Download as pdf or txt
Download as pdf or txt
You are on page 1of 21

Wave-induced motion

royalsocietypublishing.org/journal/rspa of rigid bodies: beads, boats


and buildings
P. A. Martin
Review
Department of Applied Mathematics and Statistics, Colorado School
Cite this article: Martin PA. 2023 of Mines, Golden, CO 80401, USA
Wave-induced motion of rigid bodies: beads,
PAM, 0000-0002-0384-3885
boats and buildings. Proc. R. Soc. A 479:
Downloaded from https://royalsocietypublishing.org/ on 04 March 2023

20220463. The determination of rigid-body motion caused by


https://doi.org/10.1098/rspa.2022.0463 incident waves is a familiar problem in mechanics.
Good examples are the motion of a ship in the
Received: 4 July 2022 presence of water waves and the motion of a rigid
Accepted: 26 January 2023 structure in the presence of seismic waves. The
basic goals are to determine the motion of the rigid
body and the effects of the motion on the wave
Subject Areas: field, assuming linear theory is adequate. Although
applied mathematics, wave motion, the underlying mathematical problems are similar,
several solution methods have evolved, depending on
mechanical engineering
the physical problems of interest. For ship motions,
the standard approach is to decompose the problem
Keywords: into seven subproblems, one for each of the six
rigid-body motion, ship hydrodynamics, rigid-body modes and one to take account of the
soil–structure interaction, Haskind’s relation, incident wave. This approach is reviewed and then
boundary integral equations, time-harmonic adapted to problems in acoustics and to problems
motion in elastodynamics, such as those that arise in simple
examples of soil–structure interaction. It is argued that
the resulting approach for elastodynamic problems
Author for correspondence: has clear advantages over those currently in use.
P. A. Martin
e-mail: pamartin@mines.edu

1. Introduction
The dynamics of three-dimensional rigid bodies is a
familiar topic in classical mechanics [1–4]. For example, it
is well known that the general motion of a rigid body can
always be reduced, exactly, to a translation and a rotation:
there are six degrees of freedom (DOF).
We are interested in situations where a rigid body
is coupled to an unbounded exterior medium: wave
motion in that medium causes the rigid body to move.
We assume throughout that linear theory is adequate,
so that the motion of the rigid body is small with
respect to its equilibrium position. However, we also

2023 The Authors. Published by the Royal Society under the terms of the
Creative Commons Attribution License http://creativecommons.org/licenses/
by/4.0/, which permits unrestricted use, provided the original author and
source are credited.
assume that the rigid body has finite mass, so that the equations of motion of the body are to be
2
taken into account. Determining when linear theory may be adequate requires some dimensional
analysis, but that is outside the scope of the paper.

royalsocietypublishing.org/journal/rspa Proc. R. Soc. A 479: 20220463


..........................................................
We are mainly concerned with time-harmonic problems: the time dependence is taken as e−iωt ,
where ω is the frequency. A typical problem would be: given an incident wave, determine the
scattered waves and the motion of the body. This motion is driven by time-harmonic forces and
moments applied to the surface of the rigid body (ignoring body forces, for simplicity). These
forces and moments come from the unbounded medium outside the body. The body’s motion
also generates radiated waves: therefore, the problem is genuinely coupled.
The exterior medium can be fluid or solid. In the former case, suppose first that the fluid is
inviscid and compressible: this is the realm of acoustics, where the fluid motion is governed by
the Helmholtz equation. Application areas include suspensions or aerosols, with solid particles in
Downloaded from https://royalsocietypublishing.org/ on 04 March 2023

