Vous êtes sur la page 1sur 14

720

Hydrodynamic analysis of wave interactions


with a moored oating breakwater using the
element-free Galerkin method
Jeongwoo Lee and Woncheol Cho
Abstract: This paper deals with a numerical investigation of incident wave interactions with a moored pontoon-type
oating breakwater. The element-free Galerkin method, in which only nodal data are required to analyze the problem, is
employed to solve the diffraction and radiation boundary value problems addressed by the modied Helmholtz equation.
The numerical model includes the hydrodynamic and mooring analyses, and it is validated by previous numerical and
experimental results. Using the numerical model, we are able to assess the hydrodynamic performance of a moored
pontoon-type oating breakwater in regular waves. Numerical results are presented to show the effects of wave conditions
and mooring system conguration. This paper also presents the simple forms of stiffness coefcients of a slack mooring
line. The inuence of mooring line condition on the performance of a oating breakwater is highlighted.
Key words: moored oating breakwater, element-free Galerkin method, mooring line condition.
Rsum : Cet article prsente une tude numrique des interactions des ondes incidentes sur un brise-lames ottant amarr
de type ponton. La mthode sans maillage de Galerkin, dans laquelle uniquement les donnes nodales sont requises pour
analyser le problme, est employe pour rsoudre les problmes de valeur de diffraction et de rayonnement aux limites
tels quabords par lquation dHelmholtz modie. Le modle numrique comprend les analyses hydrodynamiques et
damarrage et il est valid par les rsultats numriques et exprimentaux antrieurs. En utilisant le modle numrique,
nous pouvons valuer la performance hydrodynamique dun brise-lames ottant amarr de type ponton avec des conditions
de vagues rgulires. Les rsultats numriques sont prsents an de montrer les effets des conditions de vagues et de
la conguration damarrage. Cet article prsente galement les formes simples des coefcients de rigidit dune ligne
damarrage tension variable. Linuence de ltat de la ligne damarrage sur le rendement dun brise-lames ottant est
souligne.
Mots cls : brise-lames ottant amarr, mthode sans maillage de Galerkin, condition de la ligne damarre.
[Traduit par la Rdaction]
1. Introduction
Floating breakwaters are considered to be alternatives to con-
ventional xed breakwaters. They are useful in preserving small
marinas and recreational harbors from wave forces. The cost of
a oating system is slightly dependent on water depth and bot-
tomfoundation conditions. In addition, oating breakwaters are
ecologically advantageous because they have little interaction
withwater circulation, biological exchange, andsediment trans-
port. Thus, many studies have been undertaken to investigate
the performance of various types of oating breakwaters.
Numerical models based on the linear potential theory have
proven to be efcient tools to predict the sea-keeping behavior
of oating breakwaters. Generally, there are two main classes of
Received 10 October 2002. Revision accepted 7 March 2003. Pub-
lished on the NRC Research Press Web site at http://cjce.nrc.ca/ on
8 August 2003.
J. Lee
1
and W. Cho. Department of Civil Engineering, Yonsei
University, Seoul 120-749, Korea.
Written discussion of this article is welcomed and will be received
by the Editor until 31 December 2003.
1
Corresponding author (e-mail: ljw007@mail.yonsei.ac.kr).
numerical technique to solve the wave diffraction and radiation
boundary value problems for velocity potentials when a oat-
ing body undergoes oscillations in waves. The most popular
methods are the nite element method (FEM) and the bound-
ary element method (BEM). Both methods have been widely
used to compute the exciting forces and hydrodynamic coef-
cients due to wave interactions with a oating structure. The
FEM has been successfully tested by Chen and Mei (1974),
Bai (1975), Newton (1978), and Sharan (1986) to evaluate the
hydrodynamic coefcients. Bai and Yeung (1974) applied the
FEM technique to a wavestructure problem and compared the
results with those from other methods. Leonard et al. (1983)
used the FEM to solve the boundary value problems relating
to diffraction and radiation phenomena near the oating body,
performing hydrodynamic analyses to estimate the added mass
coefcients, damping coefcients, and responses of a oat-
ing body with varying incident wave angle, draft, depth, etc.
Leonard et al. paid particular attention to negative added mass
in heave and the corresponding sharp damping coefcients for
a dual structure at certain wave frequency ranges. Hanif (1983)
and Sannasiraj et al. (1995, 1998) presented applications of
the FEM to analyze the hydrodynamic behavior of a oating
breakwater subject to its motions. Recently, the hydrodynamic
coefcients and forces on a twin structure under oblique waves
Can. J. Civ. Eng. 30: 720733 (2003) doi: 10.1139/L03-020 2003 NRC Canada
Lee and Cho 721
have been evaluated using the numerical method (Sannasiraj et
al. 2000).
On the other hand, a numerical method based on the BEM
or wave source method was used to obtain the solution to the
wavestructure interaction problem (Garrison 1978; Sarpkaya
and Isaacson 1981; Chakrabarti 1987; Isaacson and Baldwin
1996). Au and Brebbia (1982) investigated the application of
the BEMto compute wave forces on xed and oating offshore
structures of arbitrary shapes. Isaacson and Nwogu (1987) de-
veloped a numerical procedure based on Greens theorem to
compute the exciting forces and hydrodynamic coefcients due
to the interaction of a regular oblique wave train with a long
oating cylinder. They also dealt with the effects of wave direc-
tionality on the loads and motions of long structures. Bhat and
Isaacson (1998) made BEM-based numerical and experimen-
tal assessments of the performance of moored twin-pontoon
oating breakwaters with rectangular- and circular-section pon-
toons. They evaluated the inuence of various design parame-
ters, such as relative draft, pontoon spacing, and mooring line
slackness, on the breakwater performance.
Even though the FEM and BEM, as mentioned previously,
have been widely applied to solve the free-surface ow prob-
lems, both solutions have certain limitations. The BEM has
difculty in handling nonlinear and discontinuous problems,
and it also needs high computational times because the matrix
arising in this formulation is full and nonsymmetric. The FEM
requires labour-intensive, time-consuming human resources for
mesh generation. In particular, numerical experiments in which
the geometry of the structure is varied need to be performed be-
cause the performance and response of oating breakwaters
are signicantly inuenced by the width, draft, and shape of
structure. Therefore, preprocessing tasks for mesh generation
may be reduced if meshless methods such as the element-free
Galerkin method (EFGM) are used, in which nodes represent-
ing the geometry of the structure and the boundary domain can
be easily added or removed. In the present paper, we focus on
the EFGM for the hydrodynamic analysis of wave interactions
with a oating breakwater.
The EFGMwas introduced by Nayroles et al. (1992) and was
rst called the diffuse element method (DEM). It is now better
known as the EFGM, after modication by Belytschko et al.
(1994). The main advantages of this method are no mesh, no
element connectivity, and easy pre- and post-processing tasks.
The EFGM has been increasingly applied in solid mechanics
(Belytschko et al. 1994, 1995; Belytschko and Tabbara 1996).
Recently, the method has been introduced in the elds of mul-
tiphase deforming porous media (Modaressi and Aubert 1998),
hydraulics (Du 1999, 2000), and acoustic problems (Suleau and
Bouillard 2000). No previous efforts have been made, however,
to apply the EFGM to water wave problems. Therefore, the
main objective of the present paper is to propose an EFGM-
based numerical model for simulating the wave interactions
with a oating breakwater. The formulation and numerical im-
plementation are presented. By using the numerical model, the
hydrodynamic performance of a moored oating breakwater in
regular waves is assessed. In the design of a moored oating
breakwater, it is necessary to assess the mooring forces act-
ing on the body. These forces should be considered in evaluat-
ing restoring forces depending on the mooring line stiffnesses.
Therefore, this paper also presents the simple forms of stiffness
coefcients of a slack mooring line. The inuence of mooring
line condition on the performance of a oating breakwater is
highlighted.
2. Theoretical formulation
2.1. Hydrodynamic analysis
The present paper theoretically deals with the performance
of pontoon-type oating breakwaters. Theoretical formulation
is based on the two-dimensional linear diffraction and radiation
problems due to the presence of a oating breakwater and its
motions in waves. Based on the assumptions of an inviscid and
incompressible uid and an irrotational ow, the ideal uid mo-
tion can be represented by a velocity potential that satises
the Laplace equation within the uid domain. The breakwater is
regarded as a long, rigid, horizontal cylinder oscillating in sway,
heave, and roll, with each motion periodic in time and along the
longitudinal axis. Subject to the basic assumption that the inci-
dent waves and oscillatory body motions are sufciently small,
the problem of wave interaction with a oating breakwater can
be described by a linear superposition of the diffraction and
radiation problems (Isaacson and Nwogu 1987).
The right-handed Cartesian coordinate system is employed,
with z measured upward from still water level and the y axis
parallel to the longitudinal axis of the oating body. A deni-
tion sketch of a wave oating breakwater interaction problem
is shown in Fig. 1. The system is subject to a regular, small-
amplitude wave train of height H and angular frequency .
The incident waves obliquely propagate in water of depth d in
a direction making an angle with the longitudinal axis.
The total velocity potential is assumed to be the sum of the
incident potential
0
, the scattered potential
4
, and the radi-
ation potential
j
(j = 1, 2, 3) in three modes of sway, heave,
and roll. This is mathematically expressed in the form
[1] =
0
+
4
+
3

