You are on page 1of 13

Applied OceanResearch16 (1994) 47-59

~) 1994 Elsevier Science Limited Printed in Great Britain. All rights reserved 0141-1187/94/$07.00

ELSEVIER

Wave effects on deformable bodies


J. N. Newman
Department of Ocean Engineering, MIT, Cambridge, Massachusetts 02139, USA

The linearized frequency-domain analysis of wave radiation and diffraction by a three-dimensional body in a fixed mean position is extended to a variety of deformable body motions. These include continuous structural deflections, and also discontinuous motions which can be used to represent multiple interacting bodies. A general methodology is adopted with the body deflectiondefined by an expansion in arbitrary modal shape functions, and the response in each mode is obtained as a logical extension of the usual analysis for rigid-body modes. Illustrative computations are presented for the bending of a freely-floatingbarge, and of a vertical column with cantilever support at the bottom. For these structural deflections the use of orthogonal polynomials is emphasized, as an alternative to the more conventional use of natural modes. Also presented are computations for the motions of two rigid barges connected by a hinge joint, and for a finite array of images used to approximate wall effects on a cylinder in a channel. Results from the latter problem are compared with more analytical solutions, and it is shown that practical results can be obtained for the first-order hydrodynamic force coefficientsusing only a few images.

1 INTRODUCTION Wave-induced motions of ships and offshore platforms are usually described by six rigid-body modes including translations (surge, sway, heave) parallel to the Cartesian axes, and rotations (roll, pitch, yaw) about these axes. Additional 'generalized' modes of motion are significant in various applications including continuous structural deflections, discontinuous deflections of multi-component bodies with mechanical joints, and the interactions of separate bodies. These applications are quite different in a practical context, but similar mathematically. Thus it is convenient to analyze them from a single generalized approach. Wave-induced structural loads are important for many practical applications, but in most cases the structural and hydrodynamic analyses are performed separately. This is appropriate for stiff structures, where the eigenfrequencies of elastic deflections are substantially higher than the frequencies of the first-order wave loads. In these circumstances the analysis of wave radiation and diffraction can be performed neglecting the structural modes, and the hydrodynamic analysis of these modes can be restricted to the evaluation of the corresponding added-mass coefficients using a simplified high-frequency approximation of the free-surface effect. On the other hand, in applications where the eigenfrequencies of elastic deflections fall within the spectrum of first-order wave loads, the two problems must be
47

coupled to account for wave radiation in the analysis of the structural modes. In structural analysis it is common to define a special set of natural mode shapes, which correspond to the actual elastic deflections of the body in a specified physical context. This is difficult for hydroelastic problems, where the mode shapes are affected by the hydrodynamic pressure field and cannot be specified in advance. This difficulty can be avoided if the structural deflection is represented instead by a superposition of simpler mathematical mode shapes which are sufficiently general and complete to represent the physical motion. For example, the structural deflections of a floating slender ship can be expressed by the free undamped 'wet' bending modes of the hull in water, by the 'dry' modes of the same structure in air, as recommended by Bishop and Price, 1 or more simply the orthogonal modes of a uniform beam, as shown by Gran. 2 An even simpler representation can be developed in terms of orthogonal polynomials, despite the fact that these lack a physical basis and do not satisfy the appropriate free-end boundary conditions. Analogous options exist in defining the modes for discontinuous deflection of multiple bodies. The usual definition of each mode is with one body performing a unit amplitude of motion and the others fixed. If N separate bodies oscillate independently, with six degrees of motion for each body, the ensemble has a total of 6N degrees of freedom. The individual bodies, and modes of

48

J. N. N e w m a n

motion, can be denoted by a pair of corresponding indices, denoting for example the amplitude of motion for body n in m o d e j by the symbol ~n-). However this is an awkward notation which obscures the analogy with other applications. It is more convenient to consider a single 'global body', defined geometrically by the ensemble of N separate bodies. For this example there are a total of 6N modes of the global body, or more if the individual bodies are deformable. Any specified distribution of normal velocity on the (global) body surface is considered to be a 'mode' in the present work. The decomposition of complicated body motions into separate modes is justified by linear superposition, provided suitable modal functions are included. The appropriate definition of each mode is non-unique, and subject to choice. In specific applications the choice of modes can be based on physical considerations, and computational efficiency. The case of two rigid bodies joined by a hinge or pin joint is a simple example, applicable to a pair of floating barges joined together by an equivalent mooring arrangement. The hinge constraint reduces the total number of degrees of freedom from twelve to seven, including six rigid modes of the global body plus one angular deflection of the hinge. Solving the latter problem directly avoids the unnecessary solution for five extra radiation potentials, and permits the equations of motion to be solved in a consistent manner without imposing ad hoc constraints. Similar considerations apply if the method of images is used to analyse the motions of a single body near a vertical wall, or if a finite number of images are used to approximate the effects of two parallel walls in a wave channel. If the total number of images is I, the general problem for N = I + 1 separate bodies involves 6N rigid-body motions, but in the usual application the motion of each image is either the same or opposite to that of the actual body. Solving the latter problem directly, with appropriate definitions of the normal velocity on each image, reduces the total number of radiation potentials by the factor 1/N. Section 2 formulates the linear hydrodynamic analysis for a global body with arbitrary specified modes of motion. This analysis leads to straightforward extensions of the usual added-mass, damping, and excitingforce coefficients for a rigid body. Special attention is required for the hydrostatic coefficients, following the analysis in Ref. 3. Illustrative applications are then described in Sections 3-6. For each of the above applications the hydrodynamic analysis is performed using the three-dimensional radiation/diffraction code WAMIT, based on the boundary-element 'panel' method. Further details regarding this code are given in Refs 4-7. To provide a computational basis for analyzing the general problem of a body with arbitrary specified modes, the program has been extended so that the user can specify the

functional form of generalized modes in addition to the six conventional rigid-body motions. Provision is made to specify external mass, damping, and stiffness matrices with coefficients corresponding to each of the modes of motion. For each of the applications described below, discretizations with increasing numbers of panels have been used to verify numerical convergence. For example, in the analysis of the vertical column in Section 4, two discretizations (with N = 512,2048) total panels on the column are employed, and the results compared; the maximum relative difference observed in the modal amplitudes is 2%. Assuming the numerical errors are proportional to l / N , as in Ref. 7, it follows that the results presented, based on the finer discretization, are accurate within 0"5%. Similar convergence tests and error estimates have been made for all of the results presented.
2 FORMULATION AND FIRST-ORDER PRESSURE FORCE

