Anda di halaman 1dari 16

ARTICLE IN PRESS

Ocean Engineering 34 (2007) 15161531 www.elsevier.com/locate/oceaneng

Hydroelastic analysis of exible oating interconnected structures


Shixiao Fua,b,, Torgeir Moana, Xujun Chena, Weicheng Cuib,c
a b

Centre for Ships and Ocean Structures, Norwegian University of Science and Technology, N-7491 Trondheim, Norway School of Naval Architecture, Ocean and Civil Engineering, Shanghai Jiao Tong University, Shanghai 200030, China c China Ship Scientic Research Center, Wuxi 214082, China Received 6 November 2006; accepted 9 January 2007 Available online 3 February 2007

Abstract Three-dimensional hydroelasticity theory is used to predict the hydroelastic response of exible oating interconnected structures. The theory is extended to take into account hinge rigid modes, which are calculated from a numerical analysis of the structure based on the nite element method. The modules and connectors are all considered to be exible, with variable translational and rotational connector stiffness. As a special case, the response of a two-module interconnected structure with very high connector stiffness is found to compare well to experimental results for an otherwise equivalent continuous structure. This model is used to study the general characteristics of hydroelastic response in exible oating interconnected structures, including their displacement and bending moments under various conditions. The effects of connector and module stiffness on the hydroelastic response are also studied, to provide information regarding the optimal design of such structures. r 2007 Elsevier Ltd. All rights reserved.
Keywords: Interconnected structures; Hinge modes; Linear hydroelasticity; Frequency domain; Bending moment; Displacement

1. Introduction Very large oating structures (VLFS) can be used for a variety of purposes, such as airports, bridges, storage facilities, emergency bases, and terminals. A key feature of these exible structures is the coupling between their deformation and the uid eld. A variety of VLFS hull designs have emerged, including monolithic hulls, semisubmersible hulls, and hulls composed of many interconnected exible modules. Various theories have been developed in order to predict the hydroelastic response of continuous exible structures. For simple spatial models such as beams and plates, one-, two- and three-dimensional hydroelasticity theories have been developed. Many variations of these theories have been adopted using both analytical formulations (Sahoo et al., 2000; Sun et al., 2002; Ohkusu, 1998) and numerical
Corresponding author. Centre for Ships and Ocean Structures, Norwegian University of Science and Technology, N-7491 Trondheim, Norway. Tel.: +47 73551108. E-mail address: shixiao.fu@ntnu.no (S. Fu).

methods (Wu et al., 1995; Kim and Ertekin, 1998; Ertekin and Kim, 1999; Eatock Taylor and Ohkusu, 2000; Eatock Taylor, 2003; Cui et al., 2007). Specic hydrodynamic formulations based on the modal representation of structural behaviour, traditional three-dimensional seakeeping theory, and linear potential theory have been developed to predict the response of both beam-like structures (Bishop and Price, 1979) and those of arbitrary shape (Wu, 1984), through application of two-dimensional strip theory and the three-dimensional Greens function method, respectively. Other hydroelastic formulations also exist based upon two-dimensional (Wu and Moan, 1996; Xia et al., 1998) and three-dimensional nonlinear theory (Chen et al., 2003a). Finally, several hybrid methods of hydroelastic analysis for the single module problem have also been developed (Hamamoto, 1998; Seto and Ochi, 1998; Kashiwagi, 1998; Hermans, 1998). To predict the hydroelastic response of interconnected multi-module structures, multi-body hydrodynamic interaction theory is usually adopted. In this theory, both modules and connectors may be modelled as either rigid or exible. There are, therefore, four types of model: Rigid

0029-8018/$ - see front matter r 2007 Elsevier Ltd. All rights reserved. doi:10.1016/j.oceaneng.2007.01.003

ARTICLE IN PRESS
S. Fu et al. / Ocean Engineering 34 (2007) 15161531 1517

Module and Rigid Connector (RMRC), Rigid Module and Flexible Connector (RMFC), Flexible Module and Rigid Connector (FMRC) and Flexible Module and Flexible Connector (FMFC). By adopting two-dimensional linear strip theory, ignoring the hydrodynamic interaction between modules, and using a simplied beam model with varying shear and exural rigidities, Che et al. (1992) analysed the hydroelastic response of a 5-module VLFS. Che et al. (1994) later extended this theory by representing the structure with a three-dimensional nite element model rather than as a beam. Various three-dimensional methods (in both hydrodynamics and structural analysis) have been developed using source distribution methods to analyse RMFC models (Wang et al., 1991; Riggs and Ertekin, 1993; Riggs et al., 1999; Cui et al., 2007). These formulations account for the hydrodynamic interactions between each module by considering the radiation conditions corresponding to the motion of each module in one of its six rigid modes, while keeping the other modules xed. By employing the composite singularity distribution method and three-dimensional hydroelasticity theory, Wu et al. (1993) analysed the hydroelastic response of a 5-module VLFS with FMFC. Riggs et al. (2000) compared the wave-induced response of an interconnected VLFS under the RMFC and FMFC (FEA) models. They found that the effect of module elasticity in the FMFC model could be reproduced in a RMFC model by changing the stiffness of the RMFC connectors to match the natural frequencies and mode shapes of the two models. The methods considered so far deal with modules joined by connectors at both deck and bottom levels, so that there is no hinge modes existed, or all the modules are considered to be rigid. In a structure composed of serially and longitudinally connected barges, Newman (1997a, b, 1998a) explicitly dened hinge rigid body modes to represent the relative motions between the modules and the shear force loads in the connectors (WAMIT; Lee and Newman, 2004). In addition to accounting for hinged connectors, modules can be modelled as exible beams (Newman, 1998b; Lee and Newman, 2000; Newman, 2005). Using WAMIT and taking into account the elasticity of both modules and connectors, Kim et al. (1999) studied the hydroelastic response of a ve-module VLFS in the linear frequency domain, where the elasticity of modules and connectors is modelled by using a structural three-dimensional FE modal analysis, and the hinge rigid modes are explicitly dened following Newman (1997a, b) and Lee and Newman (2004). When it comes to the more complicated interconnected multi-body structures, composed of many exible modules that need not be connected serially, it will become very difcult to explicitly dene the hinge modes of rigid relative motion and shear force. In particular, it is difcult to ensure that the orthogonality conditions of the hinge rigid modes are satised with respect to the other exible and rotational rigid modes. The purpose of this paper is to

