adopted to investigate the dynamic properties of the DPFS with
and without a sh net. The rst- to third-order of sway, heave, roll, and sea-side tension RAO (response amplitude
operator) of the DPFS with or without a sh net are investigated. Finally, for designing the rearing space of cage
aquaculture, the inuences of net depth and net width on dynamic responses are also discussed.
2. Numerical model
A DPFS, which consists of a pair of oating rectangular pontoons, is restrained by a linear symmetric mooring system, as
shown in Fig. 1, where a is the width of each pontoon, b is the spacing between two pontoons, d is the draft, (x
G
, z
G
) is the
position of the gravity center, l
G
is the roll moment arm, y
0
is the mooring line angle, and l
0
is the original length of
mooring line. The oating structure is deployed in a numerical wave tank with a constant water depth, h. A numerical
H.-J. Tang et al. / Journal of Fluids and Structures 27 (2011) 918936 919
damping zone is used at each end of the wave tank to dissipate reected and transmitted waves, where x
d1
and x
d2
are the
entrance positions.
2.1. Governing equation
The ow eld is assumed to be incompressible, inviscid, and the motion irrotational. Thus, a velocity potential exists
and satises Laplace equation,
r
2
f0: 1
Incorporating Eq. (1) into the Green second identity, the velocity potential in the uid domain can be determined by
solving the following BIE:
af
i

_
G
j
@G
ij
@n
f
j
G
ij
@f
j
@n
_ _
dG
j
, 2
where G
ij
lnr
ij
=2p is the fundamental solution to the Laplace equation and represents a ow eld generated by a
concentrated unit source acting at ith source point. r
ij
is the distance from source point (x
i
, z
i
) to eld point (x
j
, z
j
) and a is
the internal solid angle between two boundary elements. In this model, the linear element scheme and 6-point Gaussian
quadrature integration method are applied to solve the BIE.
2.2. Inow boundary condition
Based on the continuity of velocity, a theoretical particle velocity prole can be used to specify the boundary value
along the inow boundary. For nonlinear regular waves, the second-order Stokes wave is used to prevent the mismatch
between input velocity proles and real water particle velocity, as described in Koo and Kim (2004)
@f
@n

gAk
s
coshkzh
coshkh
coskxst

3
4
A
2
ks
cosh2kzh
sinh
4
kh
cos2kxst
_

_
_

_
f
m
on G
I
, 3
where A, k, and s are the amplitude, wave number, and angular frequency, respectively. g is the gravitational acceleration
and t is the time. The modulation function f
m
is used to prevent impulse-like behavior of a wave maker, and it is written as
f
m
t
1
2
1cos
pt
T
m
_ _ _ _
for t oT
m
,
1 for t ZT
m
,
_

_
4
where T
m
is the modulation duration which depends on wave steepness. For a steeper wave, the modulation duration is
usually twice as long as a regular wave period.
2.3. Free surface boundary condition
One of the most popular and successful approaches to the fully nonlinear free surface simulation is the MEL method,
which was rst presented by Longuet-Higgins and Cokelet (1976). In this method, the kinematic and dynamic free surface
boundary conditions are transformed into the Lagrangian frame. To obtain numerical solutions for wave propagation in a
Mooring line
DPFS
Fish net

l
3L
l
a
a
b
d
h
x
z
x
3L 2L 2L
d
Fig. 1. Denition sketch of a DPFS with a sh net.
H.-J. Tang et al. / Journal of Fluids and Structures 27 (2011) 918936 920
wave tank, the scheme used numerical damping zones at both ends of the wave tank to absorb the transmitted wave
energy at the end of the tank as well as to dissipate the reected waves in front of the input boundary. Following Cointe
(1990) and Tanizawa (1996), the numerical damping zones are incorporated into the free surface boundary conditions as
dx
dt

@f
@x
dz
dt

@f
@z
nxzz
e

df
dt
gz
1
2
9rf9
2
nxff
e

on G
f
,
_

_
5
where n(x) is the damping coefcient of numerical damping zone given by
nx
a
d
sx
d1
x=L
2
xrx
d1
,
a
d
sxx
d2
=L
2
xZx
d2
,
0 otherwise:
_

_
6
a
d
is the dimensionless parameter for the strength of the damping zone, and after several tests we found that a
d
set to 1 is
adequate for accurate results. L is the wavelength of the input wave, x
d1
and x
d2
are the entrance positions of each damping
zone shown in Fig. 1. z
e
and f
e
in Eq. (5) are the entrance wave elevation and potential function in the front damping zone,
only existing in xrx
d1
. Tanizawa (1996) has applied this damping zone technique to dissipate the wave energy reected
from the structure, yet without disturbing the outgoing incident waves. For practical purposes, the nonlinear analytical
solution of the second-order Stokes wave theory is adopted in the damping zone 1 to improve the computational process.
In this model, the entrance potential and wave elevation are written as
f
e

Ag
s
coshkzh
coshkh
sinkxst
3
8
A
2
s
cosh2kzh
sinh
4
kh
sin2kxst,
z
e
Acoskxst
kA
2
coshkh
4sinh
3
kh
2cosh2khcos2kxst:
_

_
7
Additionally, the nodal velocities in Eq. (5) are obtained by using the cubic spline scheme in the curvilinear coordinate
system as described in Section 2.6. The corner problem between free surface and body surface is treated according to Grilli
and Svendsen (1990) as described in Section 2.7.
2.4. Body surface boundary condition
In this model, the body surface (G
s
) is impermeable. Therefore, the uid velocity is equal to the normal velocity on the
body surface
@f
@n
n
1
_ x
G
n
2
_
z
G
n
3
_
y
G
on G
s
, 8
where (n
1
,n
2
,n
3
)(n
x
,n
z
,r
z
n
x
r
x
n
z
)is the unit normal vector on the body boundary; (r
x
,r
z
)(xx
G
,zz
G
) is the position vector
from body surface to gravity center. Subscript G designates the gravity center of the body; _ x
G
,
_
z
G
are the translational
velocities in the x and z axes (sway and heave motions) and
_
y
G
is the angular velocity about the y axis (roll motion).
2.5. Rigid boundary condition
At the end-wall (G
w
) and bottom (G
b
) of the wave tank, the boundary conditions are considered impermeable. The
normal velocities are then set to zero:
@f
@n
0 on G
b
and G
w
: 9
2.6. Curvilinear coordinate system
In this paper a curvilinear coordinate system and the cubic spline scheme are adopted to solve the spatial derivatives of
velocity potential in Eq. (5) on the free surface boundary. The relationship between the velocity components in Cartesian
and curvilinear coordinates is written as
@f
f
@x