Time-harmonic motions of small amplitude are considered, with the complex factor e i~t applied to all firstorder oscillatory quantities. The boundary conditions on the body and free surface are linearized, and potential flow is assumed. Cartesian coordinates are defined with z = 0 as the plane of the undisturbed free surface and z = - h as the bottom. Within the fluid domain is a body (or ensemble of bodies) with wetted surface Sb. The body may be floating, submerged, or bottom-mounted. Six conventional rigid-body motions (surge, sway, heave, roll, pitch, yaw) are denoted by the indices j = 1,2,3,4,5,6, respectively. The same notation is extended to other modes of body deformation (j=7,8,...,J). Each mode may be defined by a vector 'shape function' Sj(x) with Cartesian components uj, vj, wj. For rigid-body translation (j = 1,2, 3) the shape function Sy(x) is a unit vector in the corresponding direction, and for rigid-body rotations (j = 4, 5, 6) about the origin x = 0, Sj(x) = Sj_3 x x. The displacement of an arbitrary point within the body, due to motion in the jth mode with complex amplitude (y, is represented by the product ~jSj(x). The vector Sj is assumed to be continuous and differentiable near the body surface Sb, with divergence Dj = V-Sy. The divergence is zero for each rigid'body mode. The normal component of Sj on Sb is expressed in the form
nj = Sj . n = uynx + Vjny + Wjnz

(2.1)

The unit normal vector n points out of the fluid domain and into the body. The fluid velocity field is represented by the gradient of the velocity potential q~, governed by Laplace's equation in the fluid domain V2q~ = 0 (2.2)

Wave effects on deformable bodies


Appropriate boundary conditions on the free surface and bottom are 0~b matrices are defined in the form

49

Oz 02

K~b = 0, 0, on

on z

z= 0 -h

(2.3) (2.4)

OOi dS w2a i j - iwbii = -iwP I ISodpjn, dS = -P1ISo~PJ-~n


(2.13) Similarly, the generalized wave-exciting force is

0O

Here, K = w2/g, were g is the acceleration of gravity. The velocity potential of the incident wave is defined by

Xi = - i w p IJs0 qSnidS = - P J Jso (b ~n dS


(2.14) The indices i and j can take on any values within the ranges of the rigid-body modes (1-6) and extended modes (7,8,...). Green's theorem can be applied to (2.13) to establish reciprocity, and to (2.14) to derive the Haskind relations between the generalized exciting force and the corresponding radiation potential. The deformation of the body geometry must be considered in deriving the contribution to the generalized force (2.12) from the hydrostatic pressure -pgz, since this pressure is of order one. Three surfaces of integration are defined as follows: the initial wetted surface of the body prior to the modal displacement Sj is denoted as So; the corresponding deformed surface after the normal displacement in the mode j is denoted as S~; and the closed surface ~ is defined including So, S~, and the portion of the plane z = 0 lying between these two open surfaces if they intersect the free surface. On E the normal vector is defined in a consistent manner, to point out of the enclosed volume u. With these definitions, the change in the hydrostatic generalized force component Fi due to a unit displacement of the body in mode j is defined by the matrix

Ot = ig A~cOShcosh[k(Zkh + h)] e_i k(x cosl3+ y sinjS)

(2.5)

where the wavenumber k is the positive real root of the dispersion relation
032

- - = ktanhkh
g

(2.6)

and/3 is the angle between the direction of propagation of the incident wave and the positive x axis. The velocity potential ~bcan be expressed in the form = 0R + Or) where 0z) = Ot + ~bs (2.8) is the diffraction potential, Os is the scattering component representing the disturbance of the incident wave by the fixed body, and
J

(2.7)

OR = Y~ ~j~j
j=l

(2.9)

is the radiation potential due to the body's motions. In each mode, Oj is the corresponding unit-amplitude radiation potential. On the undisturbed position of the body boundary, the radiation and diffraction potentials are subject to the conditions

ci,=,gIj
cij pg

$6

,,;dS-pgJJSo z,,,dS=,gJJz.,dS
(2.15)
(2.16)

Using (2.1) and the divergence theorem

OqSj _= iwnj On 0~ =0
On

(2.10)

JI zSi'lldS-~PgJIJ v V'(zSi)dV

(2.1])

The boundary-value problem is completed by imposing a radiation condition of outgoing waves, for the potentials ~ and Oj. Corresponding to each mode of motion, generalized first-order pressure forces are defined in the form
/

For small deformations the volume u is thin, and the last integral can be approximated to first order as the surface integral of the product of the integrand and the distance nj between the two boundary surfaces So and S~. Thus

cij=PgIIsonJV'(2Si)dS=pgIIsonj(wi-k-zDi)dS
(2.17) Here the generalized force is defined in a fixed reference frame, and only the hydrostatic pressure is considered. As a result, (2.17) includes some contributions which normally are not considered, such as the roll (or pitch) moment due to a sway (or surge) displacement, both of which are equal to the displaced body volume times the moment arm associated with the corresponding displacement. Normally, for a freely-floating body, these

Fi=IIspnidS=-pIJs(iw(a+gz)nidS
b b

(2.12)

Here p is the fluid pressure, which is evaluated in the last form of (2.12) from the linearized Bernoulli equation. The contribution to (2.12) from the term iwO involves a trivial extension of the corresponding rigid-body analysis. After substituting (2.10) for the components of the radiation potential, added-mass and damping

50

J. N. Newman

contributions are balanced by the gravitational force due to the body mass. In the generalized analysis of a deformable body the corresponding mass force must be evaluated separately, for each mode, depending on the mode shape and mass distribution. The coefficients (2.17) depend not only on the normal displacement (2.1), but also on the divergence D~ and the vertical component of S~. In seneral the hydrostatic matrix (2.17) is not symmetric.3

and the stiffness matrix

Cij =

L/2 J-L~2

EIf. (x)fj (x)dx

(3.7)