demonstrate a method of predicting the hydroelastic response of a exible, oating, interconnected structure using general three-dimensional hydroelasticity theory (Wu, 1984), extending previous work to take into account hinge rigid modes. These modes are calculated through a numerical analysis of the structure based on the nite element method, rather than being explicitly dened to meet orthogonality conditions. All the modules and connectors are considered to be exible. The translational and rotational stiffness of the connectors is also considered. This method is validated by a special numerical case, where the hydroelastic response for very high connector stiffness values is shown to be the equivalent to that of a continuous structure. Using the results of this test model, the hydroelastic responses of more general structures are studied, including their displacement and bending moments. Moreover, the effect of connector and module stiffness on the hydroelastic response is studied to provide insight into the optimal design of such structures. 2. Equations of motion for freely oating exible structures Using the nite element method, the equation of motion for an arbitrary structural system can be represented as g C fU _ g K fU g fPg, M fU (1)

where [M], [C] and [K] are the global mass, damping and stiffness matrices, respectively; {U} is the nodal displacement vector; and {P} is the vector of structural distributed forces. All of these entities are assembled from the corresponding single element matrices [Me], [Ce], [Ke], {Ue}, and {Pe} using standard FEM procedures. The connectors are modeled by translational and rotational springs, and can be incorporated into the motion equations using standard FEM procedures. Neglecting all external forces and damping yields the free vibration equation of the system: g K fU g f0g. M fU (2)

Assuming that Eq. (2) has a harmonic solution with frequency o, this then leads to the following eigenvalue problem: o2 M K fDg f0g. (3)

Provided that [M] and [K] are symmetric and [M] is positive denite, and that [K] is positive denite (for a system without any free motions) or semi-denite (for a system allowing some special free motions), all the eigenvalues of Eq. (3) will be non-negative and real. The eigenvalues o2 r r 1; 2; 3; . . . ; 6n represent the squared natural frequencies of the system:
2 2 0po2 1 po2 ; ; po6n ,

(4) o2 r X0

where when [K] is positive denite, and when [K] is semi-denite. Each eigenvalue is associated with a real eigenvector {Dr}, which represents the rth

o2 r 40

ARTICLE IN PRESS
1518 S. Fu et al. / Ocean Engineering 34 (2007) 15161531

natural mode: fDr g ffDr1 g; fDr2 g ; . . . ; fDrj g ; . . . ; fDrn ggT , (5)

where fDri g is the eigenvector of the ith node which contains 6 degree of freedoms, and i runs over the n nodes of the structural FE model system. fd r g, a sub-matrix of fDr g, consists of the rth natural mode components of all the nodes associated with one particular element. The rth modal shape {ur} at any point in that element can be expressed as
* ur

conditions with respect to K gM and [M] are automatically satised in Eq. (8). Thus, these also can satisfy the orthogonality conditions with respect to [K] and [M] for the original interconnected structure. This means that eigenvalues and eigenvectors of the original system can therefore be expressed as o2 l g, fDr g fX r g. (11) (12)

l N Lfd r g fur ; vr ; wr g ,

(6)

where [L] is a banded, local-to-global coordinate transform matrix composed of diagonal sub-matrices [l], each of which is a simple cosine matrix between two coordinates. [N] is the displacement interpolation function of the structural element. For freely oating, hinge-connected, multi-module structures, Eq. (3) has zero-valued roots corresponding to the 6 modes of global rigid motion and the hinge modes describing relative motion between each module. According to traditional seakeeping theory, the rigid modes of the global system can be described by three translational components (uG, vG, wG) and three rotational components (yxG, yyG, yzG) about the center of mass in the global coordinate system coincident with equilibrium. Thus, the rst six rigid modes (with zero frequency) at any point j on the freely oating body can be expressed by fD 1 j g f 1 ; fD 2 j g f 0 ; fD 3 j g f 0 ; fD 4 j g f 0 ; 0; 1; 0; 0; 0; 1; 0; 0; 0; 0; 0; 0; 0; 0 gT ; 0 gT ; 0 gT ; 1; 0; 0; 0; 1; 0; 0 gT ; 0 gT ; 1 gT : (7)

Since usually only the rst several oscillatory modes dominate the structural dynamic response, we assume that the nodal displacement of the structure can written as a superposition of the rst m modes, fU g
m X fDr gpr t Dfpg, r1

(13)

where pr(t) refers to the rth generalized coordinate. For r 16, {Dr} represents the vector of the rst six rigid modes and pr(t) the magnitude of rigid displacement about the center of mass (xG, yG, zG). Substituting (13) into (1) and premultiplying by [D]T, the generalized equation of motion is as follows: g bfp _ g cfpg fZg, afp with a DT M D, b DT C D, c DT K D, fZg DT fPg fZ 1 ; Z2 ; . . . ; Zm g. 15 (14)

z zG ;

y yG ; x xG ; 0;

fD 5 j g f z z G ; fD6j g f y yG ;

x xG ;

These vectors correspond to the six rigid motions of the global structure: surge, sway, heave, roll, pitch and yaw, where (x, y, z) and (xG, yG, zG) are the coordinates of a point in the oating body and the center of mass, respectively. To obtain the zero-frequency hinge modes describing the relative motion between different modules, we transform the eigenvalue problem into a new one by introducing an additional articial stiffness proportional to the mass, g[M] where g can be non-zero articial real number close to the rst non-zero eigenvalue of the system. Then we have lM K 0 fX g f0g, where l o g, K 0 K gM .
2

[a], [b] and [c] are the generalized mass, damping and stiffness matrices respectively; {Z} is the generalized distributed force and can be expressed as Zr fur gT fPg. (16)