@f
f
@s
cosb
f

@f
f
@n
sinb
f
,
@f
f
@z

@f
f
@s
sinb
f

@f
f
@n
cosb
f
,
_

_
10
H.-J. Tang et al. / Journal of Fluids and Structures 27 (2011) 918936 921
where f
f
represents the potential function on the free surface, and b
f
is the angle between s, a section of free surface, and
the x axis. The normal velocities of the free surface qf
f
/qn are obtained after solving the BIE, and the angle b
f
is determined
from the following equation:
tanb
f

sinb
f
cosb
f

@z=@s
@x=@s
, 11
where qf
f
/qs, qx/qs, and qz/qs are calculated by using cubic spline interpolation in curvilinear coordinates along the free
surface.
Once the values of the time derivative of the potential function on the right side of Eq. (5) are known, the substantial
derivative equations on the left side can be used to predict the new nodal position and its corresponding potential on the
free surface boundary by employing the RK4 method as a time marching scheme. This process was repeated until the
simulation reached the steady-state condition.
Note that the node-regridding and smoothing technique is also applied in the present model by using the cubic spline
interpolation on the curvilinear coordinate system, to prevent free surface nodes from moving too close to each other, and
to prevent the saw-tooth condition occurring which may lead to the numerical instability.
2.7. Corner problem between free surface and body surface
At the intersection of body surface and free surface, the discontinuity of ux occurs due to the discontinuity of normal
direction. Although the cubic spline scheme is accurate to determine the tangential slope at the end-point with the nature
condition (curvature equal to zero) and the Lagrangian polynomial method, the requirement of continuity of ux at the
corner is still difcult to meet. To deal with this discontinuity, the double collocation node technique is often used. Grilli
and Svendsen (1990) proposed that a treatment for the corner problem at the intersection based on the continuity ux is
@f
f
@s

@f
f
@n
cosb
b
b
f

sinb
b
b
f

@f
b
@n
1
sinb
b
b
f

, 12
where the subscripts b and f denote the body and water free surface, respectively. @f
b
=@n and @f
f
=@n are the normal
velocities on the free surface and body surface, respectively; @f
f
=@s is the modied tangential velocity on the free surface
and will be used in Eq. (10) when dealing with corner problem. In this model, the input boundary angle at the front of tank
is b
b
p=2, while that of the wall boundary at the end of wave tank is b
b
3p=2.
2.8. Wave forces on the body
The hydrodynamic forces on the body can be calculated by integrating the pressure around the wetted body surface as
F
_
Gs
r f
t
gz
1
2
9rf9
2
_ _
nds,
M
_
Gs
r f
t
gz
1
2
9rf9
2
_ _
r nds,
_

_
13
where F and M are the hydrodynamic force and moment, r is the water density, n is the unit normal vector on the body
surface and points into the body, and r is the position vector from gravity center to body surface.
2.9. Acceleration potential method
In order to evaluate the hydrodynamic forces on the oating body using Eq. (13), both the gradients of velocity
potential (rf) and the unsteady f
t
terms on the wetted body surface must be determined beforehand. rf is evaluated by
the regular BEM, whilef
t
is determined by an acceleration potential method proposed by Tanizawa (1995), taking the
advantage of the feature that f
t
also satises the Laplace equation,
r
2
f
t
0: 14
Following Tanizawa (1995), the body-surface boundary condition in the acceleration eld is described as
@f
t
@n
n
1
x
G
n
2

z
G
n
3

y
G
q on G
s
, 15
where x
G
,

z
G
are the translational accelerations in the x-and z-axes and

y
G
is the angular acceleration around y-axis. q is
dened as
q n
1
_
y
G
r
x
_
y
G
2
_
z
G
2
@f
@z
_ _
n
2
_
y
G
r
z
_
y
G
2_ x
G
2
@f
@x
_ _
k
n
@f
@x
_ x
G

_
y
G
r
z
_ _
2

@f
@z

_
z
G

_
y
G
r
x
_ _
2
_ _
k
n
@f
@x
_ _
2

@f
@z
_ _
2
_ _

@f
@s
@
2
f
@n@s

@f
@n
@
2
f
@s
2
, 16
H.-J. Tang et al. / Journal of Fluids and Structures 27 (2011) 918936 922
where k
n
1/r
n
is the normal curvature along the s direction of the body surface and r
n
is the radius of curvature. Note that
@f=@s, q
2
f/(qs qn), and @
2
f=@s
2
are calculated by using cubic spline interpolation in the curvilinear coordinates along the
body surface. The above variables at the corners of the structure are considered as the natural condition (curvature equal
to zero), and those between structure and free surface are modied by the same method as described in Section 2.7.
For solving the f
t
in the acceleration eld, four methods are available, these including: (1) iterative method (Cao et al.,
1994; Yan and Ma, 2007), (2) modal decomposition method (Vinje and Brevig, 1981), (3) implicit boundary condition
method (Tanizawa, 1995), and (4) indirect method (Wu and Eatock Taylor, 1996). Detailed description for these methods
can be found in Tanizawa (2000), Koo and Kim (2004), and Ma and Yan (2009).
Following our previous study (Huang et al., 2008; Huang and Tang, 2009), the modal decomposition method with fully-
updated RK4 scheme are adopted in this model. This scheme has been proved to be accurate and robust (see Koo and Kim,
2004; Huang et al., 2008). For instance, the average CPU time for a case examined in Section 4.1 of the present paper is
about 68 min on a normal PC with a Pentium IV 3.0 GHz processor and 2G RAM.
2.10. Modal decomposition method
The modal decomposition method was rst introduced by Vinje and Brevig (1981). This method solves the BIE for the
acceleration eld. The acceleration potential function is decomposed into four modes corresponding to three unit
accelerations for sway-heave-roll motions (radiation problem) and the acceleration due to incident wave eld (diffraction
problem). Using these four modes in Eq. (17) and the boundary conditions listed in Eqs. (18)(21), the unknown modes
amplitude can be determined by solving its respective BIE. The f
t
is given by
f
t