j=1

j
The incident wave potential can be represented by the linear
wave theory given by

0
(x, y, z, t ) = Re
_

igH
2
cosh [k(z +d)]
cosh(kd)
[2]
exp [i(kx cos +ky sin t )]
_
= Re
_

igH
2

0
(x, z)
exp [i(ky sin t )]
_
where
0
(x, z) is the nondimensional incident velocity poten-
tial; i is the imaginary unit,

1; g is the gravitational acceler-


ation; and k is the wave number, which is related to the angular
frequency by the dispersion relation
[3]
2
= gk tanh(kd)
2003 NRC Canada
722 Can. J. Civ. Eng. Vol. 30, 2003
Fig. 1. Denition sketch.
For a linear diffraction problem, the scattered velocity po-
tential, which is considered to vary sinusoidally in time and the
y axis, is described by s
[4]
4
(x, y, z, t )
= Re
_

igH
2

4
(x, z) exp [i(ky sin t )]
_
where
4
(x, z) is the unknown nondimensional scattered ve-
locity potential. The boundary value problem for the scattered
potential satisfying the Laplace equation (
2

4
= 0) can be
stated as the two-dimensional modied Helmholtz equation
(Isaacson and Nwogu 1987):
[5a]

2

4
x
2
+

2

4
z
2

2

4
= 0 in the uid domain
[5b]

4
z
=

2
g

4
at z = 0
[5c]

4
z
= 0 at z = d
[5d]

4
n
=

0
n
on S
B
[5e]

4
x
= ik cos
4
, x =
where is the x component of wave number k, k sin ; n is the
outward unit normal vector from the body boundary; and S
B
is
the body boundary surface.
The wave radiation problemcan be described by the radiation
velocity potential associated with displacement or rotation in
the jth mode. The radiation potential varying periodically in
time and the y axis can be expressed as
[6]
j
(x, y, z, t ) = Re
_
i
j

j
(x, z)
exp [i (y t )]
_
, j = 1, 2, 3
where
j
(x, z) is the unknown nondimensional radiated veloc-
ity potential;
j
is the complex amplitude of the oscillation of
the body with three modes; and j = 1, 2, 3 denotes x, z direc-
tions and rotation about the y axis, respectively. The radiation
boundary value problem is dened similar to the diffraction
problem. The governing equation and boundary conditions for
the radiation velocity potential have the same forms as those
of the boundary value problem for the scattered potential as
stated in eqs. [5a][5e], except for the body surface boundary
condition. The boundary condition on the body surface is given
by
[7]

j
n
= n
j
, j = 1, 2, 3 on S
B
where n
1
, n
2
are the x, z components of the outward unit normal
vector n from the body boundary and n
3
= (z e)n
1
xn
2
, in
which (0, e) is the coordinate of the center of rotation.
Once the complex scattered potential,
4
, and the complex
radiation potential,
j
, are obtained, the hydrodynamic pres-
sures and forces can be computed as follows. From the lin-
earized Bernoulli equation, the hydrodynamic pressure can be
expressed as
[8] p =

t
= i
where is the uid density. The hydrodynamic forces and mo-
ment per unit length, F
j
, can be computed by simply integrating
the pressures along the body boundary surface, S
B
:
[9] F
j
=
_
S
B
pn
j
dS, j = 1, 2, 3
It is usually convenient to separate the hydrodynamic forces into
the wave exciting forces F
e
j
due to the diffraction potential and
the motion-induced forces F
m
j
due to the radiation potential.
The wave exciting forces and the motion-induced forces can
be expressed in eqs. [10] and [11], respectively (Isaacson and
Nwogu 1987):
F
e
j
= Re
_
1
2
gH
_
S
B
(
0
+
4
)n
j
[10]
exp [i(y t )] dS
_
= Re
_
f
j
exp [i(y t )]
_
, j = 1, 2, 3
where f
j
is the complex wave exciting force amplitudes and can
be obtained from the incident and scattered velocity potentials,