In general, the generalized coordinates {p} in Eq. (14) separate naturally into two groups, which can be denoted by {pR} and {pD} respectively, that is to say fp g fpR ; p D gT , where fp R g fp 1 ; p 2 ; . . . ; p 6 gT (18) (17)

(8)

(9) (10)

refers to the rigid body modes of the global structure as dened by Eq. (7) and fpD g fp 7 ; p 8 ; . . . ; p m gT (19)

From Eq. (8) we can get the corresponding positive eigenvalues l and eigenvectors {X}. The orthogonality

refers to the distortion modes, including both rigid hinge modes and structural distortional modes. Eq. (14) can

ARTICLE IN PRESS
S. Fu et al. / Ocean Engineering 34 (2007) 15161531 1519

therefore be written as " #( ) " R p aR 0 0 D p 0 aD 0 " 0 0

3. Potential theory and freely oating exible structures 0 bD 0 cD #( #( _R p _D p pR pD ) ) ( ) , 20 The uid around a exible freely oating body is assumed to be ideal (i.e., uniform, continuous, inviscid, incompressible and irrotational). Hence, the uid behaviour can be described by the velocity potential. In order to simplify the expressions two coordinate systems are introduced, namely the equilibrium frame Oxyz and the system of xed body axes O0 x0 y0 z0 . The origin of the Oxyz system is the point of intersection between the still water surface and the vertical line, which goes through the centre of gravity of the structure, which is also the axis Oz. The O0 x0 y0 z0 system is xed on the oating body. Assuming an ideal uid eld, the velocity potential (unsteady velocity potential) around a oating structure with zero forward speed in the equilibrium frame may be decomposed into the form (Wu, 1984): fx; y; z; t fI x; y; z; t fD x; y; z; t m X pr fr x; y; z; t,
r1

ZR ZD

where 0 is the zero matrix, aD and cD are diagonal matrices, bD is a square symmetric matrix, and aR, the inertia matrix of a rigid body, has the form 2 3 rr 0 0 " # am 0 6 7 0 rr 0 7 aR ; am 6 4 5, 0 aI 0 0 rr 2 3 I 44 I 45 I 46 6 7 7 21 aI 6 4 I 54 I 55 I 56 5. I 64 I 65 I 66 In Eq. (21), r is the density of the uid and r is the displacement of the whole structure. The components of aI are given by ZZZ ij I ij I ji 1 rb x xG i3 x xG j3 dV . (22)
V

27

It is convenient to use a viscous damping model. A further simplication option is to represent the structural damping by a Rayleigh or proportional damping model (e.g. Bathe and Wilson, 1976): C aK bM , (23)

where fI x; y; z; t, fD x; y; z; t and fr x; y; z; t denote the incident wave potential, diffraction wave potential, and radiation wave potential arising from the responses of the exible body. In the frequency domain, the rst-order unsteady velocity potential and the principal coordinates may be ( further expressed as # ) " m X f Re jI jD jr pr eiot , (28)
r1

pr t Refpr e g,

iot

(29)

where a and b are the stiffness and mass proportional damping constants, respectively. These can be associated with fractions of the critical damping x as follows: x 0:5ao b=o. (24)

a and b can be determined by choosing the fractions of critical damping (x1 and x2) at the rst two distinct nonzero frequencies (o1 and o2). The two resulting equations can then be inverted and solved:
2 a 2x2 o2 x1 o1 =o2 2 o1 , 2 b 2o1 o2 x1 o2 x2 o1 =o2 2 o 1 .

25

The degree of damping at other frequencies is obviously only approximated by this solution. Finally, the dynamic response of the discretised structure can be found by solving Eq. (20) to obtain the principal coordinate value of each mode. This gives ut
m X r1

where o is the wave circular frequency; jI and jD are components of the incident wave velocity potential and the diffraction wave potential respectively; jr (r 1,y,m) the component of the radiation wave potential arising from vibration in the rth principal dry mode of the exible body, with unit amplitude and frequency o; and pr the complex amplitude of the generalized coordinate. The diffraction potential and radiation potential individually satisfy the following boundary value problem: 8 in O : r2 f 0; > > > > < on S F : o2 f g qf 0; qz (30) qf > on S : 0 ; > B qz > > : on S : sommerfeld radiation condition: 1

fur gpr t,

(26)

where u and {ur} are the displacement and rth mode shape of the structural system, respectively.

Fig. 1. Denition of the uid and structure boundaries.

ARTICLE IN PRESS
1520 S. Fu et al. / Ocean Engineering 34 (2007) 15161531

As shown in Fig. 1, O is the uid domain around the exible body; SF, SB and SN are the free surface, bottom surface, and boundary surface at innity of the uid, respectively. Furthermore, the boundary conditions on the wetted body surface S for the diffraction and radiation potentials can be written as qfD qf I qn qn and qfr * * io u r n , qn
*

where w(t) is the vertical displacement of the oating structure at time t. The rth component of the generalized uid force on the wetted surface can be expressed as ZZ * * n u r p dS . (34) Zr
S

(31)

Substituting Eqs. (33) and (34) into Eq. (14), we have


m X k1

k o2 ark p r arr pr r 1; 2; . . . ; m ,

m X k 1

_ k F r E r Rr brk p 35

(32)

where n represents the outward-directed unit vector * normal to the wetted surface of the body, u r represents the rth modal shape vector of the corresponding point on the wetted surface, and g is the acceleration due to gravity. The diffraction and radiation potentials jD and jr can be determined using the three-dimensional Greens function method, where the source strength distribution over the wetted surfaces S can be determined numerically. The dynamic uid pressure acting on the mean wetted surface S (averaging over motions and distortions of the body) is given by the linearised Bernoulli equation in the equilibrium coordinate system:   qf px; y; z; tjS r gwt . (33) qt S

where Fr, Er and Rr are the rth generalized wave excitation forces, radiation forces and restoring forces induced by the displacement of the structure, respectively. These forces have the following forms: ZZ * * Fr r n u r iojI jD dS eiot , (36)
S

ZZ Er r
S

* * n ur

io

m X k1

pk tjk dS eiot ,

(37)

ZZ Rr r
S

n u r gw dS.