3
m 1
a
m
j
m
j
4
, 17
where a
m
is the mth mode component of generalized body acceleration (1sway, 2heave, 3roll, 4diffraction mode).
The boundary conditions in the acceleration eld for each mode are given as
@j
m
@n

n
m
m1,2,3
q m4
on G
s
,
_
18
j
m

0 m1,2,3
gz
1
2
9rf9
2
m4
on G
f
,
_
19
@j
m
@n
0 m1,2,3 on G
I
, 20
@j
m
@n
0 m1 $4 on G
b
and G
w
: 21
The inow boundary condition for mode 4 is obtained from the 2nd-order Stokes wave theory,
@j
4
@n

gAk
coshkzh
coshkh
sinkxst

3
2
A
2
ks
2
cosh2kzh
sinh
4
kh
sin2kxst
_

_
_

_
on G
I
: 22
After solving the BIE for each mode, the values of j
m
and @j
m
=@n on all the boundaries are obtained, and the remaining
unknown in Eq. (17) is a
m
.
Substituting Eq. (17) into Eq. (13), combined with Newtons second law including the wave hydrodynamic forces and
other forces (with subscript e; e.g. gravitational force, damping force, restoring force, and force on net), the equation of
motion becomes
ma
1

_
Gs
ra
1
j
1
a
2
j
2
a
3
j
3
j
4
gz
1
2
9rf9
2
n
1
dsF
ex
,
ma
2

_
Gs
ra
1
j
1
a
2
j
2
a
3
j
3
j
4
gz
1
2
9rf9
2
n
2
dsF
ez
,
I
G
a
3

_
Gs
ra
1
j
1
a
2
j
2
a
3
j
3
j
4
gz
1
2
9rf9
2
n
3
dsM
ey
:
_

_
23
After solving Eq. (23), the generalized acceleration a
m
can be obtained. The details of extra forces are described below.
2.11. Mooring force
The mooring system is considered linear and symmetric (see Fig. 1) with the spring constant K, and the
wave hydrodynamic forces on the mooring line is negligible as compared to the force on the oating dual pontoon.
H.-J. Tang et al. / Journal of Fluids and Structures 27 (2011) 918936 923
The pre-tension force of this mooring system is written as
F
T
0

2rgadlmg
2siny
0
, 24
where l is the total length of the oating dual pontoon structure in the direction of y.
2.12. Wave forces on sh net
In this model, a sh net cage is set between the pontoons and secured by a steel frame. The net is assumed not to
deform. By applying the lumped mass method (Huang et al., 2006), the sh net panel is divided into several elements and
nodes, as shown in Fig. 2, and a modied Morison equation (Brebbia and Walker, 1979) calculates the drag and inertia
forces on the net elements is given as
F
net

1
2
rC
D
A
net
V
R
9V
R
9r8
net
C
M
dV
dt
r8
net
K
M
d
_
R
dt
, 25
where r is water density, C
D
is the drag coefcient, C
M
1K
M
is the inertia coefcient, and K
M
is the added mass
coefcient. In this model, C
M
2.0 is assumed, which is generally between 1.0 and 2.0. A
net
is the projected area of the net
element, 8
net
is the volume of the net element, and V
R
V
_
R is the relative velocity between ow eld and net element. V
is the ow velocity at the center of a net element,
_
R is the central velocity of the net element, dV/dt is the uid particle
acceleration at the center of the net element, and d
_
R=dt is the central acceleration of the net element. Details about uid
particle velocity and acceleration are described in the appendix.
According to Lolands empirical formula (Loland, 1991), the drag force which is parallel to the uid motion, and the lift
force which is perpendicular to the uid motion, are given as follows:
F
D

1
2
rC
D
aA
net
9V
R
9
2
,
F
L

1
2
rC
L
aA
net
9V
R
9
2
,
_
_
_
26
where C
D
(a) and C
L
(a) are coefcients related to the angle of a between the uid particle velocity vector and the normal
vector of the net element (see Huang et al., 2006):
C
D
a 0:040:040:33S
n
6:54S
n
2
4:88S
n
3
cosa,
C
L
a 0:05S
n
2:3S
n
2
1:76S
n
3
sina,
_
27
where S
n
is the solidity ratio, and dened as
S
n

2D
line
l
net

1
2
D
line
l
net
_ _
2
, 28
in which D
line
is the twine diameter and l
net
is a half mesh size.
2.13. Consideration of damping effects
Experimental testing in a physical wave tank has uncovered that the dynamic responses near the resonant frequency of
body motions were damped signicantly; similar phenomena were also identied by Yamamoto et al. (1980) and may be
attributed to the uid viscous effect. The ow eld is assumed to be inviscid in the present model and determining the
damping coefcients of a oating structure is nearly impossible at this moment. To simplify this problem, an uncoupled
damping coefcient matrix is incorporated into the equation of motion of Eq. (23) to represent the damping forces.
G
Gravity center of
platform
Net element
node
node
Fig. 2. Denition sketch of nodes and elements of net with relation to the gravity center of DPFS.
H.-J. Tang et al. / Journal of Fluids and Structures 27 (2011) 918936 924
The equation is rewritten as
m
m
I
G
_

_
_

_
x
G

z
G

y
G
_ _

C
xx
C
zz
C
yy
_

_
_

_
_
x
G
_
z
G
_
y
G
_ _
K
x
G
z
G
y
G
_

_
_

_
F
x
F
z
M
y
_

_
_

_, 29
where [K] is the stiffness matrix and the components are determined numerically at each time step according to mooring
angle. F
x
, F
z
, and M
y
are the resultant hydrodynamic forces acting on the DPFS. C
xx
, C
zz
, and C
yy
are the damping coefcients
of sway, heave, and roll motion, respectively. Furthermore, these damping coefcients are assumed to be equal and
denoted by C, and are obtained from the damping ratio (Chopra, 2001),
z
C
2