3 BENDING OF A SLENDER BARGE As the first example of structural deflections, we consider a ship freely floating in head seas. The ship is assumed to be a slender beam, with vertical displacement Re(~(x)e i~t) along the length. The structural deflection is governed by the beam equation 1

The stiffness matrix (3.7) is derived by integrating the corresponding term in (3.4) twice by parts, and invoking the boundary conditions (3.2). Both (3.6) and (3.7) are symmetric matrices. The integral on the right-hand side of (3.4) includes contributions to the linearized pressure force Z(x) from the diffraction and radiation problems, as described in Section 2. Substituting (2.13-2.15), and moving the radiation coefficients to the left side of (3.4), gives a conventional set of equations of motion

Z ~J[-w2(aij + Mij) + iwbij + Cij + cij] = Xi


J

(3.8)

-w2m~ + ( E l ( ' ) " = Z(x)

(3.1)

where re(x) is the distribution of mass, and primes denote differentiation with respect to x. The product of the modulus of elasticity E and moment of inertia I is the stiffness factor El. Z(x) is the local pressure force acting on a vertical section of the ship. Since both ends of the ship are free, the appropriate boundary conditions are

Pertinent modes of motion in head seas include surge, heave, pitch, and the structural deflection. Two different sets of mode functions will be used for this case. The first are the natural modes for bending of a uniform beam with free ends, given by 2 1 (co__ss m2j___q cosh m2jqX~

f2j(q) = ~ ',,cos~2j ~ cosh~2j J


(3.9) 1 (Sl.n " ~2j~-lq + sinh ~2j+l q~ 5 \ s m '~2j+1

(' = O,

(EI~")' = 0,

for

x = +L/2

f2j+l(q) =

(3.2)

where L is the ship's length, and the origin is at the midship section. The deflection ~ may be expanded in an appropriate set of modes, in the form ~(x) = Z ~ f j ( x J ) (3.3)

Here the normalized coordinate is q = 2x/L, and j = 1,2,.... The degenerate functions f0 -- 1 and fl = q correspond to the heave and pitch modes respectively. The factors ~j are the positive real roots of the equation (-1) j tan ~j + tanh ~j = 0 The first four roots are ~2 = 2.3650, ~5 = 7.0686 Note thatfj(x) is an even or odd function o f x according as j is even or odd. The first four bending modes are shown as the solid lines in Fig. 1. These functions are orthogonal, with the normalized values fj (1) = 1, and the integrated values ~3 = 3.9266, ~4 = 5.4978, (3.10)

where the (complex) amplitude ~j of each mode is unknown. Adopting the method of weighted residuals (3.1) is multiplied by f~(x) and integrated along the length, to give the system of equations
~]

ej
=

fL/2

+ (ezf/, (x) ),,] dx


~(x)Z(x) dx
(3.4)

L/2 J-L~2

V_fi(q)fj(q)dq = 16ij for (i > 2) or (j > 2)


(3.11) where the Kroenecker delta 6ij is equal to 1 when i = j, otherwise zero. Assuming a uniform mass distribution, m is a constant and the mass matrix is diagonal, with the values fL/2 Mij = m J_t/2~(2x/L)fj(2x/Z) dx
= 1 M d-fllfi(q~eJ(q)dq = I M ' i J

These can be rewritten in the form


E (J[--od2Mij + Cij] =

J_L/2fi(x)z(x) dx

fL/2

(3.5)

where the coefficients on the left-hand side are the mass matrix

Mij =

fJ-L~2 mf(x)fj(x) dx

t/2

(3.12)

(3.6)

where M = mL is the total mass.

Wave effects on deformable bodies

51

1.00 ~
0.50

Table 1. Amplitude of each mode and sum (total deflection at the bow) based on the natural free-free beam modes (3.9) and on the Legendre polynomials, for the wave period 8 s

Mode, j

Natural modes

Legendre modes

'//I//

I~1
2 3 4 5 6 7 8 9 0-173821 0-010693 0"000051 0"000133 0.000011 0"000069 0"000003 0"000028

Sum
0"173821 0"173545 0"173494 0'173517 0"173528 0"173534 0"173537 0"173540

I~1
0"193646 0"013896 0"021882 0"003925 0"002072 0.000541 0.000051 0.000002

Sum
0"193646 0"193324 0-171505 0"171512 0"173581 0"173568 0'173517 0.173517

0.00

-0.50 -1.00 ~ - 1.00 v ~ q

-0.50

0.00

0.50

1.00

Fig. 1. Natural modes (3.9) (solid lines) and Legendre polynomial modes (dashed lines). The normalized coordinate q is equal to 5: I at the ends of the ship. The stiffness matrix is simplified in this case by noting that if E1 is constant, and the individual modes (3.9) satisfy the boundary conditions (3.2), the integrand of (3.7) can be written instead in the original form of (3.4) as the product f~(iv) where the second factor is the fourth derivative. From the differential equation
f/(iV)(q) _ tc4j~(q) =

0 (3.13)

and the orthogonality relation (3.11) it follows that

Cij = 4(EI/L 3)tc4 tSij

The alternative modal functions applied to this problem are the Legendre polynomials Pi(q), which are plotted for comparison in Fig. 1. These are simpler mathematically than the natural modes (3.9), but the off-diagonal elements of the stiffness matrix are nonzero for even values of m + n. As a specific example, a rectangular barge is considered with length L = 80m, beam B = 10m, and draft T = 5m. The total mass M is distributed uniformly throughout the volume, up to a 'deck' 5 m above the free surface. A constant stiffness factor 8EI/L = 0.1Ms -2 is assumed. The water depth is infinite. Table 1 compares the response in each of eight bending modes at a wave period of 8 s, using the two alternative mode shapes. The effectiveness of each is indicated by the magnitudes of the corresponding amplitudes (j, and by the convergence of the sums which represent the total magnitude of the bending motion at the bow (q = 1). The 'converged' values of these two sums agree to four significant figures, confirming that the Legendre polynomials are applicable to this problem even though they do not satisfy the boundary conditions (3.1). However the natural modes

