1-s2.0-s0029801808000115-main.pdf

Upload: chiragpalan1987

Post on 04-Jun-2018

217 views

Category:

Documents


0 download

TRANSCRIPT

  • 8/13/2019 1-s2.0-S0029801808000115-main.pdf

    1/9

    Ocean Engineering 35 (2008) 589597

    Comparison of hydroelastic computer codes based

    on the ISSC VLFS benchmark

    H.R. Riggsa,, Hideyuki Suzukib, R.C. Ertekinc, Jang Whan Kimd, K. Iijimae

    aDepartment of Civil and Environmental Engineering, University of Hawaii at Manoa, Honolulu, HI 96822, USAbDepartment of Environmental and Ocean Engineering, University of Tokyo, Tokyo 108-8639, Japan

    cDepartment of Ocean and Resources Engineering, University of Hawaii at Manoa, Honolulu, HI 96822, USAdTechnip USA, Houston, TX 77079, USAeOsaka University, Osaka 565-0871, Japan

    Received 3 August 2007; accepted 10 January 2008

    Available online 25 January 2008

    Abstract

    There has been substantial development in computer codes for linear hydroelasticity in recent years, driven in part by the motivation to

    investigate the wave-induced response of very large floating structures (VLFSs). A recent International Ship and Offshore Structures

    Congress (ISSC) state-of-the-art report on VLFS design and analysis [ISSC, 2006. Report of Specialist Task Committee VI.2, very large

    floating structures. In: Frieze, P.A., Shenoi, R.A. (eds.), Proceedings of the 16th International Ship and Offshore Structures Congress,

    Elsevier, Southampton, UK, pp. 397451] included a brief comparative study of the simulation results from different computer codes

    for a pontoon (mat-like) VLFS. The codes covered a mix of both fluid models (potential and linear GreenNaghdi) and structural models

    (3-D grillage, 2-D plate, 3-D shell). A more detailed comparison of the results from a select group of models from that study is provided

    and discussed herein. The similarities in the results increase the confidence level of the state-of-the-art in predicting the hydroelastic

    response of such structures, and the differences, including in computational efficiency, lead to an understanding of the significance of

    specific modeling assumptions and their impact on the predicted response.r 2008 Elsevier Ltd. All rights reserved.

    Keywords: Very large floating structure; VLFS; Hydroelasticity; ISSC; Potential theory; GreenNaghdi theory

    1. Introduction

    There has been substantial development in hydroelastic

    computer codes over the last two decades. Much of the

    development of 3-D formulations has followed the foun-

    dational work of Wu (Price and Wu, 1985;Wu, 1984). The

    driving force, especially in the 1990s, for the developmenthas been the Japanese Megafloat project (Suzuki, 2005)

    and the US Mobile Offshore Base project (Palo, 2005). The

    former project was concerned principally with floating

    civilian airports in protected, shallow water, and focused

    on pontoon-type very large floating structure (VLFS). The

    latter project was concerned principally with a floating

    military airbase in the deep open ocean, and hence focused

    on multi-module semi-submersible-type VLFS. This latter

    project included a comparison of the predictive response of

    different computer codes for the latter application (BNI,

    1999; NFESC, 2000; Wung et al., 1999). However, the

    computer codes were based on very similar formulations,

    for the most part, and the system consisted of flexibly

    connected rigid modules. The results are not directlytransferable to the pontoon-type VLFS and the more

    different structural and fluid models that are used for this

    class of structure. It should be noted that one of the

    computer codes (HYDRAN) used in that comparative

    study is also used here.

    The interest in VLFS prompted the International Ship

    and Offshore Structures Congress (ISSC) to establish a

    special task committee on VLFS, whose task was to

    consider the state-of-the-art in VLFS analysis and design

    procedures, including the application of hydroelasticity

    ARTICLE IN PRESS

    www.elsevier.com/locate/oceaneng

    0029-8018/$ - see front matterr 2008 Elsevier Ltd. All rights reserved.

    doi:10.1016/j.oceaneng.2008.01.012

    Corresponding author. Tel.: +1 808956 6566; fax: +1 808956 5014.