Km
p : 30
In this study, the damping ratio is set to 0.1 in order to t our experimental data, K is the spring constant, and m is the
total mass of the oating structure.
3. Physical model test
The numerical model was validated by a physical model test in a wave tank (see Fig. 3). Experiments were conducted at
the Hydrodynamics Laboratory in the Department of Marine Environment and Engineering, National Sun Yat-sen
University. The wave tank, with a piston type wave generator, measured 35 m1 m1.2 m (L WH) and enabled
waves as well as uniform currents to be generated. The physical model was composed of a DPFS, a sh net cage, and a
symmetric coil spring. A load cell was installed on the sea-side mooring line. The details of the DPFS model are shown in
Fig. 4. A thick, acrylic plate was used to connect the pontoons. Beneath the acrylic plate, two iron bars were installed to
prevent bending. Iron plates were installed within the pontoons to control the draft. The DPFS was 96.8 cm long in the y
direction while the wave tank was 100 cm wide. Thus, each gap between the model edge and tank wall under
Roller
h
Load cell
Wave gage
Wave gage
14.78m 0.36m0.36m 3.67m 1m 2.5m
Coil spring
Fish net
1 2 3 4
Fig. 3. Experimental setting in a wave tank.
3.9
1
9
50 25 25
1
50
100
3.8
4
5
.
5
Unit cm
Fish net
Fish net Fish net
Iron plate Iron plate
Acrylic plate
Fig. 4. The dual pontoon structure used in the experiments.
H.-J. Tang et al. / Journal of Fluids and Structures 27 (2011) 918936 925
consideration was 1.6 cm. This is to prevent the collision between structure and tank wall. These gaps may lead to a small
amount of energy loss (3.2%(10098.6)/100), and it was regarded as negligible in this study. However, the increase of
energy loss was observed in the vicinity of structural resonance frequencies, which may be linked to the viscous effect.
Thus, the resonant peaks of measurement are generally smaller than the numerical predictions in Figs. 58. Table 1 lists
the data of the physical model of the DPFS cage system and its corresponding material. The sh net twine was 0.175 cm in
diameter, the half mesh size was 2 cm, and the specic gravity of the net was 1.14.
4. Results and discussion
In general, the RAO is used to describe the rst-order dynamic response of body motions related to incident wave
amplitudes through an FFT analysis, as shown in Table 2. The second- and third-order dynamic responses can also be
calculated by using the same analysis, since the dynamic responses of wavebody interactions are harmonic. Therefore,
the rst- to third-order RAO are adopted to describe the nonlinear dynamic responses in this study. The incident and
0 1 2 3 4 5 6
0
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
1.8
2
S
w
a
y

R
A
O
Experiments (without net)
Experiments (with net)
Present model (without net)
Present model (with net)
Fig. 5. Comparison of sway RAO between present model and experiments for the DPFS with and without a sh net.
0 1 2 3 4 5 6
0
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
1.8
2
H
e
a
v
e

R
A
O
Experiments (without net)
Experiments (with net)
Present model (without net)
Present model (with net)
Fig. 6. Comparison of heave RAO between present model and experiments for the DPFS with and without a sh net.
H.-J. Tang et al. / Journal of Fluids and Structures 27 (2011) 918936 926
reected waves are separated by Mansard and Funkes method (Mansard and Funke, 1980), using three wave gauges
installed in front of the structure.
In addition, the appropriate mesh size and time step for the model are obtained through a preliminary convergence
test. The mesh size is L/30 (L is the wave length) on the free surface, the space between each pontoon is 10 elements, h/20
on both sides of the NWT, L/10 on the bottom, and there are 40 elements on each pontoon surface. The appropriate
marching time step of the RK4 is T/64, and the total simulated time is 50T.
4.1. Model verication
In this section, the present model is used to calculate the dynamic properties of the DPFS when interacting with waves
(see Fig. 1). To verify the accuracy of the present model, the simulation values are compared with experimental results of
the DPFS having a net and not having one. The physical parameters are given in Table 1. The incident wave height is 4 cm
in all physical and numerical cases.
Fig. 5 shows that the results for the sway RAO of the present model are similar to the results from the experimental
data for s
2
h/g41.0, although differences increase for s
2
h/gr1.0. We have found that the sway response is slightly
0 1 2 3 4 5 6
0
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
1.8
2
R
o
l
l

R
A
O
Experiments (without net)
Experiments (with net)
Present model (without net)
Present model (with net)
Fig. 7. Comparison of roll RAO between present model and experiments for the DPFS with and without a sh net.
0 1 2 3 4 5 6
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
R
e
f
l
e
c
t
i
o
n