0
and
4
, respectively, and
F
m
j
= Re
_

i
_
S
B

i
n
j
exp [i(y t )]dS
_
[11]
= Re
_

ji

i
+i
ji

i
exp [i(y t )]
_
,
i, j = 1, 2, 3
where
ji
is the added mass coefcient and
ji
is the damping
coefcient. The former is proportional to the body accelera-
tion and the latter to the body velocity. These hydrodynamic
coefcients are obtained from the radiated potential
j
.
In the frequency domain, the equations of motion of the oat-
ing body oscillating in three modes can be expressed as
[12]
3

i=1
_

2
(M
ji
+
ji
) i
ji
+C
ji
+K
ji
_

i
= f
j
, j = 1, 2, 3
2003 NRC Canada
Lee and Cho 723
where M
ji
, C
ji
, and K
ji
are the mass, hydrostatic stiffness,
and mooring stiffness matrix coefcients, respectively. The dy-
namic responses of the oating breakwater with three modes
varying sinusoidally in time and the y axis can be calculated by
solving eq. [12].
2.2. Mooring analysis
The mooring analysis is carried out in conjunction with the
hydrodynamic analysis to obtain the mooring line loads, the
conguration of each line, and the associated restraints. Jain
(1980) described the restraints of a mooring cable as a stiffness
matrix that has linearized mooring line stiffness coefcients
in sway and heave modes. These stiffness coefcients are de-
rived analytically by using the basic catenary equation for the
cable equilibrium. It is assumed that the cable is perfectly exi-
ble, inextensible, andheavy. The linearizedstiffness coefcients
are evaluated from the weight per unit length of the mooring
line in water, the initial horizontal tension, the initial tension
at the body, the initial tension at the anchor, the coordinates
of the body with respect to the anchor point, and the mooring
line conguration, such as the length of the mooring line, the
angle at the anchor point, the angle made with the body, and
the point of attachment of the mooring line with the oating
structure. The mathematical forms of stiffness coefcients are
omitted in this paper. The detailed expressions are presented
in Jain (1980). Sannasiraj et al. (1998) extended the method
of Jain by considering the restoring moment of mooring lines
against the rolling motion. Sannasiraj et al. conducted a com-
prehensive theoretical and experimental study of the behaviour
of moored single-pontoon-type oating breakwaters and pre-
sented the wave-attenuation characteristics for various moor-
ing system congurations; however, they only considered taut
mooring systems. In taut cables, there is no variable tangential
contact with the seabed, so mooring line tension may be in-
creased signicantly by a little movement of the oating body.
This steep increase of mooring tension may affect the safety of
the mooring system. Practically, a oating breakwater would be
moored by a number of slack cables in which there is variable
tangential contact with the seabed. In our research, we carried
out a mooring analysis by considering the slack mooring sys-
tem. Thus, in the present paper, mooring stiffness coefcients
are derived for the slack condition. The derivation procedure is
presented in Appendix A. The effects of mooring condition on
the wave attenuation and the mooring forces are discussed in
Sect. 4.4.
3. Element-free Galerkin method
formulation
3.1. Moving least-squares interpolant
In the EFGM, a moving least-squares (MLS) principle is used
to formulate the shape functions. No connectivity is needed,
which is one of the main differences between the EFGM and
the FEM. The interpolant u
h
(x) of a function u(x) in a domain
can be dened as
[13] u
h
(x) =
m

i=1
p
i
(x)
i
(x) = p(x)
T
(x)
where p(x) is the polynomial basis in the space x
T
= [x, z],
superscript T represents the transpose of the matrix, and (x) is
a vector of the coefcients. A linear basis and a quadratic basis
in two dimensions are provided by
[14] p = [1, x, z]
T
for m = 3
[15] p = [1, x, z, x
2
, xz, z
2
]
T
for m = 6
The coefcients (x) are determined at any point x by mini-
mizing a weighted discrete least-squares norm J:
[16] J =
n

j=1
w
j
(x)[p
T
(x
j
)(x) u
h
j
]
2
where w
j
(x) denotes the weight function, w
j
(x) w(x x
j
),
which is non-negative and decays with increasing distance to a
point j; n is the number of points in the neighborhood of x; and
u
h
j
is the nodal value of u at x = x
j
. Standard minimization
with respect to (x) gives
[17] =

C
1
u
h
with

C
1
= A
1
B
where
[18] A =
n

j=1
w
j
(x)p(x
j
)
T
p(x
j
)
[19] B = [w
1
(x)p(x
1
), w
2
(x)p(x
2
), , w
n
(x)p(x
n
)]
Hence, we have
u
h
(x)

= p
T

C
1
u
h
[20]
=
n

j=1
m

i=1
p
i
(x)
_
A
1
(x)B(x)
_
ij
u
h
j
=
n

j=1
N
j
(x)u
h
j
From eq. [20], EFGM shape functions N
j
(x) can then be ex-
pressed as
[21] N
j
(x) =
m

i=1
p
i
(x)
_
A(x)
1
B(x)
_
ij
The partial derivatives of the interpolation function N
j
(x) are
thus obtained as
[22] N
j,x
(x) =
m

i=1
_
p
i,x
(A
1
B)
ij
+p
i
(A
1
,x
B +A
1
B
,x
)
ij
_
where
[23] A
1
,x
= A
1
A
,x
A
1
[24] A
,x
=
n

j=1
w
,x
(x x
j
)p(x
j
)p(x
j
)
T
2003 NRC Canada
724 Can. J. Civ. Eng. Vol. 30, 2003
[25] B
,x
= [w
,x
(x x
1
)p(x
1
), w
,x
(x x
2
)p(x
2
),
, w
,x
(x x
n
)p(x
n
)]
In this study, the exponential weight function will be used with
a circle domain of inuence. The exponential weight function
at point j is given by
[26] w(x x
j
) = w(d
s
) =
e
(d
s
/c)
q
e
(d
m
/c)
q
1 e
(d
m
/c)
q
,
d d
m
where d
s
is the distance between point x and node x
j
, c is a con-
stant that controls the relative weights, q is a power parameter,
and d
m
is the size of the domain of inuence of x
j
. For d
m
we
have chosen the expression dened by Ouatouati and Johnson
(1999):
[27] d
m
= d
max
c
j
In eq. [27], d
max
is a scaling parameter that is generally 1.04.0,
and c
j
is the maximumdistance that is determined by searching
for enough neighbor nodes for A to be regular and invertible.
We used a uniformdistribution of nodes, so c
j
can be described
as
[28] c
j
=
_
dx
2
+dz
2
where dx and dz are the internodal distances. In our applica-
tions, a parameter c is set to c
j
. For a detailed presentation of
the EFGM, the reader is referred to Nayroles et al. (1992) and
Belytschko et al. (1994).
3.2. Application to the diffractionradiation problem
The EFGM is applied to solve the diffractionradiation
boundary value problem of wave interactions with a oating
breakwater. To obtain the discrete equations, it is necessary to
use a variational weak form of the governing equation, which
is addressed by the modied Helmholtz equation and boundary
conditions. Taking the weak form of eq. [5a], integrating by
parts, and substituting the natural and mixed boundary condi-
tions yield
[29]
_

(
i

i
+
2
)d
_

f
()d
ik cos
_

()d =
_

B
(

n
)d
where
f
is the free surface;

is the radiation boundary;


B
is
the immersedbodysurface; and is a test functioncorrespond-
ing to , which is nondimensional scattered or radiated velocity
potential ( =
j
, j = 1, 2, 3, 4). Then, a standard Galerkin
procedure is used to obtain the discrete equations correspond-
ing to the weak form of the governing equation. According to
eq. [20], the approximate velocity potential is described by
[30]
h
=
n

i=1
N
i

i
= N
where is a set of n nodal values for the velocity potential
and N
i
are the shape functions dened in eq. [21]. Substitut-
ing the expression of
h
given in eq. [30] into a weak form
of the governing equation (eq. [29]) yields a linear system of
equations:
[31] [ K
1
+K
2
+K
3
] {} = {P}
where
[32] [K
1
] =
_