    E-mail address: [email protected] (H.R. Riggs).

    http://www.elsevier.com/locate/oceanenghttp://localhost/var/www/apps/conversion/tmp/scratch_5/dx.doi.org/10.1016/j.oceaneng.2008.01.012mailto:[email protected]:[email protected]://localhost/var/www/apps/conversion/tmp/scratch_5/dx.doi.org/10.1016/j.oceaneng.2008.01.012http://www.elsevier.com/locate/oceaneng
  • 8/13/2019 1-s2.0-S0029801808000115-main.pdf

    2/9

    (ISSC, 2006). One section of that report included a

    comparative study on the wave-induced response deter-

    mined by several computer programs, each based on

    significantly different models. Space limitations prevented

    a detailed comparative study of the methodology and the

    results (ISSC, 2006). Therefore, a more detailed presenta-

    tion is made herein, including additional and some up-dated results). Such a comparison is important because:

    (1) similarities in the results increase the confidence level of

    the state-of-the-art in predicting the hydroelastic response

    of such structures, and (2) differences lead to an under-

    standing of the significance of specific modeling assump-

    tions and their impact on the predicted response. A similar

    comparative study (Jiao et al., 2006) has been published

    recently, but it was restricted to a shallow draft plate. In

    addition, it considered only displacements. For VLFS

    design, prediction of stresses is very important, which the

    current study considers.

    2. Problem description

    The model is a rectangular, pontoon structure 500 m

    long and 100 m wide. It is 2 m high, the draft is 1 m, and it

    is in 20m of seawater. Transverse and longitudinal

    bulkheads are spaced at 50 and 20 m, respectively. The

    top deck, bottom hull, bulkheads, and sides are steel plates

    with a nominal thickness of 20mm. To include the

    additional bending stiffness associated with stiffeners that

    are not modeled, the plate bending thickness is 150 mm for

    all plates. The mass density of steel is 7850 kg/m3, and the

    structural mass is based on the nominal steel volume given

    the 20 mm plate thickness. The modulus of elasticity, E,and Poissons ratio, n, for steel were taken to be 200 GPa

    and 0.3, respectively. To model the non-structural mass,

    the top and bottom plates are assigned an additional mass

    density of 17,131 kg/m3. As a result, the CG is located at

    the geometric center of the cross-section. Note that all

    dimensions are midplane dimensions consistent with a shell

    finite element model of the structure. The density of

    seawater, r, is 1025 kg/m3, and gravitational acceleration,

    g, is 9.81 m/s2. Structural damping is not included in the

    dynamic analysis. The seawater is assumed unbounded

    horizontally and the seabed is flat.

    The model has been designed to have significant flexible

    response under waves. Of interest herein is the global

    response, and in particular displacements and stresses.

    Results are reported for wave periods between 5 and 30 s,

    and for wave angles of 01 (head seas), 301, and 451.

    3. Computer codes

    Three computer programs were used to obtain the

    dynamic response: HYDRAN (OCI, 2005), VODAC

    (Iijima et al., 1997) and LGN (Kim and Ertekin, 1998).

    In all programs, the fluid is assumed to be incompressible

    and inviscid and the flow is assumed to be irrotational.

    In the first two programs, the fluid model is based on

    linear, 3-D potential theory. In the last program, the

    fluid model is based on the linear GreenNaghdi equations

    for long waves. HYDRAN uses a traditional constant

    panel Green function formulation for the fluid and a

    3-D shell finite element model for the structure. VODAC

    uses a traditional constant panel Green function formula-

    tion, together with interaction theory, for the fluid and a 3-D grillage model for the structure. LGN uses the Green

    Naghdi equations in the fluid domain, as mentioned, and a

    linear Kirchhoff plate model; the governing equations are

    matched at the juncture boundaries and they are solved by

    a combination of eigenfunction expansion method and the

    boundary-integral equation method. All of the codes

    determine the response in the frequency domain.