c
o
e
f
f
i
c
i
e
n
t
Experiments (without net)
Experiments (with net)
Present model (without net)
Present model (with net)
Fig. 8. Comparison of sea-side tension force RAO between present model and experiments for the DPFS with and without a sh net.
H.-J. Tang et al. / Journal of Fluids and Structures 27 (2011) 918936 927
different in the experimental results with or without a net, which may be due to the nets mass (1.645 kg, 1/38 of the DPFS
mass) and the net twine projected area (solidity ratio, S
n
0.0175) which are too small to affect the DPFS experimental
results.
Fig. 6 reveals that the present model overestimates the heave RAO when compared with the experimental results,
particularly near the resonant frequency of heave motions. This may be caused by the uid viscous effect and energy losses
between wavebody interaction in the physical wave tank (Weng and Chou, 2007). Fig. 7 shows relative similarity of the
RAO for roll between the present model and the physical model. Further, the results indicate that the resonant frequencies
of each mode of motion are different. For example, the sway modes resonant frequency lowers due to the mooring
effect. The heave modes resonant frequency is close to s
2
h/g4.0, and the roll modes resonant frequency is within
3.0os
2
h/go4.0. This can be explained by the different stiffness components which may lead to the different natural
frequencies of each mode. Physical and numerical results also reveal that when the DPFS has a net, lower resonant responses
occur and down shift to lower frequencies due to the extra mass of net. The damping effect also occurs because of the net cage.
The results of tension RAO are demonstrated in Fig. 8. Simulation results of DPFS with or without a net have similar
trends to the measurements, but simulated values are somewhat larger due to the inviscid ow property assumed in the
computation. Results of the simulation also indicate that the sh net hanging underneath the DPFS may reduce the
resonant peak in the mooring tension because of its damping effect. Finally, we can identify that the mooring tensions are
inuenced by the heave motion mode at higher frequencies s
2
h/gE4.0 and the sway motion mode at lower frequencies
s
2
h/go1.0, based on their similar trends of RAOs.
Fig. 9 shows the comparisons of reection coefcients. Similarities are found between these simulations and
measurements, except in the high frequency region (s
2
h/gZ4.5). This may be due to the severe energy loss in body
uid interaction in the physical wave tank. The results also show that the smallest reection coefcient is located near
s
2
h/g4.0, where the heave RAO reaches its maximum. This happens because wave energy is transformed into kinetic
energy of the oating structure. It also reveals that within the range of 1.5os
2
h/go3.8, the reection coefcient of the
DPFS with a net is relatively lower than its counterpart without a net, which may be due to the damping effect of the net.
4.2. Nonlinearity of dynamic responses
In our previous study (Huang and Tang, 2009), the dynamic responses of a moored DPFS had been investigated by using
a 2-D fully nonlinear numerical wave tank. The simulated results showed good agreement with experiments by Weng and
Chou (2007) except near the resonant frequency region, which may be caused by viscous effects and energy losses of the
physical model test. The nonlinear wave effects on the dynamic response were calculated for different incident wave
heights. It was found that as the incident wave height increases the resonant frequencies of heave and roll RAOs are
slightly shifted to the lower frequencies due to the increase of added mass and the damping effect. Besides, the results of
Table 1
Specications of the physical model.
Specications Sizes
Water depth (h) 0.80 m
Total mass of DPFS (m) 62.58 kg
Width of pontoon (a) 0.25 m
Spacing between pontoons (b) 0.50 m
Draft (d) 0.153 m
Roll moment arm (l
G
) 0.50 m
Moment of inertia (I
G
) 8.93 kg m
2
Spring constant (K) 674.93 N/m
Mooring angle (y
0
) 521
Gravity center (x
G
, z
G
) (0.0 m, 0.0861 m)
Net depth (d
net
) 0.455 m
Net mass (m
net
) 1.645 kg
Solidity ratio (S
n
) 0.0175
Twine diameter (D
line
) 0.175 cm
Half mesh size (l
net
) 2 cm
Specic gravity of net (Nylon) 1.14
Table 2
Denition of the RAOs.
Normalized parameters Denition
Normalized sway RAO x
G
/A
Normalized heave RAO z
G
/A
Normalized roll RAO l
G
y
G
/A
Normalized tension RAO F
T
/KA
H.-J. Tang et al. / Journal of Fluids and Structures 27 (2011) 918936 928
the rst- to third-order RAOs of body motion indicated that the 2nd-order RAO inuenced on the heave and roll motions,
especially at the resonant frequencies of the mooring system and the DPFS.
In this simulation, rst- to third-order sway, heave, pitch, and sea-side tension RAOs are shown in Figs. 10 and 11 for
the DPFS without and with a sh net. This paper is not only to investigate the nonlinearity of dynamic responses, but to
characterize the inuence of the sh net on the DPFS in its higher-order dynamic responses. Fig. 10 shows that the 2nd-
order RAO apparently increases as s
2
h=g approximates to 1.0 and 4.0 for heave, roll, and sea-side tension RAOs. The former
is related to the mooring effect; and the latter is due to the nonlinear interaction between waves and the DPFS. On the
0 1 2 3 4 5 6
0
0.5
1
1.5
2
2.5
3
S
e
a

s
i
d
e

t
e
n
s
i
o
n

R
A
O
Experiments (without net)
Experiments (with net)
Present model (without net)
Present model (with net)
Fig. 9. Comparison of reection coefcient between present model and experiments for the DPFS with and without a sh net.
0 1 2 3 4 5 6
0
0.5
1
1.5
2
S
w
a
y

R
A
O
1st
2nd
3rd
0 1 2 3 4 5 6
0
0.5
1
1.5
2
H
e
a
v
e

R
A
O
0 1 2 3 4 5 6
0
0.5
1
1.5
2
R
o
l
l

R
A
O
0 1 2 3 4 5 6
0
0.5
1
1.5
2
S
e
a

s
i
d
e

t
e
n
s
i
o
n

R
A
O
1st
2nd
3rd
1st
2nd
3rd
1st
2nd
3rd
Fig. 10. Comparison of rst- to third-order of (a) sway RAO, (b) heave RAO, (c) roll RAO, and (d) sea-side tension RAO for the DPFS without a sh net.
H.-J. Tang et al. / Journal of Fluids and Structures 27 (2011) 918936 929
contrary, the third-order RAOs for all responses in Fig. 10(a)(d) are found to be very small in this case. Similarly, the
results of the DPFS with sh net also reveal the same trends in Fig. 11, but the magnitudes of the rst-order dynamic
responses are found to be somewhat smaller; thus may indicate that the sh net has a damping role in the dynamic
responses of the DPFS.
4.3. Inuences of net depth
In order to investigate the inuences of the sh net on the structures dynamic properties (e.g., structure motions,
mooring tension responses, and reection coefcient), the input parameters are specied as follows: z
G
0.5d,
m2u
1
rad, I
G