(N) (N)
T
+
2
NN
T
d
[33] [ K
2
] =

2
g
_

f
NN
T
d
[34] [ K
3
] = ik cos
_

NN
T
d
The load vector for the diffraction problem is given by
[35] { P } =
_

B
N

0
n
d
For the radiation problem, the load vector in each degree of
freedom is
[36] { P } =
_

B
Nn
j
d, j = 1, 2, 3
For many problems, the imposition of an essential boundary
condition is difcult because, in general, MLS shape functions
lack the Kronecker delta property (eq. [37]). To overcome this
problem, the Lagrange multipliers method, the FEM, and the
penalty method have been used. Such a difculty is not en-
countered in the diffractionradiation problem considered in
this study, however, because no essential boundary conditions
exist in eqs. [5b][5e]:
[37]
i
(x
j
) =
ij
=
_
1, i = j
0, i = j
To numerically calculate the integrals of eqs. [32][36], the
background integral cells are used over the entire domain. In
each integration cell, 4 4 Gauss quadrature is used to eval-
uate the stiffness of the EFGM. At each quadrature point, the
domain of inuence and the nodes inside it are determined. The
discrete equations are assembled by looping over each quadra-
ture point. The EFGM shape functions and its derivatives are
calculated by eqs. [21] and [22] and assembled by looping over
each integration cell. The system equations are solved to get
velocity potentials using a polynomial basis function with an
exponential weight function.
4. Numerical tests and applications
4.1. Validation of the numerical model
The present EFGM-based numerical model has been vali-
dated by comparison with the numerical results previously re-
ported by Sannasiraj et al. (1998). Consider now the case of
a moored, single-pontoon, rectangular-section breakwater with
depth d = 2.35 m, width a = 0.4 m, and draft D = 0.1 m. The
regular wave height is 0.05 m and the frequency range is 0.3
1.5 Hz. The body is restrained by four mooring lines, each of
length L = 4.7 m. The computation is performed in a bounded
2003 NRC Canada
Lee and Cho 725
Fig. 2. Comparison of responses of amplitude operators (RAOs)
in sway as a function of normalized wave frequency. EFGM,
element-free Galerkin method; FEM, nite element method.
domain. The domain considered here is described by 3616 inner
points and 258 boundary surface points that are regularly dis-
tributed. The distance between nodes is 0.05 m. A linear basis
and circular domains of inuence with d
max
= 1.0 to 1.5 are
used.
Figures 24 compare the computed responses of amplitude
operators (RAOs) in sway, heave, and roll motions as a function
of normalized wave frequency,
2
a/2g, with the experimental
and numerical results of Sannasiraj et al. (1998), who used the
FEMto solve the corresponding boundary value problem. Here,
RAO means the ratio of motion amplitude to wave amplitude.
The computed RAOvalues agree well with the results estimated
by Sannasiraj et al. over the whole range of
2
a/2g. In partic-
ular, both numerical solutions are almost the same except at a
frequency range of
2
a/2g from0.2 to 0.3 in sway motion. The
discrepancy may be attributed to the difference of the numer-
ical scheme and also to the viscous damping, which was con-
sidered, near the resonant frequency, in the theoretical model
of Sannasiraj et al. In their simulation, an empirical damping
coefcient estimated from the experimental results was added
to damping coefcients in the equations of motion.
As shown in Fig. 5, we see that the computed transmis-
sion coefcients (K
t
) are in satisfactory agreement with the
results of Sannasiraj et al. (1998). There are some variations
between the experimental and numerical wave transmissions,
however. Sannasiraj et al. explain that these variations would
be caused by viscous damping forces and various uncertainties.
The most signicant disagreement between the numerical and
experimentally measured transmission coefcients appears to
be near
2
a/2g = 1.0 to 1.1, where the measured minimum
wave transmission is not as extreme as predicted by the numer-
ical model. In our opinion, this is because the viscous damp-
ing due to the ow separation at sharp corners damps out the
heave motion. This means that the signicantly reduced heave
motion would cause a proportional increase in the measured
wave transmission. Moreover, the reduced heave motion may
be attributed, in part, to the very taut mooring condition be-
cause mooring lines are not enough to allow for large motions.
In particular, at
2
a/2g = 1.1, the sway and heave motions
Fig. 3. Comparison of RAOs in heave as a function of
normalized wave frequency.
Fig. 4. Comparison of RAOs in roll as a function of normalized
wave frequency.
Fig. 5. Comparison of transmission coefcients (K
t
) as a function
of normalized wave frequency.
2003 NRC Canada
726 Can. J. Civ. Eng. Vol. 30, 2003
are restricted by each other because the phase angles of each
motion with respect to the incident wave are almost the same.
Discrepancies are also seen in the results for the transmission
coefcients in the range of
2
a/2g from 1.5 to 2.0, where the
wave generated by sway motion contributes a large fraction of
the entire transmitted wave. The cause of these discrepancies
for short waves is that the measured values of sway motion are
less than the predictions of the numerical model. It is difcult
to determine the reason for differences in experimentally mea-
sured and numerically computed sway motions, however. In
previous work by Isaacson and Baldwin (1996), it was found
that the horizontal response of a oating body to short waves
is dominated by the displacement due to the mean wave drift
force rather than the oscillatory motion. Considering the inu-
ence of such a drift motion, we infer that the mean position of
the oating breakwater may be moved to the leeward side by the
mean wave drift force. This horizontal excursion of the oat-
ing breakwater increases the tautness of seaward mooring lines,
which leads to restriction of sway oscillation because mooring
lines are not sufciently long to allow large movement.
4.2. Inuence of element-free Galerkin method
discretization
This section presents the inuence of the EFGM discretiza-
tion renement on the accuracy of the solution. Owing to the
discretization error, the numerical wave propagates with a phase
velocity /k
h
, where k
h
is the numerical wave number, differ-
ent from a given wave velocity /k. Figure 6 illustrates the
phase error between the exact and numerical wave velocity po-
tential proles. This numerical error must be reduced by appro-
priate discretization (e.g., suitable nodal distributions and size
of the domains of inuence). In the EFGM, there are selectable
parameters affecting the accuracy as included in eq. [26]. The
optimal choice is one of the major factors inuencing the perfor-
mance of the EFGM. The solution accuracy may signicantly
depend on the internodal spacing (h) and the sizes of the do-
mains of inuence (d
m
).
We have performed numerical tests for a dual-pontoon oat-
ing body to study the effects of the EFGM discretization. The
case of a dual-pontoon breakwater with two rectangular sec-
tions of width a = 2.0 m, draft D = 1.0 m, and pontoon
spacing b = 2.0 m located in a water depth d = 5.0 m is
considered. A regular wave height is 1.0 m, and the incident
wave direction is taken to be normal to the longitudinal axis.
The radiation boundary is xed at a distance of 10a from the
body. We consider a regular distribution of nodes (internodal
spacing h = 0.25 and 0.50 m). The linear basis is used, and the
sizes of the domains of inuence are identical for all nodes. The
parameter varied in the simulation is the size of the domain of
inuence, d
m
.
The numerical error was investigated by computing the scat-
tered velocity potential (
4
) distributions as a function of dis-
tance from the body obtained by numerically solving eq. [5a].
We are then able to determine the numerical wave number k
h
,
for a given value of k, by calculating the scattered wave length,
which is obtained from the scattered wave velocity potential
(
4
) distribution. The effects of the EFGMdiscretization on the
accuracy are shown in Figs. 7 and 8 for h = 0.25 and 0.50 m,
respectively. The sizes of the domain of inuence play an im-
portant role in the accuracy. For several values of the scaling
Fig. 6. Phase error between exact and numerical wave velocity
potential proles.
parameter d
max
, which is directly related to the size of the do-
mains of inuence, the phase differences between exact and nu-
merical waves increase when k increases, except for d
max
= 1.0
m. This means that large values of d
max
generate phase errors
for short waves. We also note that there exists a limiting point
of k unaffected by d
max
. When h = 0.25 m, the limiting value
of k is about 1.0 m
1
. For h = 0.5 m, large values of d
max
yield
phase errors when wave numbers are beyond 0.5 m
1
. It is clear
that a suitable value of nondimensionalized wave number (kh)
to reduce the EFGM discretization error should be less than
0.25. Similar results are obtained for the radiated potential, but
a discussion of the results is not presented in this paper.
4.3. Efciency of the numerical model
In this section, we examine the computational efciency of
the numerical model. The program based on the element-free
Galerkin method (EFGM) was run on an IBM compatible PC
Pentium IV running at 1.2 GHz with 524 MB RAM. For the
previous example given in Sect. 4.2, the central processing unit
(CPU) time in relation to the size of the domain of inuence
is shown in Table 1. As the size of the domain of inuence be-
comes larger and larger, the CPU time greatly increases. This
is because the computing time required for generation of the
system matrix increases and the bandwidth of the system ma-
trix becomes large. The comparison of EFGM and FEM com-
puting times is also made for identical node distributions and
identical numbers of degrees of freedom. For comparison, the
problem was also solved by FEM with four-node rectangular
elements, which is similar to EFGM with d
max
= 1.0 m. The
computational CPUtimes are comparedinTable 1. It is apparent
that more computational efforts are needed for EFGM than for
FEM. This is because generation of the system matrix is more
expensive and an additional computational routine is required
for determining the nodes inside the domains of inuence over
all Gaussian points. It should be noted, however, that the pre-
and post-processing tasks are signicantly reduced for EFGM
compared with those for FEM, since the mesh generation is
not necessary and the nodes can be easily moved, removed,
and added in the domain considered for EFGM. Modication
of mesh connectivity and corresponding renumbering of nodes
2003 NRC Canada
Lee and Cho 727
Fig. 7. Effect of the EFGM discretization on the accuracy
(h = 0.25 m).
Fig. 8. Effect of the EFGM discretization on the accuracy
(h = 0.5 m).
and elements should be required for FEM if the distribution of
nodes is changed.
4.4. Applications
4.4.1. Comparison of single- and dual-pontoon
breakwaters
Consideration is now given to the moored single- and dual-
pontoon oating breakwaters with the same cross section as that
used in Sect. 4.2. Except for a pontoon spacing of b = a/2,
all other conditions are the same as those given in Sect. 4.2.
The oating body with a longitudinal length of 40.0 m is re-
strained by 10 taut mooring lines. Each line has a weight per
unit length of w
0
= 183 N/m in water. The mooring line stiff-
ness coefcients proposed by Jain (1980) are used for moor-
ing analysis. The numerical results for a dual-pontoon break-
Table 1. Comparison of CPU times
Numerical method CPU time (s)
FEM 224
EFGM
d
max
= 1.0 318
d
max
= 2.0 350
d
max
= 3.0 435
d
max
= 4.0 643
water and the corresponding results for a single-pontoon break-
water are shown in Figs. 911. The hydrodynamic behavior of a
dual-pontoon breakwater is very different from that of a single-
pontoon breakwater. The most notable aspect is the occurrence
of a sharp dip and peak in heaving motion at intermediate wave
frequencies (Fig. 10). This is because of a negative added mass
in heave and an associated sharp peak in the damping coef-
cient. For the present case,
22
= 8027 kg/mat ka = 1.1. The
RAO in sway motion for a dual-pontoon section is lower be-
cause of the large mass and added mass coefcients. Figure 11
shows the variation of transmission coefcients K
t
for single-
and dual-pontoon breakwaters and reveals that a dual-pontoon
breakwater is more efcient than a single-pontoon breakwater
in attenuating wave forces at overall frequency ranges except
near the resonant frequency, ka = 1.2 to 1.4. As ka increases
to a value of 1.0, the transmission coefcients for the case of a
dual-pontoon section drop from 1.0 to 0.1 because of the neg-
ative added mass, large damping coefcients, and the out of
phase of velocity potentials. The transmission coefcients then
sharply increase near the resonant frequency, ka = 1.4. This re-
sults from the large heaving motion and the insignicant phase
difference between the incident wave potential and the radiated
potential in heaving mode.
4.4.2. Inuence of incident wave direction
The inuence of incident wave direction on the degree of
transmission has been assessed. Numerical results for the same
oatingbreakwater as that inSect. 4.4.1. are presentedinFig. 12.
The oating breakwater is subject to waves with incident angles
of 0