(38)

300m P5 P4 60m

P1

P2

P3

(0,0)

connectors
Fig. 2. Schematic plane view of a two-module, interconnected structure.

Table 1 Natural frequency and generalized mass of the structure for different stiffness values of the connector and bending stiffness Mn. EI 4:77E 11; krot 0.0 NF 7 8 9 10 11 12 13 14 0.0 0.430 0.500 0.624 1.017 1.396 1.581 1.722 GM 2.619 2.292 1.500 2.283 1.412 2.236 1.252 2.217 0.159 109 NF 0.115 0.430 0.500 0.720 1.019 1.396 1.581 1.843 GM 2.602 2.292 1.500 2.368 1.404 2.236 1.252 2.257 0.159 1011 NF 0.155 0.430 0.500 0.841 1.024 1.396 1.581 2.080 GM 2.308 2.292 1.500 2.274 1.384 2.236 1.252 2.187 0.156 0.430 0.507 0.845 1.036 1.398 1.604 2.090 2.303 2.292 1.500 2.270 1.383 2.236 1.252 2.183 Continuous structure NF GM krot 0:0; EI 4.77 1010 NF 0.0 0.136 0.158 0.197 0.322 0.442 0.500 0.545 4.77 1012 NF 0.0 1.360 1.582 1.973 3.216 4.415 5.000 5.446 4.77 1013 NF 0.0 4.302 5.001 6.239 10.169 13.962 15.809 17.221

Mn. denotes the Mode Number, EI denotes the bending stiffness of the modulus in N m2, Krot denotes the rotational stiffness of the connector in N m, NF. denotes the Natural Frequency in Hz, and GM. denotes the Generalized Mass 106.

ARTICLE IN PRESS
S. Fu et al. / Ocean Engineering 34 (2007) 15161531 1521

The restoring force can be expressed as Rr with ZZ C rk r


S * * n ur m X k1

pk C rk eiot

(39)

where [a], [b] and [c] are the generalized mass, damping and stiffness matrices of the oating structure respectively and [A], [B] and [C] are the generalized added mass, added damping and restoring coefcient matrices of the uid respectively. Their components are given by > > 1 = < ZZ * *0 Ark o o 2 Re , i r n u oj o d S i k r Brk o o Im > > ; : S 8 9

gwk dS ,

(40)

(42)

where the Crk are generalized restoring coefcients. When rp6 and kp6, Eq. (40) reduces to the restoring coefcient matrix of rigid motion. The other generalized coefcients can be obtained similarly. Placing all these coefcients into Eq. (14) yields the linear generalized hydroelastic equation of motion g b Bfp _ g c C fpg fF g, a Afp
Table 2 The mode shapes of hinged and continuous structure Mn 7 Mode shape with krot 0.0

(41)

where Crk are the frequency-independent coefcients of the generalized rst-order forces, as expressed in Eq. (40). Finally, the dynamic displacement response of any point on the structure can be found from Eq. (20). With given displacements, all other structural parameters such as bending

Mode shape for the continous structure

10

11

12

13

14

ARTICLE IN PRESS
1522 S. Fu et al. / Ocean Engineering 34 (2007) 15161531

moments, shearing forces, twisting moments and stresses can be determined. For example, the bending moment at a certain section of the structure becomes I M S DB k dl s , (43) where DB is the bending stiffness of the structure in a certain direction, k is the curvature of structure deformation at a given point in that direction (Chen et al., 2003b), and dls is the length of the element included in the structure cross section.

4. Numerical example 4.1. The model To illustrate this methodology, we describe a numerical example based on the continuous pontoon-type VLFS mentioned by Sim and Choi (1998). The water depth is 58.5 m, and the length, width, height, draught and mass of the VLFS are 300, 60, 2, 0.5 m and 9.225 106 kg, respectively. This VLFS is a scaled model of the Mega-Float, constructed and developed by Yokosuka (Yago and Endo, 1996) for use in sheltered waters. Sim

a
0.8 Vertical Displacement (w/A) 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0.0 0.0 0.1 0.2 0.3 0.4 0.5 x/L 0.6 0.7 0.8 0.9 1.0
=60m (/L=0.2), =0 Exp. of Continuous Structure Krot=0.159E11 (Continuous) Krot=0 Krot=0.159E9

b
1.2 Vertical Displacement (w/A) 1.0 0.8 0.6 0.4 0.2 0.0
=120m (/L=0.4), =0 Exp. of Continuous Structure Krot=0.159E11 (Continuous) Krot=0 Krot=0.159E9

0.0

0.1

0.2

0.3

0.4

0.5 x/L

0.6

0.7

0.8

0.9

1.0

c
1.6 1.4 Vertical Displacement (w/A) 1.2 1.0 0.8 0.6 0.4 0.2 0.0 0.0 0.1 0.2 0.3 0.4 0.5 x/L 0.6 0.7 0.8 0.9 1.0
=180m (/L=0.6), =0

Exp. of Continuous Structure


Krot=0.159E11 (Continuous) Krot=0 Krot=0.159E9

Fig. 3. Vertical response amplitude of the longitudinal centerline with constant module bending stiffness (EI 4:77 1011 ) and variable connector rotational stiffness.