a liquid or gas [5,6], sonar [7] and fish bioacoustics: it is thought that the otolith organs in the ears
of fish are sensitive to motion [8,9]. General formalisms for such acoustic problems were given
by Haskind (Khaskind) [10] and Olsson [11]. However, almost all calculations have been made
for a spherical rigid body and an incident plane wave: the method of separation of variables
can be used, and the equations of motion of the body reduce to a single differential equation
(see §5a). This basic method was described by Lamb ([12], §298). Subsequent papers include
[7,13–15]. There are also papers on related time-domain problems [16], where a plane sound
pulse interacts with a rigid sphere [17–19]. (For a good survey of relevant Russian work, see
[20], ch. 6.) Olsson [11] has given numerical results (using null-field/T-matrix methods) for time-
harmonic scattering by rigid bodies of various shapes including spheroids and thick spheroidal
shells.
Another class of problems concerns inviscid incompressible fluids (water) with a free surface:
gravity waves can propagate along the surface and interact with a floating body. Obviously,
such problems arise in ship hydrodynamics, where they have been studied extensively [21,22],
but there are more modern applications such as to floating wind turbines [23–25]. The textbook
procedure is to decompose the full problem into seven problems, one for each of the six DOF of the
rigid-body motion and one for scattering the incident wave by the body in its fixed equilibrium
position (the diffraction problem). (Of course, symmetries can reduce the number of independent
problems to be solved.) This decomposition (which dates to the 1940s) is a simple consequence
of linearity but, surprisingly, it does not seem to have been exploited for acoustic problems
(for which it was introduced by Haskind [10]), or for elastodynamic problems (for which it
was introduced by Thau [26]). (Elastodynamic problems will be described below.) It turns out
that there is no need to solve the diffraction problem completely: the corresponding forces and
moments can be computed in terms of the solutions of the six basic radiation problems; see §4a.
This result [27], which is a consequence of Green’s theorem, is associated with the name of M. D.
Haskind (1913–1963), an engineer from Odessa, Ukraine.
Suppose next that the medium outside the rigid body is an elastic solid, extending to infinity
in all directions (full-space problem). Waves in the solid are incident on the rigid body, causing it
to oscillate, scattering waves away from the body. As before, the goal is to calculate the scattered
waves and the motion of the rigid body.
In 1900, Lamb ([28], §6) considered the scattering of a shear wave by a movable rigid sphere
in an incompressible solid. The general problem of elastodynamic scattering by a movable rigid
sphere has been solved [13,29,30], ([31], p. 622). For spheroids, see [32–35]. For the limiting case
of a thin circular disc, see [36,37]. (For scattering of flexural waves by movable rigid inclusions in
a thin plate, see [38].)
Another important class of problems arises when the external medium is a half-space (z > 0,
say) with a traction-free flat boundary (at z = 0). The rigid body may be completely buried, it may
be partially buried (analogous to the geometry for floating-body problems) or it may be attached
to the flat surface. Such problems are basic in the context of soil–structure interaction (SSI); for
reviews, see [39–43], ([44], §5) and ([45], §7.4). For good descriptions of early work, see [41,42]
and ([46], ch. 7). Much of this is concerned with the forced motion of a massless structure. The
literature on circular discs oscillating on the flat boundary, usually with simplifying non-welded
3
continuity conditions between the disc and the half-space, is extensive [41]; for a rigid disc welded
to the half-space, see [47].

royalsocietypublishing.org/journal/rspa Proc. R. Soc. A 479: 20220463


..........................................................
SSI problems can be tackled directly, as one large coupled system, and this may be appropriate
if the structure is complicated or if nonlinear phenomena are important; see [48] for a review.
However, SSI problems are often decomposed into subproblems in a variety of ways; see ([41], §7),
([42], §2), ([45], §7.4.3) and §6b. Most of these are distinct from that developed for floating-
body problems. In fact, taking that route was suggested by Thau [26] in 1967; in general, his
decomposition leads to seven radiation problems, each with prescribed displacements on the
surface of the rigid body (see §6a below). Then, it is perhaps not surprising that there is an
analogue of Haskind’s relation, implying that the diffraction problem (incident wave but fixed
rigid body) does not have to be solved completely; for a statement, see (6.10).
Downloaded from https://royalsocietypublishing.org/ on 04 March 2023

The paper begins in §2 with a simple one-dimensional example, in which a wave along a
stretched string interacts with a point mass (a bead) attached to the string. This problem can
be solved directly but it can also be solved using the kind of decomposition into subproblems
described above.
Then, as we are interested in rigid-body dynamics, we give a summary of the relevant
theory in §3. We include this section because we could not find a clear presentation in the
literature, keeping in mind that in order to couple to the exterior medium, we need to know
the displacement of the rigid body’s boundary (without introducing the angular velocity
vector). Note that for problems involving fluids, coupling is via the normal velocity but this is
proportional to the normal component of the surface displacement for time-harmonic problems.
Section 4 reviews methods for treating problems involving water waves and floating bodies.
The velocity potential in the water is decomposed into the sum of seven potentials. Each one
corresponds to a certain forced motion of the body, with normal velocity specified on the body’s
surface and outgoing surfaces waves far away. The potentials are combined in such a way that
the body’s equations of motion are satisfied. Haskind’s relation is derived, so that six potentials
have to be computed (rather than seven). These can be computed using standard software for
water-wave radiation problems.
The basic ideas coming from resolving floating-body problems in §4 are adapted to other
physical situations, starting with acoustic problems in §5. In fact, for those problems, all the
formulae from §4 apply without any substantial changes: Laplace’s equation is replaced by the
Helmholtz equation. One example is worked out in detail (plane wave scattering by a rigid
sphere): the calculations are straightforward and serve to show how the general formalism gives
the expected results. In the general case, again, the six potentials needed can be computed by
solving a standard boundary integral equation.
Elastodynamic problems are the subject of §6. We start with the full-space problem: a rigid
body is embedded in an elastic solid. Half-space SSI problems are discussed later (§6c). The
displacement in the exterior region is decomposed into the sum of seven displacement fields: this
is Thau’s decomposition [26]. Each one corresponds to a certain forced motion of the body, with
displacement specified on the body’s surface and outgoing waves far away. These elastodynamic
boundary value problems have to be solved for the associated traction vectors on the body.
Again, there is a form of Haskind’s relation, which eliminates the need to solve one problem.
The remaining six problems can be solved by standard numerical methods [44,49].
Alternative decompositions are discussed in §6b. It is argued that these are inferior to Thau’s
decomposition, even for half-space SSI problems. There are some concluding remarks in §7.