, 15

, 30

, and 45

. All cases show steep curves because


of a sharp dip and peak in heaving motion, as expected for a
dual-pontoon oating breakwater. The degree of wave attenu-
ation generally increases as the incident angle increases at
intermediate frequencies. Figure 12 also shows that the peak of
K
t
tends to move slightly towards higher frequency ranges.
4.4.3. Inuence of mooring condition
The computed wave transmissions and mooring forces are
compared for both taut and slack mooring lines to evaluate the
effects of mooring condition on the performance of a oating
breakwater. For the condition of a taut mooring line, the stiff-
ness coefcients proposed by Jain (1980) are included in the
equations of motion. On the other hand, for the condition of a
slack mooring line, we use the slack mooring stiffness coef-
cients derived in Appendix A. The arc length of a mooring line
above the seabed L

is set at 8.0 m, and the line weight in wa-


ter per unit length w
0
is set at 183 N/m for both taut and slack
mooring lines. The angle at the anchor point and the angle made
with the body for a taut mooring line are 5

and 50

, respec-
tively. The coordinates of attachment points of the mooring line
2003 NRC Canada
728 Can. J. Civ. Eng. Vol. 30, 2003
Fig. 9. Comparison of variation of RAOs in sway between single
and dual pontoons.
Fig. 10. Comparison of variation of RAOs in heave between
single and dual pontoons.
Fig. 11. Comparison of variation of transmission coefcients
between single and dual pontoons.
Fig. 12. Variation of transmission coefcients with incident angles
of 0