    4. Fluid modelsHYDRAN, VODAC

    The fluid models in these programs are based on the

    linear potential theory. The fluid motion is assumed small,and hence the total velocity potential F for a regular wave

    at frequency o can be written as a function of position

    r [x1,x2,x3]T and structural displacements u as

    F fI fD io/Rueiot (1)

    in which fI, fD, and /Rand the incident, diffraction, and

    radiation potentials, respectively. The diffraction and

    radiation potentials must satisfy the same governing

    equation in the fluid domain, O, and the same boundary

    conditions on the free surface and at the seabed

    r2f 0; r2 O, (2)

    o2f gqf

    qx3 0 on x3 0, (3)

    limx3!1

    qf

    qx30. (4)

    They must also satisfy Sommerfelds radiation condition

    on the still water surface at infinitely large radial distances

    from the body. On the wetted body surface, S, the

    diffraction potential must satisfy

    qfDq

    n

    qfIq

    n

    on S, (5)

    where n is the outward-directed unit normal vector to the

    wetted surface, while the radiation potentials must satisfy

    qfRqn

    nn on S. (6)

    n* is the generalized normal, i.e., the normal displacement

    of the body. Once the velocity potential is determined, the

    relationship between velocity and pressure, p, from Eulers

    integral is used to determine the linear, dynamic pressure

    on the structure

    p rqF

    qt . (7)

    ARTICLE IN PRESS

    H.R. Riggs et al. / Ocean Engineering 35 (2008) 589597590

  • 8/13/2019 1-s2.0-S0029801808000115-main.pdf

    3/9

    5. Fluid modelLGN

    In the LGN model, a simple kinematics of fluid motion

    is used under the assumption that the wave length of the

    free-surface motion is much longer than water depth, h.

    The horizontal and vertical velocity profiles along the

    depth are assumed to be constant and to vary linearly,respectively. In the case of small amplitude wave motion,

    the horizontal velocity field ~Vx;y; t and the verticalvelocity component w(x,y,z,t) are given in terms of the

    complex mean velocity potential c(x,y) (see Ertekin and

    Kim, 1999)

    Vx;y

  • 8/13/2019 1-s2.0-S0029801808000115-main.pdf

    4/9

    7. Structural modelVODAC

    VODAC was originally developed for the analysis of

    semi-submersible type VLFS supported by multiple floa-

    ters. Although the structural modeling available in the code

    is 3-D beamcolumn grillage, the structural model for this

    study is a simpler plane grillage, as shown in Fig. 3.

    Standard, 2-node 3-D beamcolumn elements are used, six

    degrees of freedom per node (three translations and three

    rotations). Member forces are axial force, vertical and

    horizontal shears, torsion, and vertical and horizontal

    bending moments. VODAC uses a substructure method to

    enhance computational efficiency. This approach is effec-

    tive especially for VLFS, which often have a repetitive

    structure.For the structural model used herein, the nodal spacing

    is 12.5 m in the longitudinal direction (40 elements per

    section line) and 10m in the transverse direction (10

    elements per section line), resulting in a total of 400

    elements. The elements are located so as to determine the

    deflections along the longitudinal and transverse bulk-

    heads. The axial, shear, torsional, and bending rigidities

    are determined so as to be equivalent to the structural

    properties of the pontoon-type VLFS, including the

    Poisson effect. For example, the bending rigidity EI for

    the transverse cross-section is approximately 880 GN-m2

    according to the following formula:

    EI E

    1 n21

    2BH2t

    , (12)

    where B is breadth of the structure, H is height of the

    structure, and t is the actual plate bending thickness; this

    considers contributions from the top and bottom plates

    only. The bending rigidity is equally distributed to the

    beam elements that represent the cross-section; i.e., all

    beams in a given direction have the same properties.