1
2
u
2
rad
1
3
d
2

1
3
a
2
ab
2
, l
G
ab/2, a/hd/h0.25, b/h1.0, y
0
601, and K/rgh0.03, referring to
Fig. 1. Coefcients u
1
and u
2
are dependent on the density and shape of the structure, and are given the same constant value
0.9 in this study. The water depth is h1 m.
Three different conditions are simulated (i.e., without net, d
net
/h0.3 and d
net
/h0.6), selected to clarify the inuence
on dynamic responses of the DPFS. Fig. 12 shows that these three cases have almost the same results, meaning the
inuence of the net on sway RAO is negligible. Fig. 13 reveals that when the DPFS has a net, there are slightly lower
resonant responses of heave RAO than its counterpart, but the inuence of net depth is insignicant. Fig. 14 indicates that
the deeper the net, the lower the resonant response of roll RAO, which may be caused by the damping effect of the net.
However, the result also shows that the peak value of roll RAO of d
net
/h0.3 is slightly larger than the DPFS without a sh
net in this particular case; but the case of d
net
/h0.6 leads to the opposite effect. Compared to the experimental and
numerical results in Section 4.1, Fig. 7, the resonant response of roll RAO at s
2
h/gE4.0 of the DPFS with a net (where
d
net
=h is 0.57, referring to Table 1) is smaller than that of the DPFS without a net, and is consistent with the prediction of
d
net
/h0.6. These opposite predictions for d
net
/h0.3 and 0.6 may imply that the depth of sh net does not always reduce
the resonant response of roll motion and should be determined on a case-by-case basis.
Fig. 15 shows the comparison of the reection coefcient with different net depths. The three cases investigated have
peaks at the same values of s
2
h/g, close to 2.0 and 4.0, but with different magnitudes. The trend seems to be that the DPFS
with deeper net has the lower reection coefcient at the peaks due to the energy dissipated in the uidnet interaction.
The results of sea-side mooring tension RAO with different net depths is displayed in Fig. 16. The mooring force decreases
as net depth increases, especially in the vicinity of the resonant frequency. Again, this phenomenon can be induced by the
net damping effect on the body motions, which may also reduce the magnitudes.
0 1 2 3 4 5 6
0
0.5
1
1.5
2
S
w
a
y

R
A
O
0 1 2 3 4 5 6
0
0.5
1
1.5
2
H
e
a
v
e

R
A
O
0 1 2 3 4 5 6
0
0.5
1
1.5
2
R
o
l
l

R
A
O
0 1 2 3 4 5 6
0
0.5
1
1.5
2
S
e
a

s
i
d
e

t
e
n
s
i
o
n

R
A
O 1st
2nd
3rd
1st
2nd
3rd
1st
2nd
3rd
1st
2nd
3rd
Fig. 11. Comparison of rst- to third-order of (a) sway RAO, (b) heave RAO, (c) roll RAO, and (d) sea-side tension RAO for the DPFS with a sh net.
H.-J. Tang et al. / Journal of Fluids and Structures 27 (2011) 918936 930
0 1 2 3 4 5 6
0
0.5
1
1.5
2
S
w
a
y

R
A
O
Fig. 12. Comparison of sway RAO for the DPFS with different net depths.
0 1 2 3 4 5 6
0
0.5
1
1.5
2
H
e
a
v
e

R
A
O
Fig. 13. Comparison of heave RAO for the DPFS with different net depths.
0 1 2 3 4 5 6
0
0.5
1
1.5
2
R
o
l
l

R
A
O
Fig. 14. Comparison of roll RAO for the DPFS with different net depths.
H.-J. Tang et al. / Journal of Fluids and Structures 27 (2011) 918936 931
4.4. Inuence of the net width
In this section, the present model is used to investigate the inuence of the net width (or space between pontoons) on
the dynamic properties of the structure. A DPFS with net depth (d
net
/h0.6) was chosen to investigate the gap effect of the
pontoons (b/h0.5, 0.75, 1.0, 1.25, and 1.5). The rest of the input parameters of the DPFS are the same as in Section 4.3.
Comparisons of motion modes are shown in Figs. 1719. Fig. 17 shows that the primary resonant peaks of these three
cases occur at the same normalized frequency s
2
h/g0.23, which is caused by the mooring line characteristics, not the gap
between the pontoons. However, the sway RAO may create a secondary peak as the gap increases. The secondary peak
occurred when the distance between the sea-side surfaces of front and rear pontoons (i.e., ab in Fig. 1) is close to the
incident wave length. For example, in the case of b/h1.5, the secondary peak occurred near s
2
h/g3.6, where the
incident wave length is about 1.74 m and ab is 1.75 m. Under this condition, two individual wave crests simultaneously
act on the front and rear pontoons and generate the strong sway force. Thus, a secondary peak of sway motion occurs.
Fig. 18 indicates that the downshifts of resonant frequencies are more signicant than the changes of peak values when
the gap increases. Fig. 19 demonstrates that the larger the gap, the higher the peak value of the roll RAO when the gap
increases from b/h0.5 to 1.25. However, the situation changes when the gap increases from b/h1.25 to 1.5. In these
cases, the moments of inertia (I
G
) are 17.0, 29.3, 45.1, 64.5, and 87.3 (kg m
2
), and roll moment arms (l
G
) are 0.5, 0.625, 0.75,
0.875, and 1.0 (m) for b/h0.5, 0.75, 1.0, 1.25, and 1.5, respectively (these two parameters are dened in Section 4.3). Both
these parameters increase with gap widths. However, the roll motion is inversely proportional to the moment of inertia,
0 1 2 3 4 5 6
0
0.2
0.4
0.6
0.8
1
R
e
f
l
e
c
t
i
o
n

c
o
e
f
f
i
c
i
e
n
t
Fig. 15. Comparison of reection coefcient for the DPFS with different net depths.
0 1 2 3 4 5 6
0
0.5
1
1.5
2
2.5
3
3.5
4
S
e
a

s
i
d
e

t
e
n
s
i
o
n

R
A
O
Fig. 16. Comparison of sea-side tension RAO for the DPFS with different net depths.
H.-J. Tang et al. / Journal of Fluids and Structures 27 (2011) 918936 932
0 1 2 3 4 5 6
0
0.5
1
1.5
2
S
w
a
y

R
A
O
Fig. 17. Comparison of sway RAO for the DPFS with different net widths.
0 1 2 3 4 5 6
0
0.5
1
1.5
2
2.5
H
e
a
v
e

R
A
O
Fig. 18. Comparison of heave RAO for the DPFS with different net widths.
0 1 2 3 4 5 6
0
0.5
1
1.5
2
R
o
l
l