, 15

, 30

, and 45

.
with the oating structure are (x
A
, z
A
) = (2.5m, 1.0m). In
slack condition, the angle made with the body is 53.1

and the
length of the mooring line lying on the seabed l is 8.0 m, that
is, the total mooring line length L = 16.0 m.
The mooring force amplitudes F
M
can be evaluated from
the product of the mooring stiffness matrix multiplied by the
displacement vector. The results in Fig. 13 show very large
differences between taut and slack mooring force amplitudes
for all frequency ranges. This is because the stiffness of the
slack mooring line is much less than that of the taut mooring
line. For both mooring conditions, there are distinctive peaks in
mooringforces at ka = 1.4that are relatedtomaximumheaving
motion. Even though mooring forces are different from each
other, there is no signicant inuence of mooring condition for
wave attenuation as shown in Fig. 14 because the mooring line
stiffness is negligible compared with the mass and hydrostatic
stiffnesses.
Upward and lateral forces loaded to the oating body with
a slack mooring system tend to lift the mooring line from the
seabed and reduce the length lying on the seabed, which leads
to the unexpected increase in mooring force. It is therefore im-
portant that the conguration of the mooring line should be
designed to achieve safe mooring conditions, i.e., conditions
such that each mooring line at its attachment to the anchor lies
along the seabed. We calculated the maximum and minimum
lengths of a mooring line remaining on the seabed when the
oating body moves to the farthest location from the initial po-
sition. Figure 15 shows the variation in length of a mooring line
on the seabed with nondimensionalized wave number ka. For
long waves and near the resonant frequency, large variations in
l are apparent. These variations result from the large sway mo-
tion at short wave frequencies and large heave motion near the
resonant frequency. At ka = 1.4, l reaches zero, which means
that the slack mooring line changes to a taut condition. We also
performed iterative computations to obtain the critical length
l
c
for a mooring system to be in slack condition. If the length
lying on the seabed is initially less than the critical length, a
slack mooring line changes to a taut condition by movement
of the oating body. Figure 16 shows the relationship between
the critical lengths and wave frequencies. The predicted crit-
2003 NRC Canada
Lee and Cho 729
Fig. 13. Comparison of variation of mooring forces between taut
and slack conditions.
Fig. 14. Comparison of variation of transmission coefcients
between taut and slack conditions.
ical lengths are shortened at high wave frequencies except at
ka = 1.4. As the total mooring line length Lincreases, the crit-
ical length also increases. We see that the ratios of the critical
length to the total mooring line length are within the range of
0.10.6 in our case study. From Fig. 16, a sufcient length of
line lying on the seabed is required for the mooring system to
be in the slack condition for overall frequencies.
5. Summary and conclusions
This paper presents a numerical formulation and an imple-
mentation of the element-free Galerkin method (EFGM) for
calculating hydrodynamic coefcients of a moored pontoon-
type oating breakwater. The numerical model is based on
two-dimensional linear wave diffractionradiation theory for
normal and oblique incident wave interactions with a oating
breakwater. The EFGM-based numerical model is applied to
example problems to investigate the performance of a moored
pontoon-type oating breakwater. Numerical results are pre-
Fig. 15. Variation in length of a mooring line on the seabed.
Fig. 16. Critical lengths of a mooring line lying on the seabed.
sented to show the effects of wave conditions and the cong-
uration of the mooring system. In this study, simple forms of
slack mooring line stiffness coefcients are derived to assess
the effects of mooring line condition on the performance of a
oating breakwater. The main conclusions from this research
are as follows.
(1) Element-free Galerkin method parameters such as the size
of the domain of inuence and the internodal spacing play
an important role in the accuracy of the numerical solu-
tion. Large values of the sizes of the domains of inuence
generate phase errors for the case of short waves. There
also exists a limiting point of wave number k unaffected
by the size of the domain of inuence. To reduce the EFGM
discretization errors, nondimensionalized wave number kh
should be under the value of kh