2. A simple one-dimensional example


Consider a string stretched
 along the x-axis, with uniform tension T, density per unit length 
and wavespeed c = T/. The lateral displacement of the string from its equilibrium position is
u(x, t). Time-harmonic waves of the form u(x, t) = Re{U e±ikx e−iωt } can propagate along the string;
here U is a constant and k = ω/c.
incident wave m
4

royalsocietypublishing.org/journal/rspa Proc. R. Soc. A 479: 20220463


..........................................................
u(x, t)
x
0

Figure 1. A snapshot of the bead on the displaced string at time t.

Now, attach a bead (a point mass m) at x = 0. Its equation of motion is


Downloaded from https://royalsocietypublishing.org/ on 04 March 2023

∂ 2u ∂u ∂u
m (0, t) = T (0+, t) − T (0−, t), (2.1)
∂t 2 ∂x ∂x
where the right-hand side is the restoring force provided by the tension in the string. In addition,
the displacement must be continuous at the bead, giving

u(0+, t) = u(0−, t). (2.2)

There is a wave incident on the bead from the left. This causes the bead to move laterally, with a
reflected wave and a transmitted wave. See figure 1. Suppressing the time dependence, we have

U(eikx + R e−ikx ), x < 0,
u(x) = (2.3)
U T eikx , x > 0,

where R and T are the (unknown) reflection and transmission coefficients. From (2.2), we have
1 + R = T, whereas (2.1) gives −mω2 T = Tik{T − (1 − R)}. Solving this pair of equations gives
−1 K 2ikT
R= , T= and K= . (2.4)
1+K 1+K mω2
For an alternative approach, one that will generalize to the rest of the paper, we introduce two
wavefields. The first (later called the diffraction field) is the solution of the problem when the mass
is held fixed: it is 
U(eikx − e−ikx ), x < 0,
ud (x) =
0, x > 0.

This says the incident wave is perfectly reflected, with no transmitted waves; note ud (0±) = 0.
The second wavefield is the solution of the radiation problem when the bead oscillates with
unit amplitude, in the absence of the incident wave. It is given by

1 e−ikx , x < 0,
u (x) =
eikx , x > 0;

waves are radiated symmetrically; note u1 (0±) = 1.


Finally, to determine the solution of the original problem, we take a linear combination,

u(x) = ud (x) + ξ1 u1 (x) for all x,

and then determine the constant ξ1 using the bead’s equation of motion. Note first that, by design,
u(0+) = u(0−) = ξ1 , so (2.2) is satisfied. Then (2.1) gives −mω2 ξ1 = Tik{ξ1 − (2U − ξ1 )}. Solving for
ξ1 gives ξ1 = UT, with T given by (2.4), as before.
This example is very simple because (i) the general solution of the governing wave equation
is available, and (ii) the bead can move with just one degree of freedom. These factors make a
direct treatment feasible (as seen when we started from (2.3)), whereas the alternative approach
has more general applicability. To recap, we introduced two solutions, ud and u1 , and then we
determined the correct linear combination of these two using the bead’s equation of motion. In
more complicated three-dimensional situations, we will have to solve seven problems, one for
5
each of the six DOF of the rigid body, and one to take account of the incident wave.

royalsocietypublishing.org/journal/rspa Proc. R. Soc. A 479: 20220463


..........................................................
3. Rigid bodies
(a) Basic formulation
Consider a rigid three-dimensional body B with boundary S and mass density . Introduce a fixed
system of Cartesian coordinates Ox1 x2 x3 with associated unit vectors e1 , e2 and e3 . Let x = x1 e1 +

x2 e2 + x3 e3 = xi ei , with the usual summation convention. The body has mass m = B (x) dx. Its

centre of mass is at the point G with position vector X G = m−1 B x(x) dx with respect to O. The
Downloaded from https://royalsocietypublishing.org/ on 04 March 2023

moment of inertia tensor relative to G, JG , is defined by (see [50, eqns (2.39) and (2.40)])