(3.9) provide a more effective expansion. For example, the results using one or two natural modes give three or four decimals accuracy, respectively, in the deflected amplitude at the bow. By comparison, five Legendre modes are required to achieve the same precision. Additional computations have been performed using Chebyshev polynomials for the mode functions. There is no significant change in the rate of convergence relative to the Legendre polynomials. An advantage of the Legendre modes, for a uniform mass distribution, is that the mass matrix (3.12) is diagonal. For the Chebyshev modes there is coupling in the mass matrix between all even or odd modes, including coupling of heave or pitch with the corresponding symmetric or antisymmetric bending modes. Computed values of the response-amplitude operators in heave, pitch, and the first two bending modes are shown in Fig. 2 for the range of wave periods between 5 and 12s. The rigid-body modes of heave and pitch dominate the total vertical motions. For other values of the stiffness factor the bending-mode amplitudes are modified, approximately in inverse proportion to El.
2.00

1.50 PITCH 1.00

E
0.50

0*00

III I llIIil

Illllll

Jllll

III llll llll I lll~lllll

Vl III I I W l l l l l l l

I l l ' I~I I I 1 1 1

5.00

6.00

7.00

8.00

9.00

10.00

11.00

12.00

PER I OD

Fig. 2. Vertical amplitude at the bow due to heave, pitch, and the first two bending modes.

52

J. N. Newman
1.00

These results are based on a discretization with 2304 panels on the submerged portion of the barge. A total of 64 longitudinal, 16 transverse, and 8 vertical subdivisions are used, with cosine weight-factors to provide finer spacing near the comers and the free surface. Convergence tests comparing these results with a relatively coarse discretization (576 panels) indicate an accuracy on the order of 1% for the results in Table 1 and Fig. 2.

0.50

0.00

4 BENDING OF A VERTICAL COLUMN


The second example used to illustrate the analysis of structural deflections is a slender vertical column with its free upper end at the plane of the free surface z = 0, and with damped support at the bottom z = - h . For simplicity it is assumed that the column is structurally uniform along its length, with a constant stiffness factor El. A uniform distribution of mass m is assumed along the column, and a concentrated mass m0 is placed at the free surface z = 0 to account for the 'superstructure'. Regular incident waves of frequency w are assumed, causing a horizontal deflection Re(~(z)e i~t) of the column. The beam equation (3.1) is applicable, with the coordinate x replaced by z and the horizontal pressure force X'(z) on the right-hand side. However the boundary conditions are different in this case. Appropriate boundary conditions at the bottom are

-0.50

--l,00

1 I

$ I

I I

'

0.00

0.20

0.40

0.80

0.80

1.00

Fig. 3. Mode functions (4.6) for i -- 1,2,3,4. The normalized coordinate q is zero at the bottom and 1 at the free surface. Here the normalized vertical coordinate q increases from zero at the bottom to 1 at the free surface, P7 is a polynomial of degree n, and the factor ,/2 is introduced to satisfy the boundary conditions (4.I). The polynomials P f (q) can be derived from the Jacobi polynomials P(a'~)(2q- 1). s The orthogonalization relation implies that ~ = 0 and/3 = 4. The first four members are

=0,

~'=0

at z = - h

(4.1)

P~= I P~ = 6q - 5 P~ = 28q 2 - 4 2 q + 15
P~ = 120q 3 - 252q 2 + 168q - 35 More generally

At the upper end, where there is no bending moment and a concentrated inertial load

~" = O,

EI~" = -w2mo~

at

z= 0

(4.2)

Applying the method of weighted residuals, as in (3.33.8), gives the linear system

E ~J[--w2(aij + giJ) + itvbij + Cij] : yi


J

(4.3)

?n(q)

m=0

n (4 + 2n - m)! qn-m z--~(-1)mm!(n - m)!(4 + n - m)! (4.7)

where the mass matrix and stiffness matrix are defined by

Mij = m ~_hfi(z)fj(z) dz + m0ft.(0)fj(0)

(4.4) (4.5)

These polynomials are normalized with P~(1) = 1. The corresponding orthogonality relation is Ii p;(q)pT(q)q4 dq = 2i~Su +5 (4.8)

I -h

Here (4.5) is derived in a manner similar to (3.7) by partial integration, and the second boundary condition (4.2) results in an extra contribution to the mass matrix (4.4). Since the deflection is horizontal, and depends only on the vertical coordinate, there is no contribution from the hydrostatic matrix (2.17). An appropriate set of orthogonal polynomials fj(z) (j = 1,2,3,...) can be defined to represent the deflection, in the form
,5(z) = q2 Pj-l(q),

(q = 1 + z / h )

(4.6)

The first four functions (4.6) are plotted in Fig. 3. Each mode shape is proportional to q 2 = (1 + z / h ) 2 near the bottom, and equal to 1 at the free surface. For intermediate depths they oscillate with ( j - 1) zeros, within a range of approximately + 1 / 2 except close to the free surface where the amplitude increases to 1. In the notation of Meirovitch, 9 the modal functions (4.6) are admissible since they satisfy the geometric boundary conditions (4.1). With the modal functions defined by (4.6) the mass matrix (4.4) can be evaluated. After using the

Wave effects on deformable bodies


Table 2. C ~ n t s C. of the stiffness matrix (4.5); each entry in this table shoul~ be multiplied by the factor E1/i~
i j=l j=2 j=3 j=4

53

Table 3. Amplitude of each mode

Period 5"0 6-0 6"5 7"0 8"0 9"0 10-0

Mode 1 0"04813 0"29934 1"14505 0-44875 0-22035 0"16106 0"13055

Mode 2 0.00423 0.02551 0.09698 0"03795 0"01875 0.01391 0.01151

Mode 3 0.00010 0"00041 0.00135 0-00046 0-00019 0.00014 0.00013

Mode 4 0.00002 0-00007 0.00024 0"00008 0"00004 0.00004 0.00004

2 3 4

16 32 52

16 172 380 640

32 380 7356 5 12496 5

52 640 12496 5 220652 35

orthogonality relation (4.8)