R
A
O
Fig. 19. Comparison of roll RAO for the DPFS with different net widths.
H.-J. Tang et al. / Journal of Fluids and Structures 27 (2011) 918936 933
and is proportional to the roll moment arm if the external forces are the same. Moreover, the moment of inertia (I
G
) is
found to increase faster than that of the roll moment arm (l
G
) as the gap increases, and thus the peak value of roll RAO may
decrease. A sharp decrease of roll RAO is found in the cases b/h1.0, 1.25, and 1.5. Because this phenomenon occurs near
the secondary peak of sway RAO (Fig. 17), it may imply that some amount of rolling energy is transferred to sway motion
and thus suddenly suppresses the rolling motion.
Fig. 20 shows the comparison of the reection coefcients. The results illustrate that a relative minimum value occurs
in the vicinity of s
2
h/g3.2, which corresponds to the frequency of the maximum heave and roll motion responses of the
DPFS (Figs. 1719); most of the wave energy is absorbed by the oating structure at this frequency which reduces the
reection coefcient. Another interesting phenomenon is associated with the cases b/h0.75, 1.0, and 1.25: these have
the smallest reection coefcients at s
2
h/g5.8, 4.7, and 4.0, corresponding to the second peaks of the sway motions
(see Fig. 17).
Fig. 21 shows the comparison of sea-side tension RAO with different net widths. The peak values of sea-side tension
RAO increase as gap increases, while the peaks of resonant frequency downshift. This matches the observation that the
drag force increases as the net size increases, and the total mass increase will reduce the resonant frequency.
5. Conclusions
In this study, a fully nonlinear NWT is developed to investigate the dynamic response of the moored DPFS cage in
waves. The mooring system is considered to be linearly elongated and symmetrically installed. This model is solved by the
0 1 2 3 4 5 6
0
0.2
0.4
0.6
0.8
1
R
e
f
l
e
c
t
i
o
n

c
o
e
f
f
i
c
i
e
n
t
Fig. 20. Comparison of reection coefcient for the DPFS with different net widths.
0 1 2 3 4 5 6
0
0.5
1
1.5
2
2.5
3
S
e
a

s
i
d
e

t
e
n
s
i
o
n

R
A
O
Fig. 21. Comparison of sea-side tension RAO for the DPFS with different net widths.
H.-J. Tang et al. / Journal of Fluids and Structures 27 (2011) 918936 934
BEM, and the free surface nodes are tracked by the MEL approach with a cubic spline scheme and the RK4 method.
Damping zones are arranged at both ends of the tank to absorb reected wave energy and to dissipate the transmitted
wave energy. The instantaneous oating body motion is calculated by an acceleration potential method and a modal
decomposition method.
This numerical model is validated with experimental data. Results indicate that the simulations of the present model
are similar to the experimental data, except near the resonant frequencies of heave motion, which is due to the viscosity
effect and energy loss from bodyuid interaction. Both experimental and numerical data show that resonant responses
observed in the motions and mooring tension of the DPFS with a net are slightly smaller than that of the DPFS without a
net. The second-order RAOs for body motions and tension force are also found at the same resonant frequencies, and these
responses of the DPFS with a net are generally smaller than that of the case without a net. The inuences of the net depth
and the net width on the dynamic response of the DPFS cage system are discussed. The resonant responses of roll and
tension RAO generally decrease as net depth increases, but the magnitudes of these changes are very small. The inuence
of net width on the dynamic motions is not only larger, but also more complicated than the inuence of net depth.
In engineering practices, special attention should be given to the gap in a dual pontoon oating structure.
Acknowledgments
The authors acknowledge the support in part by the National Science Council (NSC99-2221-E-110-092), and sponsored
by the Fisheries Agency (98A10083), Council of Agriculture of Taiwan. Also, discussions with Professor John R.C. Hsu for
several concepts are deeply appreciated.
Appendix
After solving the BIE for velocity potential eld, all the values f
j
and qf
j
/qn on the boundaries are obtained. The internal
horizontal and vertical velocities at position (x
i
,z
i
) are then given as
ux
i
,z
i

N
j 1
_
G
j
@h
ij
@x
f
j

@g
ij
@x
@f
j
@n
_ _
dG
j
,
wx
i
,z
i

N
j 1
_
G
j
@h
ij
@z
f
j

@g
ij
@z
@f
j
@n
_ _
dG
j
,
_

_
31
where N is the total number of boundary elements, and h
ij
and g
ij
are dened as
h
ij

@
@n
1
2p
lnr
ij
_ _

1
2p
xx
i
n
x
zz
i
n
z
r
ij
2
_ _
,
g
ij

1
2p
lnr
ij
:
_

_
32
Substitute Eq. (32) into Eq. (31) the internal velocity at any location in the ow eld can be achieved.
From the total derivative equation, the internal horizontal and vertical accelerations are dened as
du
dt

@u
@t
u
@u
@x
w
@u
@z
,
dw
dt

@w
@t
u
@w
@x
w
@w
@z
,
_

_
33
in which qu/qt and qw/qt are acquired from taking partial derivative of Eq. (31) with respect to t,
@u
@t
x
i
,z
i