    A similar approach is used for the beams that model the

    longitudinal cross-section. The following equivalent sec-

    tion properties are used: for longitudinal elements,

    A 0.8m2, Iyy 0.88 m4, Izz 26.7 m4, J 1.6 m4; and

    for transverse elements, A 1.0m2, Ixx 1.1 m4, Izz

    52.1 m4, J 2.0 m4. The moments of inertia include the

    term 1/(1n)2 to account for the Poisson effect.

    8. Structural modelLGN

    The rectangular pontoon is modeled as a homogeneous,

    isotropic plate with uniform mass distribution (per unit

    area), m, and flexural rigidity, D Et3eq=121 n2, where

    teq 0.792 m is the effective plate thickness. This thickness

    is defined so as to result in the same moment of inertia

    as the actual built-up, yz section of the pontoon. The

    moment of inertia of thexzsection is very similar, and so

    the isotropic assumption is reasonable. The verticaldisplacement, z(x,y,t) of the plate is assumed to be

    governed by the thin-plate theory (see e.g., Rayleigh, 1894):

    mq

    2z

    qt2 DD2z pf (13)

    in whichpfis the spatial part of the time-harmonic pressure

    on the bottom of the plate and D q2/qx2+q2/qy2 is the

    two-dimensional Laplacian on the horizontal plane. Since

    the plate is freely floating, the bending moment and shear

    force should vanish at the edges of the plate

    q2zqn2

    n q2z

    qs20; q

    3zqs3

    2 n q3z

    qs2qn0 on J, (14)

    where nand s denote the normal and tangential directions

    on the boundary J, andn is Poissons ratio. Eq. (14) can be

    written alternatively as

    Dz 1 nq

    2z

    qs20;

    q

    qn Dz 1 n

    q2z

    qs2

    0 on J.

    (15)

    At the corners of the plate, there can be concentrated

    shear forces to compensate for the torsional moment

    along the edges of the plate. The vanishing of this shear

    ARTICLE IN PRESS

    Fig. 3. VODAC structural model.

    H.R. Riggs et al. / Ocean Engineering 35 (2008) 589597592

  • 8/13/2019 1-s2.0-S0029801808000115-main.pdf

    5/9

    force leads to

    q2z

    qxqy 0. (16)

    Combining Eqs. (9), (10), and (13), the following coupled

    equation can be obtained in Region II:

    DD3cII rgDcII m rh

    3

    o2DcII

    ro2

    h cII 0.

    (17)

    In Region I, the atmospheric pressure is neglected, i.e.,

    pf=0, and therefore, we have

    DcI k2cI 0. (18)

    Along the juncture boundary, J, the continuity of mass

    flux and depth-mean pressure leads to the following match-

    ing conditions:

    qcI

    qn

    qcII

    qn ; cI cII on J. (19)

    9. Solution procedureHYDRAN

    HYDRAN uses the 3-D source distribution, or the

    Green function, method to determine the velocity poten-

    tials. The Green function satisfies all but the body

    boundary conditions (Wehausen and Laitone, 1960), and

    so only the wetted surface must be discretized. It uses the

    constant-strength panel formulation. This well-known

    method has been used for rigid body hydrodynamics by

    many authors (Faltinsen and Michelsen, 1974; Garrison,1974, 1975) and more recently for hydroelasticity (Price

    and Wu, 1985;Wu, 1984).

    To reduce the number of radiation potentials that must

    be solved, a reduced-basis approach is used. In particular,

    the structural displacements are represented by a combina-

    tion of a limited number of the dry (i.e., in-air) normal

    modes of vibration (Price and Wu, 1985; Wu, 1984).

    Specifically,u Wp, where W is the matrix of normal mode

    shapes and p is the vector of generalized displacements.