Mij =

mh~

+mo

(4.9)

All elements of the stiffness matrix (4.5) are non-zero rational fractions. Table 2 gives the elements of this matrix for (i,j) < 4. As an illustration, consider a vertical circular cylinder of radius 10 m, in water of depth 200 m. The distributed mass along the length of the cylinder is assumed to be half of its displaced mass. A concentrated mass m0 equal to the total displaced mass (twice the distributed mass) is located at the free surface. The stiffness of the cylinder is chosen such that the ratio EI/h3= 0.41m0s-2; this results in a resonance of the first bending mode at a wave period of 6.5 s. Computations have been performed including the first eight modes (4.6), at a sequence of wave periods between 5 and 10 s. The amplitudes of the first and second modes are plotted in Fig. 4. The amplitudes of the higher modes are too small to indicate in this figure. Also shown in Fig. 4 is the total response at the upper end of the column. The first and second modes generally are out of phase. As a result, the total response is about 10% less than the amplitude of the first mode. The most important feature in Fig. 4 is the highly1.2
IL
I t I I t

1.0

0.8

0.6

0.4

0.2

0.0 5

Period

10

Fig. 4. Amplitudes of the first bending mode (short dashed curve), second mode (long dashed curve), and the total response of the vertical column at the upper end (solid curve).

tuned resonance at 6"5 s, with a maximum horizontal deflection at the free surface slightly greater than the incident wave amplitude. The only damping which is included here is that due to linear wave radiation, which at resonance is 2"6% of the critical damping. The occurrence of this resonance can be attributed to the fact that the wave excitation, and damping are confined to a relatively small part of the column near the free surface. This is in contrast to the floating barge in Section 3, where the entire vessel is near the free surface. Table 3 shows the computed amplitudes of the first four modes at representative wave periods, confirming the rapid convergence of the modal expansion. It should be noted that in the numerical approach followed, the boundary conditions (4.2) are not explicitly imposed as constraints on the solution. Instead they are utilized only in the weaker context of the end conditions in the partial integration which yields the mass and stiffness matrices (4.4-4.5). One can verify numerically that the boundary conditions (4.2) are in fact satisfied by the numerical solution, but the relatively slow convergence of higher derivatives hinders this confirmation. For example, computed values of I~"(0)[ decrease from a maximum of 0.005 to 0.0004 when the number of modes is increased from four to eight. Even with eight modes included, the two sides of the second boundary condition (4.2) differ in the third significant figure (in the second significant figure for the lowest wave period). It is interesting to compare the use of natural modes with the method adopted here, based on the orthogonal polynomial modes (4.6). Natural modes could be used, as in Section 3, accounting for the boundary conditions (4.2) at the upper end of the column. Since the second condition (4.2) would only be satisfied at the corresponding eigenfrequency for each mode, this approach would not offer any computational advantage relative to the simpler polynomial modes used here. The computations shown here are based on a discretization of the column with a total of 2048 panels, corresponding to 128 panels uniformly spaced around the circumference and 64 vertical subdivisions between the free surface and bottom. The vertical spacing uses a cosine weight factor to give a finer subdivision near the free surface and a relatively coarse spacing near the bottom. As noted in Section 1, a comparison with computations using N = 512 panels

54
~Otn

J. N. Newman 40m

~m

Fig. 5. Configuration of the hinged barges. shows differences on the order of 2%, indicating that the accuracy of the results presented here with N = 2048 is approximately 0"5%. For this special case, where the body is a vertical circular cylinder of constant radius extending from the bottom to the free surface, the hydrodynamic analysis could be performed more directly using Havelock's 1 orthogonal eigenfunction expansion. The panel method used here has the advantage that it is applicable to any practical body of arbitrary three-dimensional form. In applications where the stiffness factor E I and mass distribution m are non-uniform, their functional dependence on z must be included in the integrals (4.4-4.5). This may require numerical evaluation of these integrals, and will result in non-zero off-diagonal elements of the mass matrix, but no fundamental difficulties are envisaged.
5 MOTIONS OF A HINGED BARGE

If separate rigid vessels are connected mechanically, the resulting modes of motion of the global body include both rigid modes of the ensemble and discontinuous relative motions between the separate vessels. As an example we consider two rectangular barges, each of which is 40m long by 10m beam and 5m draft, connected end-to-end by a simple hinge with a gap of 10m separating the two barges (Fig. 5). The hinge axis is located at the origin (x = 0, z = 0), midway between the two barges in the plane of the free surface. Each barge is assumed to be in static equilibrium, with uniform mass distribution throughout its length, beam, and vertically in the range ( - 5 < z < 5m). The water depth is assumed to be infinite. Incident head waves are considered, with resultant motions in surge (~1), heave (~3), pitch (~5), and in the additional hinged mode (~7)- The first three modes are defined conventionally, with heave the vertical motion at the hinge point and pitch the average angular position of the two barges. In the hinged mode the right-hand and left-hand barges are pitched about the y axis to angles +~7, respectively. Since the geometry is symmetric about the hinge axis, and the displacement in this mode is symmetric, there is no coupling with surge or pitch, and the latter modes are not affected by the hinge deflection. The response-amplitude operators for heave and the hinge deflection are shown in Fig. 6. These two modes are dynamically similar, and their phase angles (not shown) are practically identical. Since a positive hinge deflection implies negative vertical motion at the outer

ends of the barges, there is substantial cancellation resulting from these two modes. To illustrate this point, the heave amplitude at the midship position of each barge (x = +25 m) is also shown in Fig. 6. Also shown by the dashed line in Fig. 6 is the heave amplitude of the same pair of barges when the hinge is rigid (~7 = 0). Except for a small shift of the peak wave period and amplitude, this response is practically identical to the heave response of the hinged barges as measured at their midship sections. At a wave period of 8 s, corresponding to a wavelength of 100m, there is practically no heave motion of the rigid configuration, or of the hinged configuration at the midship sections. The results shown in Fig. 6 are based on a descretization with 2460 panels on both barges, using 32 longitudinal subdivisions on each barge, 16 transverse subdivisions and 8 vertical subdivisions. Symmetry about the planes x = 0 and y = 0 permit the solutions for canonical symmetric and antisymmetric potentials, in one quadrant consisting of one side of one barge. Thus the total number of unknowns in each canonical sub-problem is reduced to 640. Cosine spacing is used to give smaller panels near the corners of the barge and near the free surface. Comparison with the results for a