ARTICLE IN PRESS
S. Fu et al. / Ocean Engineering 34 (2007) 15161531 1523

d
1.8 1.6 Vertical Displacement (|w|/A) 1.4 1.2 1.0 0.8 0.6 0.4 0.2 0.0 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
=240m (/L=0.8), =0 Exp. of Continuous Structure Krot=0.159E11 Krot=0 Krot=0.159E9

x/L

e
2.0 1.8 Vertical Displacement (|w|/A) 1.6 1.4 1.2 1.0 0.8 0.6 0.4 0.2 0.0 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
=300m (/L=1.0), =0 Exp. of ContinuousStructure Krot=0.159E11 Krot=0 Krot=0.159E9

x/L
Fig. 3. (Continued)

and Choi (1998) estimated its bending rigidity to be EI 4:77 1011 N m2 . As there is not enough detailed structural information on interconnected multi-module structures in the open literature, we create an interconnected structure model by dividing the above-mentioned continuous VLFS model into two halves. The nature of the connection is shown in Fig. 2, which also illustrates the unit amplitude incident wave direction. The connectors at the middle of longitudinal direction of the structure are parallel to the width direction with equivalent space. Each connector is characterized by stiffness values in the x, y, and z directions and a rotational stiffness value associated with the relative pitch of the adjacent modules, which are denoted kx, ky, kz and krot, respectively. In this paper, each translational connector stiffness is taken to be constant at kx ky kz 4:6065 1012 N m. It is of course very difcult to actually dene the structural damping realistically. The fractions of critical damping in the rst two distinct non-zero frequency modes are chosen to be x1 x2 5%. We have tried different damping ratio from 1% to 6%. The results show that for our case the damping ratio has negligible effects on the responses. Totally, 4500 shell elements with 2 m 2 m are used to model the dry structure, considering membrane and bending forces simultaneously (Hibbit et al., 2004). The ABAQUS (Hibbit

et al., 2004) is used as the eigen-solver. Similar panel elements are used to model the wetted surfaces of the two modules. Five points are selected to analyse the characteristics of hydroelastic structural response, marked as P1, P2, P3, P4 and P5 in Fig. 2 with coordinates (0,30), (75,30), (150,30), (225,30) and (0,60), respectively. 4.2. Modal analysis of the dry structure Table 1 shows the natural frequency and generalized mass of the structure for different values of the connector and the module stiffness. We should note that if the rotational stiffness of the connector is zero, then for a given mode the structure will have the same mode shape and generalized mass no matter what the stiffness of the module is. Consequently, in Table 1 only one generalized mass is listed for each mode for the krot 0.0 case. The corresponding mode shapes for the hinge-connected and continuous structures can be found in Table 2. As can be seen from Tables 1 and 2, increasing the rotational stiffness of the connector tends to equalize the natural frequency and generalized mass values for the hinge-connected and continuous structures. For a given module stiffness, the existence of a hinge connector alters the natural frequencies and shapes of the vertical bending

ARTICLE IN PRESS
1524 S. Fu et al. / Ocean Engineering 34 (2007) 15161531

modes more than for the torsional modes. Theoretically, the rotational stiffness of the connectors needs to approach innity for the interconnected and continuous structures to have fully equivalent stiffness. These data show that for a rotational stiffness of 0:159 1011 N m, however, the natural frequencies and mode shapes of the interconnected structure are very close to the corresponding continuous values. The interconnected structure with a rotational stiffness of 0:159 1011 N m can therefore be considered to respond similarly to the continuous structure. Unless otherwise noted, in the following numerical results a superposition of the rst 35 modes is used in the hydroelastic analysis. 4.3. Numerical results Fig. 3 shows the vertical displacement amplitude along the longitudinal centerline for experimental and numerical results. The agreement between experimental and numerical results for the high-stiffness, continuous structure is good. The corresponding validation for a continuous model has also been made by comparing the results of the present method, WAMIT and experiments (Taghipour et al., 2006). As there are no experimental results available

for a exible oating interconnected structure to validate our methodology, we rely on well-designed numerical simulations. Fig. 3 shows the vertical displacement amplitude along the longitudinal centerline for different values of krot and wavelength l where the translational stiffness of the connectors is kept as constant. Fig. 4 shows the vertical displacement amplitude of two different points as a function of the incident wave frequency. In both Figs. 3 and 4, the bending stiffness of the structure is 4:77 1011 Nm2 . As the rotational stiffness increases, the predicted hydroelastic response of the interconnected structure using the methodology described in this paper should approach the continuous structure response and experimental results. This fact can be observed in all the panels of Fig. 3. Hence, the methodology presented in this paper can also be used to simulate the hydroelastic response of exible interconnected oating structures. Figs. 3 and 4 show that as the rotational stiffness of the connector decreases the vertical response amplitude of the connecting points will increase, reaching its maximum value when the rotational stiffness is zero. It can also be seen that the rotational stiffness of the connectors has more effect on the vertical displacement of points near the

a
1.4 Vertical Displacement (|w|/A) 1.2 1.0 0.8 0.6 0.4 0.2 0.0 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 Incident Wave Freq. (rad/s) 1.6 1.8 2.0
Vertical Disp.Amp. of P1 Krot=0.159E11 Krot=0.159E9 Krot=0

b
1.4 Vertical Displacement (|w|/A) 1.2 1.0 0.8 0.6 0.4 0.2 0.0 0.0
Vertical Disp.Amp. of P3 Krot=0.159E11 Krot=0.159E9 Krot=0

0.2

0.4

0.6

0.8 1.0 1.2 Incident Wave Freq. (rad/s)

1.4

1.6

1.8

2.0

Fig. 4. The effect of connector rotational stiffness on the vertical response, with constant bending stiffness (EI 4:77 1011 ). (a) Vertical displacement amplitude at point P1 in Fig. 2; (b) Vertical displacement amplitude at point P3 in Fig. 2.

ARTICLE IN PRESS
S. Fu et al. / Ocean Engineering 34 (2007) 15161531 1525

connectors than on other positions. Moreover, the vertical response amplitude on the upstream side of the structure is always larger than that on the downstream side. The dynamic response of the fully hinged structure (when the rotational stiffness of the connectors is zero) will be very different from that of the continuous structure. The hinge connectors increase the vertical displacement of connecting points beyond the wave amplitude, a phenomenon, which is never experienced in a continuous structure.