N
j 1
_
G
j
@h
ij
@x
f
tj

@g
ij
@x
@f
tj
@n
_ _
dG
j
,
@w
@t
x
i
,z
i

N
j 1
_
G
j
@h
ij
@z
f
tj

@g
ij
@z
@f
tj
@n
_ _
dG
j
,
_

_
34
where f
tj
and qf
tj
/qn are obtained by solving the BIE of acceleration eld.
Similarly, by taking the derivatives of Eq. (31) with respective to x and z, then qu/qx, qu/qz, qw/qx, and qw/qz in Eq. (33)
are acquired.
References
Bai, W., Eatock Taylor, R., 2006. Higher-order boundary element simulation of fully nonlinear wave radiation by oscillating vertical cylinders. Applied
Ocean Research 28, 247265.
H.-J. Tang et al. / Journal of Fluids and Structures 27 (2011) 918936 935
Bai, W., Eatock Taylor, R., 2007. Numerical simulation of fully nonlinear regular and focused wave diffraction around a vertical cylinder using domain
decomposition. Applied Ocean Research 29 (12), 5571.
Bai, W., Eatock Taylor, R., 2009. Fully nonlinear simulation of wave interaction with xed and oating ared structures. Ocean Engineering 36, 223236.
Boo, S.Y., 2002. Linear and nonlinear irregular waves and forces in a numerical wave tank. Ocean Engineering 29, 475493.
Brebbia, C.A., Walker, S., 1979. Dynamic Analysis of Offshore Structures. Newnes-Butterworths, London, pp. 109143.
Cao, Y., Beck, R.F., Schultz, W.W., 1994. Nonlinear computation of wave loads and motions of oating bodies in incident waves. In: Proceedings of Ninth
Workshop on Water Waves and Floating Bodies, pp. 3337.
Chopra, A.K., 2001. Dynamics of Structures: Theory and Application to Earthquake Engineering. Prentice-Hall Inc., Upper Saddle River.
Contento, G., 2000. Numerical wave tank computations of nonlinear motions of two-dimensional arbitrarily shaped free oating bodies. Ocean
Engineering 27, 531556.
Cointe, R., 1990. Numerical simulation of a wave channel. Engineering Analysis with Boundary elements 7 (4), 167177.
Fredriksson, D.W., Swift, M.R., Irish, J.D., Tsukrov, I., Celikkol, B., 2003. Fish cage and mooring system dynamics using physical and numerical models with
eld measurements. Aquacultural Engineering 27, 117146.
Grilli, S.T., Svendsen, I.A., 1990. Corner problems and global accuracy in the boundary element solution of nonlinear wave ows. Engineering Analysis
with Boundary Elements 7 (4), 178195.
Huang, C.C., Tang, H.J., 2009. Dynamic response of moored oating dual pontoon structure in a fully nonlinear numerical wave tank. In: Proceedings of the
19th International Offshore and Polar Engineering Conference, vol. 3, pp. 414421.
Huang, C.C., Tang, H.J., Chen, W.M., 2008. On the interaction between random waves and a freely oating body in a fully nonlinear numerical wave tank.
In: Proceedings of the 18th International Offshore and Polar Engineering Conference vol. 3, pp. 148155.
Huang, C.C., Tang, H.J., Liu, J.Y., 2006. Dynamical analysis of net cage structures for marine aquaculture: numerical simulation and model testing.
Aquacultural Engineering 35, 258270.
Huang, C.C., Tang, H.J., Liu, J.Y., 2007. Modeling volume deformation in gravity-type cages with distributed bottom weights or a rigid tube-sinker.
Aquacultural Engineering 37, 144157.
Isaacson, M.De St.Q., 1982. Nonlinear-wave effects on xed and oating bodies. Journal of Fluid Mechanics 120, 267281.
Kim, C.H., Cle ment, A.H., Tanizawa, K., 1999. Recent research and development of numerical wave tanksa review. International Journal of Offshore and
Polar Engineering 9 (4), 241256.
Koo, W.C., 2009. Nonlinear time-domain analysis of motion-restrained pneumatic oating breakwater. Ocean Engineering 36, 723731.
Koo, W.C., Kim, M.H., 2004. Freely oating-body simulation by a 2-D fully nonlinear numerical wave tank. Ocean Engineering 31, 20112046.
Koo, W.C., Kim, M.H., 2007. Fully nonlinear wavebody interactions with surface-piercing bodies. Ocean Engineering 34, 10001012.
Lee, H.H., Wang, W.S., 2005. The dragged surge motion of tension-leg type sh cage system subjected to multi-interactions among wave and structures.
IEEE Journal of Oceanic Engineering 30 (1), 5978.
Loland, G., 1991. Current Forces On and Flow through Fish Farms, Division of Marine Hydrodynamics. Norwegian Institute of Technology, Trondheim,
Norway, pp. 8595.
Longuet-Higgins, M.S., Cokelet, E., 1976. The deformation of steep surface waves on water: I. A numerical method of computation. Proceedings of Royal
Society of London A350, 126.
Ma, Q.W., Yan, S., 2009. QALE-FEM for numerical modeling of non-linear interaction between 3D moored oating bodies and steep waves. International
Journal for Numerical Methods in Engineering 78, 713756.
Mansard, E.P.D., Funke, E.R., 1980. The measurement of incident and reected spectra using a least squares method. In: Proceedings of the 17th Coastal
Engineering Conference, ASCE, pp. 154172.
Moe, H., Fredheim, A., Hopperstad, O.S., 2010. Structural analysis of aquaculture net cage in current. Journal of Fluids and Structures 26, 503516.
Sun, L., Eatock Taylor, R., Taylor, P.H., 2010. First- and second-order analysis of resonant waves between adjacent barges. Journal of Fluids and Structures
26, 954978.
Tanizawa, K., 1995. A nonlinear simulation method of 3-D body motions in waves (1st Report). Journal of the Society of Naval Architect Japan 178,
179191.
Tanizawa, K., 1996. Long time fully nonlinear simulation of oating body motions with articial damping zone. Journal of the Society of Naval Architects
of Japan 180, 311319.
Tanizawa, K., 2000. The state of the art on numerical wave tank. In: Proceedings of the Fourth Osaka Colloquium on Seakeeping Performance of Ships,
pp. 95114.
Vinje, T., Brevig, P., 1981. Nonlinear ship motions. In: Proceedings of the Third International Conference on Numerical Ship Hydrodynamics, pp. IV3-1IV3-10.
Weng, W.K., Chou, C.R., 2007. Analysis of response of oating dual pontoon structure. China Ocean Engineering 21 (1), 91104.
Williams, A.N., Abul-Azm, A.G., 1997. Dual pontoon oating breakwater. Ocean Engineering 24 (5), 465478.
Wu, G.X., Eatock Taylor, R., 1996. Transient motion of a oating body in steep water waves. In: Proceedings of Eleventh International Workshop on Water
Waves and Floating Bodies, Hamburg.
Yan, S., Ma, Q.W., 2007. Numerical simulation of fully nonlinear interaction between steep waves and 2D oating bodies using the QALE-FEM method.
Journal of Computational Physics 221, 666692.
Yamamoto, T., Yoshida, A., Ijima, T., 1980. Dynamics of elastically moored oating objects. Applied Ocean Research 2 (2), 8592.
H.-J. Tang et al. / Journal of Fluids and Structures 27 (2011) 918936 936

Вам также может понравиться