3.00

2.50

2.00

! .50

1.00

0.50

0.00 5.00

8.00

7.00

8.00

9.00

tO.O0

it.O0

12.00

PER I OD

Fig. 6. Amplitudes of the symmetric modes of the hinged

barge. Heave ((3) is normalized by the incident wave amplitude A and angular deflection in the hinged mode (~7) is normalized by the wave slope KA. The lower solid curve shows the heave amplitude at the midship section of each separate barge, and the dashed line shows the heave amplitude of the same pair of barges with a rigid connection instead of a hinge.

Wave effects on deformable bodies


coarse discretization using a total of 640 panels shows maximum differences in the range 1-5%. This implies an accuracy of 1% or better for the results plotted in Fig. 6. In summary, the vertical motions of the hinged and rigid configurations are practically identical at the midship sections of each barge, but substantially greater for the hinged configuration at the hinge axis (by a factor of about four at resonance).

55

Fig. 7. Perspective of the vertical cylinder and four images

6 WALL EFFECTS IN A CHANNEL Wave interactions with a body in a channel of rectangular cross-section may be analyzed by the method of images. To illustrate this type of problem we consider the case of a vertical circular cylinder of finite draft, centered in a rectangular channel. This problem has been studied by Yeung and Sphaier H and also by Linton and Evans; 12 in both of these works the geometry of the body is restricted, but the images are accounted for in a more analytic manner. Here we employ a finite number of image bodies, and study the convergence as this number is increased. Assuming this procedure can be carried out successfully, it offers the possibility of performing computations for more general body shapes in a channel. To permit comparison with the results in the above references, the dimensions are chosen such that T = 2a, d = 2a, and h = 10a, where a is the cylinder radius, T its draft, d the channel half-width, and h the channel depth. A single array of image cylinders is required, with the axes at y = + 2 n d ( n = 1 , 2 , 3 , . . . , N ) . Thus there is a total of 2N + 1 cylinders including the central body and its images on both sides. The computations are performed with N = 0, 1,2, 4, 8, corresponding to arrays with 1,3,5,9,17 cylinders. Each cylinder is discretized in an identical manner, with 32 azimuthal subdivisions, 8 vertical subdivisions on the side, and 8 radial subdivisions on the bottom, giving a total of 512 panels on each cylinder and up to 8704 panels for the largest array with 17 separate bodies. Since the geometry is symmetric about x = 0 and y = 0, the number of unknowns is reduced by a factor of four. The configuration with N = 2 is shown in Fig. 7. Singular features must be expected as a result of the channel walls. The simplest of these are the cut-off frequencies, which for symmetric propagating waves occur at kd = 7r, 27r, 37r,..., corresponding to k a - 1.57, Y 1 4 , 4 - 7 1 , . . . . For antisymmetric waves, kd = lr/2,31r/2,..., corresponding to /ca = -0-79, 2 - 3 6 , . . . . Here the relevant plane of symmetry is the channel centerplane, hence for a body with geometric symmetry about this plane the modes of surge, heave and pitch and the diffraction problem are symmetric whereas the modes of sway, roll and yaw are anti-

showing the discretization with 512 panels on each body. The channel walls are indicated by the straight lines, in the plane of the free surface.
symmetric. In addition to the cut-off frequencies, Linton and Evans n have shown that an antisymmetric trapped wave exists at k d - 1.41, corresponding to k a - 0.70. Irregular frequencies also must be anticipated in the present computations, which are based on the solution of a boundary-integral equation using the free-surface Green's function. The irregular frequencies which affect the heave force correspond to the zeros of the Bessel function Jo(ka), occuring at ka'-2.40,5.52, . . . . For surge and sway, the zeros of J~(ka) are relevant, occurring at ka-" 3.83,7.02,... Methods exist to remove the effects of the irregular frequencies, as described by Lee and Sclavounos, 13 but the present results are shown without this correction to illustrate the resulting numerical errors. Figure 8 shows the surge exciting force (pitch is similar). The computations for the five different arrays are indicated by dashed lines of increasing length, and the results of Yeung and Sphaier are shown by the solid line. The wall effects are relatively small, and the results depend only weakly on the number of images.
4.00 / /.--\ ,\

3.00 2.00i
1.00

"

0.00 0.00

1.00

2.00

ka

3.00

4.00

5.00

Fig. 8. Surge exciting force for the cylinder in a channel. The dashed lines of increasing length denote arrays with 1,3,5,9, and 17 bodies. The solid line is reproduced from Yeung and Sphaier.I i

56
3.00

J. N. Newman

5.00

4.00
2.00

'"

i~

1.00

/
,.00
\.\

/
1.00

"~
.... ..x

,/

0.00
0.00

. . . . . . . . .

"r

. . . . .

,eT";-,

. . . .

0.00

1.00

2.00

3.00

0.00

,,,,,,,,,i,,,,,,,,,i,,,,,,,,,t,,,,,,,,,iT,,,,,,,,i
2.00
ka

3.00

4.00

5.00

Lo

Fig. 9. Heave exciting force for the cylinder in a channel. (See Fig. 8 for details.) Increasing the number of images tends to improve the comparison with the results of Yeung & Sphaier, and the most apparent difference occurs in the vicinity of ka = 7r, corresponding to the second cut-off frequency for the channel. A weaker singular feature is apparent in the vicinity of ka = 4-7, near the third symmetric cut-off frequency. The results near the first cut-off frequency are remarkably uniform. Figure 9 shows the corresponding results for the heave exciting force, where the wall effects are minor. A very small 'hump' is apparent at the first irregular frequency (ka = 2.40). (Both exciting forces are computed using the Haskind relations, with the integration carried out over the body surface.) The added-mass and damping coefficients in surge, sway, and heave are shown in Figs 10-15. For surge