The effect of the incident wave frequency and module stiffness on the dynamic response amplitude and the generalized coordinate amplitudes of each mode are shown in Figs. 5 and 6, respectively, where the rotational stiffness of the structure is set to zero. In Figs. 5 and 6, the term Only Rigid Modes are Considered means that all the modules are assumed to be rigid, so only the rst 7 rigid modes are taken into account in the hydroelastic analysis, rather than the rst 35.

a
1.8 Vertical Displacement (|w|/A) 1.6 1.4 1.2 1.0 0.8 0.6 0.4 0.2 0.0 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0 2.2
Vertical Disp.of P1 EI=4.77E13 EI=4.77E12 EI=4.77E11 EI=4.77E10 Only Rigid Modes are Considered

Incident W ave Freq. (rad/s)

b
1.6 Vertical Displacement (|w|/A) 1.4 1.2 1.0 0.8 0.6 0.4 0.2 0.0 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0 2.2
Vertical Disp. of P3 EI=4.77E13 EI=4.77E12 EI=4.77E11 EI=4.77E10 Only Rigid Modes are Considered

Incident W ave Freq. (rad/s)

c
1.8 Vertical Displacement (|w|/A) 1.6 1.4 1.2 1.0 0.8 0.6 0.4 0.2 0.0 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0 2.2
Vertical Disp. of P5 (=45) EI=4.77E13 EI=4.77E12 EI=4.77E11 EI=4.77E10 Only Rigid Modes are Considered

Incident W ave Freq. (rad/s)

Fig. 5. The effect of module bending stiffness on vertical displacement, for connectors with zero rotational stiffness. Only Rigid Modes are Considered means that all the modules are assumed to be rigid, so only the rst 7 rigid modes are taken into account in the hydroelastic analysis. (a) Vertical displacement amplitude at point P1 in Fig. 2; (b) Vertical displacement amplitude at point P3 in Fig. 2; (c) Vertical displacement amplitude at point P5 in Fig. 2.

ARTICLE IN PRESS
1526 S. Fu et al. / Ocean Engineering 34 (2007) 15161531

As can be seen from Figs. 5 and 6, the dynamic response amplitude of points P1 and P3 is greatly affected by the bending stiffness of the module. The vertical amplitude has its maximum value when the modules are assumed to be rigid and only the rst 7 rigid modes are included. In the domain of relatively low-incident wave frequencies, in-

creasing the module stiffness will generally lead to higher response amplitudes (the exibility of the modules reduces the response amplitude of the structures). In the domain of high-incident wave frequencies, on the other hand, increasing the module stiffness will decrease the response amplitude. This can be explained by the fact that in the

a
1.2 1.0 0.8 |P3|/A 0.6 0.4 0.2 0.0 0.0
EI=4.77E13 EI=4.77E12 EI=4.77E11 EI=4.77E10 Only Rigid Modes are Considered

0.2

0.4

0.6

0.8

1.0

1.2

1.4

1.6

1.8

2.0

Incident Wave Freq. (rad/s)

b
2.2 2.0 1.8 1.6 1.4 |P7|/A 1.2 1.0 0.8 0.6 0.4 0.2 0.0 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0
EI=4.77E13 EI=4.77E12 EI=4.77E11 EI=4.77E10 Only Rigid Modes are Considered

Incident Wave Freq. (rad/s)

c
1.6 1.4 1.2 1.0 |P8|/A 0.8 0.6 0.4 0.2 0.0 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0
EI=4.77E13 EI=4.77E12 EI=4.77E11 EI=4.77E10

Incident Wave Freq. (rad/s)

Fig. 6. The effect of bending stiffness on the generalized coordinate response amplitude, for connectors with zero rotational stiffness. Only Rigid Modes are Considered means that all the modules are assumed to be rigid, so only the rst 7 rigid modes are taken into account in the hydroelastic analysis. (a) Generalized coordinate response of mode 3; (b) Generalized coordinate response of mode 7; (c) Generalized coordinate response of mode 8; (d) Generalized coordinate response of mode 10; (e) Generalized coordinate response of mode 12; (f) Generalized coordinate response of mode 14.

ARTICLE IN PRESS
S. Fu et al. / Ocean Engineering 34 (2007) 15161531 1527

1.4 1.2 1.0


EI=4.77E13 EI=4.77E12 EI=4.77E11 EI=4.77E10

|P10|A

0.8 0.6 0.4 0.2 0.0 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0 Incident Wave Freq. (rad/s)

0.8
EI=4.77E13 EI=4.77E12 EI=4.77E11 EI=4.77E10

0.6

|P12|/A

0.4

0.2

0.0 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0 Incident Wave Freq. (rad/s)

0.5
EI=4.77E13 EI=4.77E12 EI=4.77E11 EI=4.77E10

0.4

0.3 |P14|/A

0.2

0.1

0.0 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0 Incident Wave Freq. (rad/s)

Fig. 6. (Continued)

low-frequency domain diffraction wave forces and rigid modes dominate the response of the structure, while in the high-frequency domain radiation wave forces and exible modes dominate the response. Figs. 6(a) and (b) show that the stiffness of the modules will not affect the response amplitude of rigid body modes; that is to say, the rigid modes will contribute equally to the dynamic response of the interconnected structure whether it is assumed to be rigid or exible. As can be

seen in Fig. 6(cf), increasing the stiffness of each module decreases the response amplitude of each exible mode; the more exible the module is, the greater the contribution its exibility will make. This contribution will not always increase the nal response amplitude of the structure, however, which depends strongly on the relative phase angle of the response. This behaviour can be seen by comparing Figs. 5 and 6. The torsional modes will not be excited in head sea, then the corresponding

ARTICLE IN PRESS
1528
0.5

S. Fu et al. / Ocean Engineering 34 (2007) 15161531

0.4

=45 Mode 9 Mode 11 Mode 13

0.3

|P|/A
0.2 0.1 0.0 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0

Incident Wave Freq. (rad/s)

Fig. 7. The torsional mode (mode 9, 11 and 13) response amplitude in oblique sea.

a
1.8 Vertical Displacement (|w|/A) 1.6 1.4 1.2 1.0 0.8 0.6 0.4 0.2 0.0 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 Incident Wave Freq. (rad/s) 1.6 1.8 2.0 2.2
Vertical Disp. of P1 7 modes 8 modes 10 modes 14 modes 35 modes

b
Vertical Displacement (|w|/A)