    This modal-superposition-type strategy reduces the radia-

    tion potentials from the order of the number of structural

    degrees of freedom to a relatively small number. It alsomeans that the radiation potentials correspond to global

    displacement shapes rather than local finite element shape

    functions. The equation of motion becomes

    o2 Mns Mn

    f

    i Cns oC

    n

    f

    KnS K

    n

    f

    h ip FnW

    (20)

    in which

    Mns WTMsW; K

    n

    S WTKSW; K

    n

    f WTKfW,

    whereMs,KS, andKfare the structural mass, stiffness, and

    hydrostatic stiffness matrices, respectively, and Cns is the

    diagonal matrix of modal hysteretic damping coefficients

    (not used herein). Mns and Kn

    S are diagonal matrices

    because the dry normal modes are used. The terms in the

    vector of wave exciting forces are obtained from

    FnWj iro

    Z fI fDn

    n

    j dS, (21)

    while the terms in Mn

    f and Kn

    fare given respectively by

    mijr

  • 8/13/2019 1-s2.0-S0029801808000115-main.pdf

    6/9

    condition required for Kagemotos interaction theory.

    However, it has been shown from experiments and

    simulations that VODAC is applicable to pontoon-type

    VLFS with good accuracy (Iijima et al., 1999). The present

    problem is formulated as a hydrodynamic interaction

    problem between 100 (20 in longitudinal direction 5 in

    transverse direction) floating bodies. Each body has 220panels, based on a mesh discretization of 10 10 3 in the

    x, y, and z directions, respectively. Diffraction character-

    istics (Goo and Yoshida, 1990), which represent the re-

    lations between incident wave and scattered wave from a

    unit body, are calculated in advance by using the normal

    singularity distribution method. Once the diffraction

    characteristics are obtained, we have only to calculate the

    diffraction coefficients representing the hydrodynamic

    interactions by solving a system of equations that describe

    the relations between the diffraction coefficients. The

    source distribution is hierarchically calculated using the

    diffraction coefficients and the diffraction characteristics.

    As mentioned, the structural modeling is also hierarch-

    ical. The substructure method is applied to the dynamic

    analysis case, and the whole structure is expressed as a

    repetition or an assembly of the substructures. The stiffness

    matrix of a substructure is condensed into a smaller matrix

    with all the degrees of freedom eliminated except those of

    the boundary nodes. In this process, dynamic condensation

    is performed instead of static condensation. The final

    matrix to be solved is obtained by summing the condensed

    matrices regarding the boundary nodes. The displacements

    of the boundary nodes are solved first, together with the

    diffraction coefficients. Then the displacements of the inner

    nodes are calculated through backward calculations usingthe relations obtained in the condensation stage.

    Additional details of the solution procedure can be

    found inIijima et al. (1999).

    The CPU time for a single frequency and wave heading

    was 18.86 s on a Pentium 4 PC, 2.16 GHz with 2 GB RAM,

    running Windows XP.

    11. Solution procedureLGN

    The numerical solution of Eq. (18) defined in Region I,

    where there is no floating plate, can be given by

    distribution of the Green function along the boundary J

    cIIx;y cwx;y p

    2i

    ZJ

    H10 kjx isjsis ds, (24)

    wherecw(x,y) is the velocity potential for the incident wave,

    which is known. H10 is the Hankel function and s(s) is the

    unknown source strength defined along the juncture

    boundary, J. To obtain the numerical solution, the jun-

    cture boundary is discretized into a finite number of line

    elements, where the source strengths are assumed to be

    piecewise constant. The unknown source strengths are

    determined by matching the solution with the solution in

    Region II through the matching conditions given by Eq. (19).

    For this problem, 480 line elements (200 and 40 along each

    edge of the plate in x and y directions, respectively) were

    used. No other discretization is necessary because the model

    uses semi-analytical models for both the fluid and the plate.

    In Region II, the governing equation, Eq. (17), can be

    solved by the eigenfunction-expansion method because the

    domain is a simple rectangular one. The coefficients of the

    eigenfunctions are obtained by enforcing the matchingcondition, Eq. (19), and the free-edge conditions given by

    Eqs. (15) and (16). The details of this approach are given in

    Kim and Ertekin (1998).