Fig. 11. Surge damping coefficientfor the cylinder in a channel. (See Fig. 8 for details.) (Figs 10 and 11) the principal feature occurs near the first symmetric cut-off frequency k a - 1 . 5 7 . There is no corresponding feature apparent at ka - 4.71. For sway a much stronger feature is present near the trapped-wave frequency k a - 0 . 7 0 , and first antisymmetric cut-off frequency k a - 0 7 9 . When the channel walls are accounted for exactly, the sway damping is zero below the first cut-off frequency. The finite arrays demonstrate Stokes' phenomenon, oscillating about zero in an attempt to mimic the sharp cut-off. A weaker feature is apparent at the second and third cut-off frequencies k a - 2.36, 3.92. The heave added mass and damping do not exhibit such distinctive features. There is a substantial increase in the high-frequency added mass due to the channel walls, but even the array with two
10.00

7.00

/''t ',
6.00 //I! 5.00 ~\ 5.00

p.( ...
..... "

iii!
*" .. ....... .-~4,.~ ~_ ~. ...... ..2~..--~ ..-

i
4.00

7 / / '~ 1
'

~, \\

0.00

'~

, (/
!
-5.00 II

3.00

2.00

r' } ~/
Ii I
..... I~ , , . . . . . . . . . , ......... ~......... ,,,, ......

1.00

-10.00

0.00

! .00

2.00
La

3.00

4.00

5.00

0.00

1.00

2.00
La

3.00

4.00

5.00

Fig. 10. S u r g e a d d e d - m a s s coefficient f o r t h e c y l i n d e r in a

Fig. 12. S w a y a d d e d - m a s s coefficient f o r t h e c y l i n d e r in a

channel. (See Fig. 8 for details.)

channel (See Fig. 8 for details.)

Wave effects on deformable bodies


20.00 2.00

57

t% 15.00
1.50

\~,~
10.00
1.00
.....

fi.O0
~ ,,"~.~

0.50

\ \\

~"N \\ "',

......._~ ~2

'.~

0.00

.....

........

0.00

1.00

2.00
Lo

3.00

4.00

5.00

0.00 0.00

0.20

0.40 to

0.60

0.80

1.00

Fig. 13. Sway damping coefficientfor the cylinder in a channel. (See Fig. 8 for details.) images gives a close approximation to the added mass in this regime. For low frequencies the added mass and damping are influenced significantly by the walls, and the different arrays show substantial differences. The low-frequency limit of the added mass is unbounded due to the walls, as in the case of a two-dimensional body. Except in the low-frequency regime there is good agreement between the present results and the more exact computations of Linton and Evans. These results indicate that the role of the cut-off frequencies varies for different force coefficients. This is an obvious conclusion for different modes, for example between surge and sway, but a comparison of different coefficients for the surge mode reveals a more subtle
2.50

Fig. 15. Heave damping coefficient for the cylinder in a channel. (See Fig. 14 for details.) distinction. As noted above, the exciting force shown in Fig. 8 is remarkably uniform at the first cut-off frequency, with a substantial irregular feature at the second cut-off frequency, whereas the added mass and damping (Figs 10 and 11) are precisely the opposite. To attempt to explain this difference, it may be noted that the pressure force due to a free wave which is symmetrical about the centerplane y = 0, and subject to cut-off in a channel with wails at y = +d, is proportional to the integral Ji ~rcos 0 sin (kxa cos 0) cos (kya sin O)dO Here k 2 + ~ = k 2 , k y = m r / d ( n = l , 2 , 3 , . . . ) , and 0 denotes the polar angle around the cylinder. With respect to the added mass and damping, one would expect that the discontinuity near the nth cut-off frequency might be qualitatively proportional to this integral, whereas the Haskind exciting force would contain an extra weight factor cos(kacosO) associated with the incident-wave potential. Thus one might anticipate that the differences in irregularity near the first two cut-off frequencies migh be associated with different relative magnitudes of these integrals. This is borne out by computations, insofar as the integral for the added mass and damping is about four times greater than for the Haskind exciting force at the first irregular frequency, and about four times smaller at the second. Note however that all of these integrals vanish at the cut-off frequencies, in proportion to the factor kxa. Thus the possibility remains that this is not a complete explanation of the above differences. One computational problem which should be recognized is that the total number of panels, and unknowns, is proportional to the number of image bodies. For the largest array, with 17 separate bodies and 512 panels on

I L r, ,,
s

2.00

1.50

jd

l.O0

Jvl ,llllll

Jll'lrllll

JlITl'lJtlrllrl

Jilt

Jl41rlUUlt

t.00

1.00

2.00
ka

3.00

4.00

5.00

Fig. 14. Heave added-mass coefficientfor the cylinder in a


channel. The dashed lines of increasing length denote arrays with 1,3,5,9,and 17 bodies. The solid line is reproduced from Linton and Evans. i2

58

J. N. Newman

each body, a total of 8704 panels are used. Exploiting the geometric symmetry of this configuration reduces the total number of unknowns to 2176. One possibility which has not been explored is to decrease the number of panels on the images as the distance from the central body increases. In this manner it may be possible to reduce substantially the total number of panels, without degrading the accuracy of the computed forces on the body itself. The use of a truncated image array with a finite number of image bodies has been criticized in the more analytic work of Linton and Evans, due to the incorrect form of the resulting solution in the far field. However for computations of the first-order force coefficients acting on a body in a channel, and for other local flow parameters, the error in the far field seems irrelevant. The present results suggest that practical computations can in fact be performed with a relatively small number of images, except for the singular features which occur near cut-off frequencies. The significance of these singular features is reduced in practice by imperfect wall reflections, and by transient effects which attenuate the resonant reflections. Two issues which have not been addressed here are the effects of channel walls on the local pressure field and on the second-order mean drift force. Mclver 14'15 has noted that these quantities are affected strongly by wall effects. The requirement of large numbers of panels may be a significant barrier for computations of the mean drift force at the relatively large wavenumbers shown in Figs 10-15.