1.6 1.4 1.2 1.0 0.8 0.6 0.4 0.2 0.0 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 Incident Wave Freq. (rad/s) 1.6 1.8 2.0 2.2
Vertical Disp. of P3 7 modes 8 modes 10 modes 14 modes 35 modes

Fig. 8. The effect of the number of modes used in calculationon vertical displacement, for connectors with zero rotational stiffness. (a) Vertical displacement amplitude at point P1 in Fig. 2, (b) Vertical displacement amplitude at point P3 in Fig. 2.

generalize coordinates response amplitude are zero and are not shown in Fig. 6. However, these modes will be excited in oblique sea as shown in Fig. 7, where the bending stiffness of the module and rotational stiffness of the connectors are set to be 4:77 1011 N m2 and zero, respectively. Fig. 8 shows the dependence of the hydroelastic response on the number of modes used in the analysis, where the bending stiffness of the module and rotational stiffness of the connectors are set to 4:77 1011 N m2 and zero,

respectively. In Fig. 8, N modes means that the rst N modes are selected to describe the deformations of the structure. As can be seen from Fig. 8, the number of modes selected to describe the structural deformations affects the vertical displacement amplitude. Whether or not the exible modes are considered will greatly affect the response, however, compared with the higher modes contributions, the rst few exible modes will contribute much to the nal response.

ARTICLE IN PRESS
S. Fu et al. / Ocean Engineering 34 (2007) 15161531
250
Cross Section at P2 Cross Section at P4

1529

Bending Moment Mx (MN.m)

200

150

100

50

0 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 Incident Wave Freq. (rad/s) 1.6 1.8 2.0

Fig. 9. Bending moment amplitudes in cross-sections P2 and P4 as shown in Fig. 2 versus incident wave frequency, for interconnected structures with zero rotational connector stiffness.

240 210 Bending Moment Mx (MN.m) 180 150 120 90 60 30 0 0.0 0.1 0.2 0.3 0.4 0.5 x/L 0.6 0.7 0.8

=0.3rad/s =0.65rad/s =1.4rad/s

0.9

1.0

Fig. 10. Bending moment amplitude distributions along the length of the interconnected structure, with zero rotational connector stiffness.

240 210 Bending Moment Mx (MN.m) 180 150 120 90 60 30 0 0.0


=0.65rad/s Krot=0.0 Krot=0.159E9 Krot=0.159E10 (Continous)

0.1

0.2

0.3

0.4

0.5 x/L

0.6

0.7

0.8

0.9

1.0

Fig. 11. The effects of rotational connector stiffness on the bending moment amplitude along the length of the structure.

Fig. 9 shows variations of the bending moment amplitude for two different cross sections with different incident wave frequencies. Bending moment amplitudes along the length of the structure are shown in Figs. 10

and 11, where the bending stiffness of the module is kept constant at 4:77 1011 N m2 . As can be seen in Fig. 9, the upstream and downstream modules have almost the same bending moment amplitude

ARTICLE IN PRESS
1530 S. Fu et al. / Ocean Engineering 34 (2007) 15161531 Chen, X.J., Jensen, J.J., Cui, W.C., Fu, S.X., 2003b. Hydroelasticity of a oating plate in multidirectional waves. Ocean Engineering 30 (15), 19972017. Cui, W.C., Yang, J.M., Wu, Y.S., Liu, Y.Z., 2007. Theory of Hydroelasticity and its Application to Very Large Floating Structures. Shanghai Jiao Tong University Press, to be published. Eatock Taylor, R., 2003. Wet or Dry Modes in Linear Hydroelaticity Why Modes? Hydroelasticity in Marine Technology, Oxford, UK, pp. 239250. Eatock Taylor, R., Ohkusu, M., 2000. Green functions for hydroelastic analysis of vibrating free-free beams and plates. Applied Ocean Research 22, 295314. Ertekin, R.C., Kim, J.W., 1999. Hydroelastic response of oating mattype structure in oblique, shallow-water waves. Journal of Ship Research 43 (4), 241254. Hamamoto, T., 1998. 3D hydroelastic analysis of module linked large oating structures using quadratic BE-FE hybrid model, hydroelasticity in marine technology. In: Proceedings of the Second International Conference on Hydroelasticity in Marine Technology, Fukuoka, Japan, pp. 3637. Hermans, A.J., 1998. A boundary element method to describe the excitation of waves in a very large oating exible platform. Hydroelasticity in Marine Technology, pp. 6976. Hibbit, Karlsson, Sorenson, 2004. ABAQUS Users Manual 6.5. ABAQUS Inc. Kashiwagi, M., 1998. A B-Spline Galerkin scheme for calculating the hydroelastic response of a very large oating structure in waves. Journal of Marine Science and Technology 1, 3749. Kim, D., Chen, L., Blaszkowski, Z., 1999. Linear frequency domain hydroelastic analysis for McDermotts mobile offshore base using WAMIT. In: Proceedings of the Third International Workshop on Very Large Floating Structures, VLFS99, Honolulu, pp. 105113. Kim, J.W., Ertekin, R.C., 1998. An eigenfunction-expansion method for predicting hydroelastic behavior of a shallow-draft VLFS. In: Proceedings Second International Conference on Hydroelasticity in Marine Technology (ICHMT98), Fukuoka, Japan, pp. 4759. Lee, C.H., Newman, J.N., 2000. An assessment of hydroelasticity for very large hinged vessels. Journal of Fluid and Structures 14, 957970. Lee, C.H., Newman, J.N., 2004. WAMIT Users Manual 6.2. WAMIT Inc, MIT, MA, USA. Newman J.N., 1997a. Wave effects on hinged bodies. Part Ibody motions, technical report /http://www.wamit.com/publications. htmS. Newman, J.N., 1997b. Wave effects on hinged bodies. Part IIhinge loads, technical report /http://www.wamit.com/publications.htmS. Newman, J.N., 1998a. Wave effects on hinged bodies. Part IIIhinge loads vs. number of modules, technical report /http://www.wamit. com/publications.htmS. Newman, J.N., 1998b. Wave effects on hinged bodies. Part IVvertical bending modes, technical report /http://www.wamit.com/publications. htmS. Newman, J.N., 2005. Efcient hydrodynamic analysis of very large oating structures. Marine Structures 16, 169180. Ohkusu, M., 1998. Hydroelastic behavior of oating articial islands in waves. Journal of the Society of Naval Architects of Japan 183, 239248. Riggs, H.R., Ertekin, R.C., 1993. Approximate methods for dynamic response of multi-module oating structures. Marine Structures 6, 117141. Riggs, H.R., Ertekin, R.C., Mills, T.R.J., 1999. Impact of stiffness on the response of a multimodule mobile offshore base. International Journal of Offshore and Polar Engineering 9 (2), 126133. Riggs, H.R., Ertekin, R.C., Mills, T.R.J., 2000. A comparative study of RMFC and FEA models for the wave-induced response of a MOB. Marine Structures 13, 217232. Sahoo, T., Yip, T.L., Chwang, A.T., 2000. On the interaction of surface wave with a semi-innite elastic plate. In: Proceedings of the 10th International Offshore and Polar Engineering Conference, Seattle, USA, pp. 584589.