= 0.25.
(2) The hydrodynamic behaviors of a dual-pontoonbreakwater
are different from those of a single-pontoon breakwater.
The most notable aspect is the occurrence of a sharp dip and
peak in heaving motion of a dual-pontoon breakwater. It is
also seen that the transmission coefcients are signicantly
2003 NRC Canada
730 Can. J. Civ. Eng. Vol. 30, 2003
affected by the heaving motion.
(3) As the incident angle increases, the degree of wave atten-
uation generally increases and the peak transmission coef-
cient at intermediate frequencies tends to move slightly
toward higher frequency ranges.
(4) Large differences between taut and slack mooring force
amplitudes for all frequency ranges are seen in our exam-
ple. However, there is no signicant inuence of mooring
condition on wave attenuation because of the small moor-
ing line stiffness relative to the mass and hydrostatic stiff-
nesses. To achieve a safe slack mooring condition, more
than half of the total mooring line length must be lying on
the seabed.
References
Au, M.C., and Brebbia, C.A. 1982. Numerical prediction of wave
forces using the boundary element method. Applied Mathematical
Modelling, 6: 218228.
Bai, K.J. 1975. Diffraction of oblique waves by an innite cylinder.
Journal of Fluid Mechanics, 68: 513535.
Bai, K.J., and Yeung, R.W. 1974. Numerical solutions to free surface
ow problems. In Proceedings of the 10th Symposium on Naval
Hydrodynamics, Cambridge, Mass., 2428 June 1974. Ofce of
Naval Research, Washington, D.C. pp. 609647.
Belytschko, T., and Tabbara, M. 1996. Dynamic fracture using
element-free Galerkin methods. International Journal for Numer-
ical Methods in Engineering, 39: 923938.
Belytschko, T., Lu,Y.Y., and Gu, L. 1994. Element-free Galerkin meth-
ods. International Journal for Numerical Methods in Engineering,
37: 229256.
Belytschko, T., Lu, Y.Y., and Gu, L. 1995. Crack propagation by
element-free Galerkin methods. Engineering Fracture Mechanics,
51(2): 295315.
Bhat, S., and Isaacson, M. 1998. Performance of twin-pontoon oating
breakwaters. In Proceedings of the 8th International Offshore and
Polar Engineering Conference, Montral, Que., 2429 May 1998.
International Society of Offshore and Polar Engineers (ISOPE),
Calif. Vol. 3, pp. 584589.
Chakrabarti, S.K. 1987. Hydrodynamics of offshore structures. Com-
putational Mechanics Publications, Southampton, U.K.
Chen, H.S., and Mei, C.C. 1974. Oscillations and save forces in a
man-made harbour in the open sea. In Proceedings of the 10th Sym-
posium on Naval Hydrodynamics, Cambridge, Mass., 2428 June
1974. Ofce of Naval Research, Washington, D.C. pp. 537596.
Du, C.J. 1999. Finite-point simulation of steady shallow water ows.
ASCE Journal of Hydraulic Engineering, 125(6): 621630.
Du, C.J. 2000. An element-free Galerkin method for simulation of
stationary two-dimensional shallow water ows in rivers. Compu-
tational Methods inApplied Mechanical Engineering, 182: 89107.
Garrison, C.J. 1978. Hydrodynamic loading of large offshore struc-
tures. Three dimensional source distribution methods. In Numeri-
cal methods in offshore engineering. Edited by O.C. Zienkiewicz,
R.W. Lewis, and K.G. Stagg. Wiley, Chichester, U.K., pp. 97140.
Hanif, M. 1983. Analysis of heaving and swaying motion of a oating
breakwater by nite element method. Ocean Engineering, 10(3):
181190.
Isaacson, M., and Baldwin, J.F. 1996. Moored structures in waves and
currents. Canadian Journal of Civil Engineering, 23(2): 418430.
Isaacson, M., and Nwogu, O.U. 1987. Wave loads and motions of
long structures in directional seas. Transactions of the ASME, 109:
126132.
Jain, R.K. 1980. A simple method of calculating the equivalent stiff-
nesses in mooring cables. Applied Ocean Research, 2: 139142.
Leonard, J.W., Huang, M.-C., and Hudspeth, R.T. 1983. Hydrody-
namic interference between oating cylinders in oblique seas. Ap-
plied Ocean Research, 5(3): 158166.
Modaressi, H., and Aubert, P. 1998. Element-free Galerkin method
for deforming multiphase porous media. International Journal for
Numerical Methods in Engineering, 42: 313340.
Nayroles, B., Touzot, G., and Villon, P. 1992. Generalizing the nite
element method: diffuse approximation and diffuse elements. Com-
putational Mechanics, 10: 307318.
Newton, R.E. 1978. Finite element analysis of two-dimensional added
mass and damping. In Finite elements in uids I. Edited by R.H.
Gallagher et al. John Wiley and Sons, NewYork, pp. 219232.
Ouatouati, A.E., and Johnson, D.A. 1999. A new approach for numer-
ical modal analysis using the element-free method. International
Journal for Numerical Methods in Engineering, 46: 127.
Sannasiraj, S.A., Sundar, V., and Sundaravadivelu, R. 1995. The hydro-
dynamic behaviour of long oating structures in directional seas.
Applied Ocean Research, 17: 233243.
Sannasiraj, S.A., Sundar, V., and Sundaravadivelu, R. 1998. Mooring
forces and motion responses of pontoon-type oating breakwaters.
Ocean Engineering, 25(1): 2748.
Sannasiraj, S.A., Sundaravadivelu, R., and Sundar, V. 2000.
Diffractionradiation of multiple oating structures in directional
waves. Ocean Engineering, 28: 201234.
Sarpkaya, T., and Isaacson, M. 1981. Mechanics of wave forces on
offshore structures. Van Nostrand Reinhold, NewYork.
Sharan, S.K. 1986. Modelling of radiation damping in uids by nite
elements. International Journal for Numerical Methods in Engineer-
ing, 23: 945957.
Suleau, S., and Bouillard, Ph. 2000. One dimensional dispersion anal-
ysis for the element-free Galerkin method for the Helmholtz equa-
tion. International Journal for Numerical Methods in Engineering,
47: 11691188.
List of symbols
a width of breakwater
b pontoon spacing
c constant that controls the relative weights
C
ji
hydrostatic stiffness matrix coefcients
d water depth
d
m
size of the domain of inuence
d
max
scaling parameter
d
s
distance between point x and node x
j
dx, dy internodal distances
D draft
f
j
complex wave exciting force amplitude
F
j
hydrodynamic forces and moment per unit length
F
e
j
wave exciting forces due to the diffraction
potential
F
m
j
motion-induced forces due to the radiation
potential
2003 NRC Canada
Lee and Cho 731
F
M
mooring force
g gravitational acceleration
h internodal spacing
H wave height
i imaginary unit
J weighted discrete least-squares norm
k wave number
k
h
numerical wave number
K
ji
mooring stiffness matrix coefcients
K
t
transmission coefcient
l length of a mooring line lying on the seabed
l
c
critical length
L total mooring line length
L

arc length of mooring line above the seabed


M
ji
mass matrix coefcients
n outward unit normal vector
N
j
(x) EFGM shape functions
N
j,x
(x) partial derivatives of the interpolation function
p hydrodynamic pressure
p(x) polynomial basis in the space x
T
= [x, z]
P load vector
q power parameter
s arc length
S
B
body boundary surface
T tension in the cable
T
h
, T
v
horizontal and vertical components of tension
T
0
horizontal component of tension
u function in domain
u
h
interpolant of u
w
0
cable weight in water per unit length
w
j
(x) weight function
X horizontal coordinate from an origin B
Z vertical coordinate from an origin B
z

vertical coordinate from an origin O

(x) vector of the coefcients


incident angle
test function corresponding to
x component of wave number k
nondimensional scattered or radiated velocity
potential

h
approximate velocity potential

j
(x, z) nondimensional radiated velocity potential

0
(x, z) nondimensional incident velocity potential

4
(x, z) nondimensional scattered velocity potential
velocity potential

j
radiation potential

0
incident potential

4
scattered potential

B
immersed body surface

f
free surface

radiation boundary

j
complex amplitude of the oscillation of the body

ji
damping coefcient

ji
added mass coefcient
uid density
angle between horizontal line and tangential line
at any point on the cable
angular frequency
domain
set of n nodal values for the velocity potential
Appendix A:
The closed form of stiffness coefcients for a slack cable has
been derived following the procedure given by Jain (1980), who
presented a simplied method of calculating the stiffnesses of
a taut mooring cable. It is assumed that the cable is inextensible
and heavy, the cable offers no resistance to bending, and the
current drag force acting on the cable is negligible.
A.1. Basic equations of cable equilibrium
For cable equilibrium, horizontal and vertical components of
tension are
[A.1] T
h
= T cos = T
0
[A.2] T
v
= T sin = w
0
s
where T is the tension in the cable, s is the arc length, w
0
is
the cable weight in water per unit length, T
0
is the horizontal
component of tension, and is the angle between horizontal
line and tangential line at any point on the cable. Cable tension
can be expressed in terms of arc length
[A.3] T = T
0
_
1 +
_
w
0
s
T
0
_
2
_
1/2
Cartesian coordinates of a point on the cable can be described
from the catenary equation
[A.4] x =
T
0
w
0
sinh
1
_
w
0
s
T
0
_
[A.5] z =
T
0
w
0
_
_
_
_
1 +
_
w
0
s
T
0
_
2
_
1/2
1
_
_
_
A.2. Stiffness coefcients for slack cable
The basic equations for the cable equilibrium are used to
derive the slack mooring line stiffnesses. In the case of the slack
condition, the lower end of the line is tangential to the seabed
and any change of tension causes the cable to partially lift from
or drop to the seabed. Thus the length of a mooring line lying
on the seabed will vary with top end movement. Figure A.1
shows the sectional sketch of a mooring cable with length L to
an anchor point B. Three sets of Cartesian coordinates (X, Z),
(x, z), and (x