    The CPU time for each frequency and each wave

    heading was 1.85s on a Pentium 4 PC, 2.0 GHz, with

    384 MB RAM, running Windows 2000.

    12. Results

    12.1. Natural periods and modes

    As mentioned previously, the HYDRAN solution used

    the first 30 dry normal modes for the reduced basis. The

    first dry deformation mode of the shell model corre-

    sponded to vertical bending in the longitudinal direction

    and had a natural period of 24.6s, while the second

    bending mode had a period of 8.96 s and the third bending

    mode had a period of 4.62 s. The first bending mode in the

    transverse direction had a natural period of 1.32 s. The

    30th mode had a natural period of 0.81 s and a spatial

    variation unlikely to attract significant forces for the wave

    periods considered.

    Although not needed for the analysis, the program also

    estimates the wet natural periods and modes, which include

    the added mass and the hydrostatic stiffness. The programfound 6 wet natural periods between 8 and 30 s: 8.2, 8.4,

    8.6, 10.5, 14.50, and 15 s.

    ARTICLE IN PRESS

    0

    0.2

    0.4

    0.6

    0.8

    1

    -250

    HYDRAN

    VODAC

    LGN

    Maximumverticaldisplacem

    entin10secwave(m/m)

    Position (m)

    Wave angle = 0

    -200 -150 -100 -50 50 100 150 200 2500

    Fig. 4. Maximum vertical displacement in 10 s head sea.

    H.R. Riggs et al. / Ocean Engineering 35 (2008) 589597594

  • 8/13/2019 1-s2.0-S0029801808000115-main.pdf

    7/9

    The other two programs do not require natural periods

    and mode shapes, so no comparison can be made. The

    results are reported principally to reveal a little bit more the

    fundamental characteristics of the structure.

    12.2. Wave-induced response

    The maximum displacements along the centerline

    induced by a wave with a period of 10 s and an incidence

    angle of 01(propagating in the xdirection) are shown in

    Fig. 4. The shape is strongly dependent on the wave angle,

    as can be seen by comparing Figs. 4 and 5. These results

    show generally good comparisons between the programs,

    especially considering the different numerical models that

    were used to obtain the results. However, in the oblique sea

    case,Fig. 5, the results for LGN differ significantly in the

    interior of the plate. The reason for this is unknown, but itmay be that the twisting response induced by the oblique

    ARTICLE IN PRESS

    0

    10

    20

    30

    40

    50

    60

    70

    80

    90

    5

    HYDRAN

    VODAC

    LGN

    Longitudinalstressatcenter(MPa/m)

    Wave Period (s)

    Wave angle = 0

    Wave angle = 45

    Wave angle = 30

    10 15 20 25 30

    Fig. 8. Longitudinal stress at center.

    0

    0.2

    0.4

    0.6

    0.8

    1

    1.2

    1.4

    5

    HYDRAN

    VODAC

    LGN

    Verticaldisplacementatfore(m/m)

    Wave Period (s)

    Wave angle = 0

    Wave angle = 45

    Wave angle = 30

    10 15 20 25 30

    Fig. 7. RAO of vertical displacement at bow.

    0

    0.2

    0.4

    0.6

    0.8

    1

    1.2

    1.4

    5

    HYDRAN

    VODAC

    LGN

    Verticaldisplaceme

    ntatcenter(m/m)

    Wave Period (s)

    Wave angle = 0

    Wave angle = 45

    Wave angle = 30

    10 15 20 25 30

    Fig. 6. RAO of vertical displacement at center.

    0

    0.2

    0.4

    0.6

    0.8

    1

    1.2

    -250

    HYDRAN

    VODAC

    LGN

    Maximumverticaldisplacementin10secwave(m/m)

    Position (m)

    Wave angle = 45

    -200 -150 -100 -50 0 50 100 150 200 250

    Fig. 5. Maximum vertical displacements in 10 s oblique sea.

    H.R. Riggs et al. / Ocean Engineering 35 (2008) 589597 595

  • 8/13/2019 1-s2.0-S0029801808000115-main.pdf

    8/9

    sea is not captured as well by the plate model. All three

    models agree quite well in the end displacements, however.