for frequencies up to 0.45 rad/s. This means that rigid body motion dominates the interconnected structures, and these frequencies do not contribute to the bending moments of the structures. At higher frequencies the upstream module will always have a greater bending moment than the downstream module. At a frequency of about 0.6 rad/s the bending moments of the upstream and downstream modules reach their maximum values, which can be observed by comparing Figs. 5, 6 and 9. Near this frequency, bending modes number 8, 10, 12 and 14, which contribute to the bending moment of the module, are all excited by the incident wave. As can be seen in Fig. 10, for low frequencies of the incident wave the upstream and downstream modules will have almost the same bending moment amplitude. This is due to the fact that the rigid body modes and diffraction wave forces dominate the motion and deection of the interconnected structure. As the incident wave frequency increases, the distribution of bending moments in the two modules becomes more closely related to the exible modes of each module. This phenomenon can also be observed by comparing Figs. 5, 6 and 10. Fig. 11 shows that the rotational connector stiffness also affects the bending moment amplitude along the structure. As the rotational stiffness decreases from a very high value (representing a continuous structure) to zero, the bending moment in the middle of the structure decreases from its maximum value to zero. The bending moment within the upstream and downstream modules, on the other hand, will increase slightly. 5. Conclusions This paper presents a hydroelastic theory of oating structures composed of exible, interconnected modules. A numerical example of a two-module structure is presented. The effects of connector and module stiffness on the hydroelastic response of the structure are studied, as well as the characteristics of the bending moment distribution. The results show that the stiffness of the connectors and the modules is very important in determining the hydroelastic response of the structure. References
Bathe, K.J., Wilson, E.L., 1976. Numerical Methods in Finite Element Analysis. Prentice-Hall, Englewood Cliffs, NJ. Bishop, R.E.D., Price, W.G., 1979. Hydroelasticity of Ships. Cambridge University Press, UK. Che, X.L., Wang, D.Y., Wang, M.L., Xu, Y.F., 1992. Two-dimensional hydroelastic analysis of very large oating structures. Marine Technology 29 (1), 1324. Che, X.L., Riggs, H.R., Ertekin, R.C., 1994. Composite 2D/3D hydroelastic-analysis method for oating structures. Journal of Engineering Mechanics 120 (7), 14991520. Chen, X.J., Wu, Y.S., Cui, W.C., Tang, X.F., 2003a. Nonlinear hydroelastic analysis of a moored oating body. Ocean Engineering 30 (8), 9651003.

ARTICLE IN PRESS
S. Fu et al. / Ocean Engineering 34 (2007) 15161531 Seto, H., Ochi, M. 1998. A hybrid element approach to hydroelastic behavior of a very large oating structure in regular wave. Hydroelasticity in Marine Technology. In: Proceedings of the second international conference on hydroelasticity in marine technology, Fukuoka, Japan, pp.185193. Sim, I.H., Choi, H.S., 1998. An analysis of the hydroelastic behavior of large oating structures in oblique waves. Hydroelasticity in marine technology. In: Proceedings of the Second International Conference on Hydroelasticity in Marine Technology, Fukuoka, Japan, pp. 165169. Sun, H., Song, H., Cui, W.C., 2002. On the interaction of surface waves with an elastic plate of nite length in head seas. China Ocean Engineering 16 (1), 2132. Taghipour, R., Fu, S.X., Moan, T., 2006. Validated two and three dimensional linear hydroelastic analysis using standard software. In: Proceedings of the 16th International Offshore and Polar Engineering Conference, San Francisco, California, USA, pp. 101107. 1531 Wang, D.Y., Riggs, H.R., Ertekin, R.C., 1991. Three-dimensional hydroelastic response of a very large oating structure. International Journal of Offshore and Polar Engineering 1 (4), 307316. Wu, C., Watanabe, E., Utsunomiya, T., 1995. An eigenfunction expansion matching method for analysis wave-induced response of a large oating plate. Applied Ocean Research 17, 301310. Wu, M.K., Moan, T., 1996. Linear and nonlinear hydroelastic analysis of high-speed vessels. Journal of Ship Research 40 (2), 149163. Wu, Y.S., 1984. Hydroelasticity of oating bodies, Ph.D. Thesis, Brunel University, UK. Wu, Y.S., Wang, D.Y., Riggs, H.R., Ertekin, R.C., 1993. Composite singularity distribution method with application to hydroelasticity. Marine Structures 6, 143163. Xia, J.Z., Wang, Z.H., Jensen, J.J., 1998. Non-linear wave loads and ship responses by a time-domain strip theory. Marine Structure 11 (3), 101123. Yago, K., Endo, H., 1996. On the hydroelastic response of boxshaped oating structure with shallow draft. Journal of the Society of Naval Architects of Japan 180, 341352.

Anda mungkin juga menyukai