, z

) are selected which have origins B, O, and


O

, respectively. The coordinates of point A with respect to the


origin O are
[A.6] x
A
=
T
0
w
sinh
1
_
L l
T
0
w
0
_
=
T
0
w
0
sinh
1
_
w
0
L

T
0
_
z
A
=
T
0
w
0
_
_
_
_
1 +
_
L l
T
0
w
0
_
2
_
1/2
1
_
_
_
[A.7]
=
T
0
w
0
_
_
_
_
1 +
_
w
0
L

T
0
_
2
_
1/2
1
_
_
_
2003 NRC Canada
732 Can. J. Civ. Eng. Vol. 30, 2003
Fig. A.1. Mooring line coordinate systems.
Z z
x , x X, B O
A
X
Z
z'
A
O
where Lis the total cable length, l is the length of a mooring line
lying on the seabed (BO), and L

= Ll. As point A moves to


A

by small movements X and Z, the position of origin O


is also changed to O

(OO

= BOBO

= l l

= l). The
coordinate of point A

with respect to O

is (x
A
+ x
A
, z
A
+
z
A
) = (x
A
+X l, z
A
+Z). Here, x
A
and z
A
are
the horizontal and vertical variations, respectively; and l is
the small amount of movement of the origin, which is equal to
the small variation in l. The small movement of point A causes
the variations in the length l and the horizontal tension T
0
. From
eqs. [A6] and [A7]
[A.8] x
A
=
dx
A
dT
0
T
0
+
dx
A
dl
l =
T
0
T
0
x
A

T
0
w
0
_
w
0
T
0
l +
w
0
L

T
2
0
T
0
__
1 +
_
w
0
L

T
0
_
2
_
1/2
[A.9] z
A
=
dz
A
dT
0
T
0
+
dz
A
dl
l =
T
0
T
0
z
A
L

_
w
0
T
0
l +
w
0
L

T
2
0
T
0
__
1 +
_
w
0
L

T
0
_
2
_
1/2
where l and T
0
are small variations in l and T
0
, respectively.
Equations [A8] and [A9] are rewritten with respect to the (X, Z) coordinate system:
[A.10] X =
T
0
T
0
(X
A
l)
T
0
w
0
_
w
0
T
0
l +
w
0
L

T
2
0
T
0
_
T
0
T
A
+l =
T
0
T
0
(X
A
l) +
_
1
T
0
T
A
_
l
L

T
A
T
0
[A.11] Z =
T
0
T
0
Z
A
L

_
w
0
T
0
l +
w
0
L

T
2
0
T
0
_
T
0
T
A
=
T
0
T
0
Z
A

w
0
L

T
A
l
w
0
L
2
T
0
T
A
T
0
where (X
A
, Z
A
) = (x
A
+ l, z
A
) is the coordinate of A with respect to B, X = x
A
+ l, Z = z
A
, and T
A
is the
initial tension at point A. From eq. [A3],
[A.12] T
A
= T
0
_
1 +
_
w
0
L

T
0
_
2
_
1/2
A.3. Stiffness components
Linearized stiffness coefcients for a slack mooring line are evaluated as follows.
A.3.1. Horizontal stiffness due to a vertical movement
The horizontal stiffness due to a vertical movement of point A is dened as
[A.13] K
12
=
T
0
Z
As X = 0,
[A.14] l =
_
T
0
T
0
(X
A
l)
L

T
0
T
A
_ _
T
A
T
A
T
0
_
Insert eq. [A14] into eq. [A11], then
Z =
T
0
T
0
Z
A
+
w
0
L

T
A
__
T
0
T
0
(X
A
l)
L

T
0
T
A
__
T
A
T
A
T
0
__

w
0
L
2
T
0
T
A
T
0
[A.15]
=
T
0
T
0
_
Z
A

w
0
L
2
T
A
+
w
0
L

T
A
T
0
_
(X
A
l)
L

T
0
T
A
__
[A.16] K
12
=
T
0
Z
= T
0
_
Z
A

w
0
L
2
T
A
+
w
0
L

T
A
T
0
_
(X
A
l)
L

T
0
T
A
__
1
2003 NRC Canada
Lee and Cho 733
Using eqs. [A7] and [A12] and z
A
= Z
A
, then
[A.17] Z
A

w
0
L
2
T
A
=
T
A
T
0
w
0

w
0
L
2
T
A
=
1
w
0
_
T
A
T
0

T
2
A
T
2
0
T
A
_
=
T
0
w
0
_
T
A
T
0
T
A
_
=
T
0
T
A
Z
A
From eqs. [A16] and [A17], K
12
can be rewritten as
[A.18] K
12
=
T
0
Z
= T
0
_
w
0
L

T
A
T
0
_
X
A
l
L

T
0
T
A
_

T
0
T
A
Z
A
_
1
A.3.2. Vertical stiffness due to a horizontal movement
The stiffness component K
21
, vertical stiffness due to a horizontal movement, is dened as
[A.19] K
21
=
T
v
X
As Z = 0,
[A.20] l =
T
A
w
0
L

_
T
0
T
0
Z
A

w
0
L
2
T
A
T
0
T
0
_
=
T
A
w
0
L

T
0
T
0
_
Z
A

w
0
L
2
T
A
_
=
T
0
w
0
L

Z
A
=
T
0
w
0
L

_
T
A
T
0
w
0
_
Insert eq. [A20] into eq. [A10]; then
X =
T
0
T
0
(X
A
l)
_
T
A
T
0
T
A
_
T
0
w
0
L

Z
A

T
A
T
0
[A.21]
=
T
0
w
0
_
w
0
T
0
_
X
A
l
L

T
0
T
A
_

_
T
A
T
0
T
0
L

_
T
0
T
A
Z
A
_
=
T
0
w
0
__
T
A
T
0
T
0
L

___
w
0
L

T
A
T
0
__
X
A
l
L

T
0
T
A
_

T
0
T
A
Z
A
__
From eq. [A2],
[A.22] K
21
= w
0
(l)
X
= T
0
_
w
0
L

T
A
T
0
_
X
A
l
L

T
0
T
A
_

T
0
T
A
Z
A
_
1
= K
12
A.3.3. Horizontal stiffness due to a horizontal movement
The horizontal stiffness resulting from a horizontal movement, K
11
, can be calculated from eqs. [A18] and [A21] as
follows.
As Z = 0,
[A.23] K
11
=
T
0
X
=
_
w
0
L

T
A
T
0
_
K
12
A.3.4. Vertical stiffness due to a vertical movement
From X = 0 and eqs. [A14] and [A18],
[A.24] K
22
=
T
v
Z
= w
0
(l)
Z
= w
0
(l)
T
0
T
0
Z
=
w
0
T
0
_
T
A
T
A
T
0
__
X
A
l
T
0
L

T
A
_
K
12
2003 NRC Canada

Vous aimerez peut-être aussi