    The programs agree well for the RAOs for the vertical

    displacement at the center and bow of the structure (again

    along the centerline) for the 3 wave angles of 01, 301, and

    451; seeFigs. 6 and 7.

    RAOs of the longitudinal stresses at the center and 50 m

    back from the bow, also along the centerline, obtainedfrom the programs show strikingly good agreement; see

    Figs. 8 and 9. The difference in peak stresses is less than

    10% for 01 and 301, while it is about 15% for 451.

    Considering the strikingly different structural models, the

    agreement is surprising, especially in oblique seas.

    13. Conclusions

    The comparative study considered three programs for

    the linear hydroelastic response of a pontoon-type VLFS.

    Two of the programs were based on potential theory, while

    one used linear GreenNaghdi theory for long waves. All

    three programs used very different structural models. The

    results for this pontoon-type VLFS, considering a wide

    range of wave periods and wave angles, showed good

    agreement between the programs. Even the stresses agreed

    quite well. This demonstrates that all three models can be

    used with confidence, especially for preliminary studies.

    The computer times show that LGN is the most com-

    putationally efficient, although there is little practical

    difference between time requirements for it and VODAC.

    HYDRAN is the most computationally demanding. LGN,

    however, is limited to a structure that can be modeled

    as a uniform plate and for relatively shallow water. The

    LGN theory is established for shallow-water waves, and

    although the agreement of the LGN results to the linear

    potential theory results for smaller wave periods is

    remarkable, it is likely that for deeper water the results

    will show a larger difference. (For more discussion of this

    issue, the reader is referred toKim and Ertekin (2001), who

    compared the LGN predictions with the linear potential

    theory predictions for the dispersion relation for waterwaves and for hydroelastic waves on a mat-like structure.)

    The present results also confirm the conclusion that

    VODAC, which uses interaction theory of separate bodies,

    can be applied to a pontoon VLFS that consists of a single

    body (Iijima et al., 1999).

    References

    BNI, 1999. Mobile Offshore Base (MOB) Design Tools and Procedures:

    Benchmark Analysis Report. Bechtel National, Inc., San Francisco.

    Ertekin, R.C., Kim, J.W., 1999. Hydroelastic response of a floating, mat-

    type structure in oblique, shallow-water waves. Journal of ShipResearch 43 (4), 241254.

    Faltinsen, O.M., Michelsen, F.C., 1974. Motions of large structures in

    waves at zero Froude number. In: Proceedings of the International

    Symposium on the Dynamics of Marine Vehicles and Structures in

    Waves, University College, London, pp. 91106.

    Garrison, C.J., 1974. Hydrodynamics of large objects in the sea. Part

    IHydrodynamic analysis. Journal of Hydronautics 8, 512.

    Garrison, C.J., 1975. Hydrodynamics of large objects in the sea.

    Part IIMotion of free-floating bodies. Journal of Hydronautics 9

    (2), 5863.

    Goo, J.-S., Yoshida, K., 1990. A numerical method for huge semisub-

    mersible responses in waves. SNAME Transactions 98, 365384.

    Huang, L.L., Riggs, H.R., 2000. The hydrostatic stiffness of flexible

    floating structures for linear hydroelasticity. Marine Structures 13 (2),

    91106.Iijima, K., Yoshida, K., Suzuki, H., 1997. Hydrodynamic and hydroelastic

    analyses of very large floating structures in waves. In: Proceedings of

    the 16th Offshore Mechanics and Arctic Engineering Conference,

    Yokohama, pp. 139145.

    Iijima, K., Suzuki, H., Yoshida, K., 1999. Structural design methodology

    of VLFS from the viewpoint of dynamic response characteristics. In:

    Proceedings of the Third International Workshop on Very Large

    Floating Structures, Honolulu, pp. 249258.