7 DISCUSSION This paper extends the methodology of three-dimensional radiation and diffraction based on the panel method of computation to the analysis of various deformable body motions. The continuous deflections which are illustrated for the vertical column and floating barge are typical of structural deflections. The discontinuous motions of the hinged barge and array of cylinders are examples where the physical application is quite different. The specific examples used for illustration are geometrically simple, but the computational approach is applicable to more general bodies. For structural deflections the practical importance of a complete hydrodynamic analysis depends on the particular application. For relatively stiff freely-floating bodies, including conventional ships, the frequencies of hull vibrations are substantially higher than those of first-order wave excitation, and the amplitudes of these motions are much smaller than the rigid-body modes. Thus the hydrodynamic and structural analyses can be carried out separately. However this simplified procedure is not appropriate if the body is more flexible. The slender barge analysed in Section 3 illustrates the

methodology which can be used in such cases. For the slender vertical column treated in Section 4, a highlytuned resonant bending deflection occurs within the range of significant wave energy, increasing the importance of the structural deflection. As the development of offshore platforms extends into deeper waters this type of problem may become particularly important. One point emphasized in both of these examples is the feasibility of using mathematical mode shapes which differ from the more conventional physical modes of the body. The vertical column is an example where the use of orthogonal polynomials is particularly effective, and where the corresponding natural modes would be relatively complicated due to the concentrated mass at the free surface. Conversely, the floating barge illustrates the relative efficiency of using natural modes of a beam with free ends, although the use of orthogonal polynomials also is feasible. It is customary in engineering applications to perform a dynamic analysis of the specific structure, ignoring the effects of the surrounding fluid to determine the corresponding natural 'dry' modes. In the case of a ship, for example, these would be the actual physical modes of free vibration of the hull, at different eigenfrequencies, accounting for the distributions of structural stiffness and mass throughout the ship and neglecting all hydrodynamic and hydrostatic pressure forces. For cases where the stiffness and mass are substantially continuous, such an approach seems unnecessarily complicated by comparison with the alternatives of representing the waveinduced deflections in terms of either the natural modes (3.9) for a simple uniform beam, or appropriate orthogonal polynomials. The hinged barge, which is considered in Section 5, illustrates the methodology for studying multiple bodies with mechanical constraints. This type of problem could be analysed in a more general manner by considering each separate body to have six independent degrees of freedom, and imposing the constraint in a special postprocessor. Thus the four relevant modes of surge, heave, pitch, and hinge deflection for the hinged barge would be represented by six modes (surge, heave and pitch for each separate body), and two constraints would be imposed stating that the horizontal and vertical motion at the hinge point is the same for both barges. Such a procedure is awkward from the computational standpoint, since the radiation potential of the extra modes must be evaluated, and the constraints are imposed in an ad hoc manner. The present approach avoids these complications. The problem of wall effects on a cylinder, treated in Section 6, is another application of discontinuous modes. In this case generalized modes are used to represent the appropriate motions of each image body, so as to satisfy the boundary conditions on the walls. This problem also could be treated with independent modes of each image body, but the relatively large

Wave effects on deformable bodies

59

number of images and corresponding independent modes would substantially increase the computational burden. ACKNOWLEDGEMENTS This work was performed as part o f the Joint Industry Project 'Wave effects on offshore structures', supported by the Chevron Oil Field Research Company, Conoco Norway, Mobil Oil Company, the National Research Council of Canada, Norsk Hydro, Offshore Technology Research Center, Petrobr~s, Saga Petroleum, the Shell Development Company, Statoil, and Det Norske Veritas Research. Tabular data files for the solid curves shown in Figs 9 and 10 were kindly provided by the authors of Refs 11 and 12. REFERENCES 1. Bishop, R. E. D. & Price, W. G., Hydroelasticity of ships, Cambridge University Press, Cambridge, 1979. 2. Gran, S. A., Course in Ocean Engineering, Elsevier, Amsterdam, 1992. 3. Newman, J. N., Deformable floating bodies. 8th International Workshop on Water Waves and Floating Bodies, St John's, Newfoundland, 1993. 4. Korsmeyer, F. T., Lee, C.-H., Newman, J. N. & Sclavounos, P. D., The analysis of wave interactions with tension leg platforms. Proc. Conference on Offshore Mechanics and Artic Engineering, Houston, 1988.

5. Newman, J. N. & Sclavounos, P. D., The computation of wave loads on large offshore structures. Proc. Conference on the Behaviour of Offshore Structures, Trondheim, 1988. 6. Lee, C.-H., Newman, J. N., Kim, M.-H. & Yue, D. K. P., The computation of second-order wave loads. Proc. Offshore Mechanics and Arctic Engineering Conference, Stavanger, 1991. 7. Newman, J. N. & Lee, C.-H., Sensitivity of wave loads to the discretization of bodies. Proc. Conference on the Behaviour of Offshore Structures, London, 1992. 8. Abramowitz, M. & Stegun, I. A., Handbook of Mathematical Functions, US Government Printing Office, Washington, DC 1964. 9. Meirovitch, L., Analytical Methods in Vibrations, Macmillan, New York, 1967. 10. Havelock, T. H., Forced surface-waves on water. Philosophical Magazine, 8 (1929) 569-76. 11. Yeung, R. W. & Sphaier, S. H., Wave-interference effects on a truncated cylinder in a channel. J. Engineering Mathematics, 23 (1989) 95-117. 12. Linton, C. M. & Evans, D. V., The radiation and scattering of surface waves by a vertical circular cylinder in a channel. Phil. Trans. Roy. Soc. Lond., A, 338 (1992) 325-57. 13. Lee, C.-H. & Sclavounos, P. D., Removing the irregular frequencies from integral equations in wave-body interactions. J. Fluid Mech. 207 (1989) 393-418. 14. McIver, P., The wave field around a vertical cylinder in a channel. 7th International Workshop on Water Waves and Floating Bodies, Val de Reuil, France, 1992. 15. McIver, P., Recovery of open-sea results from narrow tank tests. 8th International Workshop on Water Waves and Floating Bodies, St John's, Newfoundland, 1993.

You might also like