    ISSC, 2006. Report of Specialist Task Committee VI.2, very large floating

    structures. In: Frieze, P.A., Shenoi, R.A. (eds.), Proceedings of the

    16th International Ship and Offshore Structures Congress, Elsevier,

    Southampton, UK, pp. 397451.

    Jiao, L.-L., Fu, S.-X., Cui, W.-C., 2006. Comparison of approaches for

    the hydroelastic response analysis of very large floating structures.

    Journal of Ship Mechanics 10 (3), 7191.Kagemoto, H., Yue, D.K.P., 1986. Interactions among multiple three-

    dimensional bodies in water waves: an exact algebraic method. Journal

    of Fluid Mechanics 166, 189209.

    Kim, J.W., Ertekin, R.C., 1998. An eigenfunction-expansion method for

    predicting hydroelastic behavior of a shallow-draft VLFS. In:

    Proceedings of the Hydroelasticity in Marine Technology, Kyushu

    University, Fukuoka, Japan, pp. 4759.

    Kim, J.W., Ertekin, R.C., 2001. Hydroelasticity of an infinitely long plate

    in oblique waves: linear GreenNaghdi theory. Journal of Engineering

    for the Maritime Environment 216, 179197.

    Newman, J.N., 1994. Wave effects on deformable bodies. Applied Ocean

    Research 16, 4759.

    NFESC, 2000. Mobile offshore base (MOB) science and technology

    program final report. TR-2125-OCN, Naval Facilities Engineering

    Service Center, Port Hueneme, CA.

    ARTICLE IN PRESS

    0

    10

    20

    30

    40

    50

    60

    70

    5

    HYDRAN

    VODAC

    LGN

    Longitudinalstressfore(MPa/m)

    Wave Period (s)

    Wave angle = 45

    Wave angle = 0

    Wave angle = 30

    10 15 20 25 30

    Fig. 9. Longitudinal stress 50m back from bow.

    H.R. Riggs et al. / Ocean Engineering 35 (2008) 589597596

  • 8/13/2019 1-s2.0-S0029801808000115-main.pdf

    9/9

    OCI, 2005. HYDRAN: A Computer Program for the HYDroelastic

    Response ANalysis of Ocean Structures, vol. 1.52. OffCoast, Inc.,

    Kailua, HI.

    Palo, P.A., 2005. Mobile offshore base: hydrodynamic advancements and

    remaining challenges. Marine Structures 18 (2), 133147.

    Price, W.G., Wu, Y., 1985. Hydroelasticity of marine structures.

    In: Niordson, F.I., Olhoff, N. (Eds.), Theoretical and Applied

    Mechanics. Elsevier Science Publishers B.V., Amsterdam,pp. 311337.

    Rayleigh, J.W.S., 1894. The Theory of Sound. Dover Publications (1945),

    New York.

    Suzuki, H., 2005. Overview of megafloat: concept, design criteria, analysis,

    and design. Marine Structures 18 (2), 111132.

    Tessler, A., 1990. A C0-anisoparametric three-node shallow shell element.

    Computer Methods in Applied Mechanics and Engineering 78, 89103.

    Wehausen, J.V., Laitone, E.V., 1960. Surface waves. In: Laitone, E.V.

    (Ed.), Handbuch der Physik. Springer, Berlin, pp. 446776.

    Wu, Y., 1984. Hydroelasticity of floating bodies. Ph.D. Dissertation,

    Brunel University.

    Wu, Y.S., Wang, D.Y., Riggs, H.R., Ertekin, R.C., 1993. Composite

    singularity distribution method with application to hydroelasticity.Marine Structures 6 (2&3), 143163.

    Wung, C. C., Manetas, M., Ying, J., 1999. Mobile offshore base (MOB)

    design and analysis requirements and hydrodynamic tools evaluations

    and modeling guidelines. In: Proceedings of the Third International

    Workshop on Very Large Floating Structures, Honolulu, HI, pp. 5159.

    ARTICLE IN PRESS

    H.R. Riggs et al. / Ocean Engineering 35 (2008) 589597 597