JG = (tr E)I − E, where (3.1)



E= (x − X G ) ⊗ (x − X G )(x) dx (3.2)
B

is the Euler tensor ([4], §7.8) and I is the identity; both JG and E are symmetric tensors. (Recall that
the tensor product u ⊗ v is defined by its action on an arbitrary vector a by (u ⊗ v)a = (v · a)u;
see [51], p. 21 and [4], appendix A.3.) Explicitly, let us introduce a body-fixed system of Cartesian
coordinates Gx1 x2 x3 with associated unit vectors e1 , e2 and e3 , and put x = xi ei . Then JG = JijG ei ⊗
ej where

JijG = (|x |2 δij − xi xj ) dV = JjiG ;
B

see, for example, ([3], eqn (10.140)), ([45], eqn (1.29)) and ([22], p. 156, eqn (144)), together
with (3.11).
We are interested in motions of B. Suppose points Xi in B have position vectors X i with respect
to O. During the motion, Xi moves to a point with position vector xi (t). As B is a rigid body,
|x1 (t) − x2 (t)| = |X 1 − X 2 |. Moreover, as orientations are preserved,

x1 (t) − x2 (t) = Q(t)(X 1 − X 2 ), (3.3)

where Q is a (dimensionless) rotation tensor. This means that Q is a proper-orthogonal second-


order tensor: it satisfies det Q = 1 and QQT = QT Q = I. Such tensors can be parameterized using
three independent parameters.
When x1 = x and X 1 = X, we can write (3.3) as

x(t) = Q(t)X + q(t) with q(t) = x2 (t) − Q(t)X 2 . (3.4)

In particular,
xG (t) = Q(t)X G + q(t), (3.5)

where xG (t) locates the centre of mass at time t. Subtracting this equation from (3.4), or using (3.3),
gives
x(t) = xG (t) + Q(t)(X − X G ). (3.6)

This expresses a well-known fact: the motion of a rigid body can be decomposed into a translation
of the centre of mass together with a rotation about the centre of mass.
The formalism adopted here is used in publications by Chadwick, Casey, O’Reilly and others;
for example, (3.4) is ([51], p. 52), ([50], eqn (2.1)) and ([4], eqn (7.4)); (3.5) is ([50], eqn (2.15)) and
([4], eqn (7.11)); and (3.6) is ([50], eqn (2.17)1 ) and ([4], eqn (7.12)).
For small oscillations (leading to a linear theory), we can approximate Q. Start by putting
Q = I + S, where S is another second-order tensor. We assume that S is small, and retain terms
that are linear in S; see ([2], §4–7). For example, enforcing QQT = QT Q = I to first order shows
6
that S is skew-symmetric, S = −ST . Therefore, in component form, we can write
⎛ ⎞ ⎛ ⎞

royalsocietypublishing.org/journal/rspa Proc. R. Soc. A 479: 20220463


−s3

..........................................................
S11 S12 S13 0 s2
⎜ ⎟ ⎜ ⎟
S = ⎝S21 S22 S23 ⎠ = ⎝ s3 0 −s1 ⎠ , (3.7)
S31 S32 S33 −s2 s1 0

where s1 , s2 and s3 are unknowns at this stage. They can be related to Eulerian angles. Thus, using
the 3-2-1 set of Euler angles defined by O’Reilly ([4], §6.8.1), we have s1 = φ, s2 = θ and s3 = ψ
in his notation, whereas comparison with ([45], eqn (1.39)) gives sj = −θj , j = 1, 2, 3, in Kausel’s
notation.
Downloaded from https://royalsocietypublishing.org/ on 04 March 2023

(b) Equations of motion


Consider a rigid body subjected to a force and a moment applied to its surface S. (Here, we ignore
body forces.) Denoting the applied force by F and time derivatives by overdots, the first equation
of motion is
F = mẍG = mS̈X G + mq̈,

using (3.5) and Q = I + S. Combining (3.7) with X G = xG e1 + yG e2 + zG e3 , we obtain


⎛ ⎞⎛ ⎞ ⎛ ⎞
0 −s̈3 s̈2 xG −s̈3 yG + s̈2 zG
⎜ ⎟⎜ ⎟ ⎜ ⎟
S̈X G = ⎝ s̈3 0 −s̈1 ⎠⎝yG ⎠ = ⎝ s̈3 xG − s̈1 zG ⎠
−s̈2 s̈1 0 zG −s̈2 xG + s̈1 yG
⎛ ⎞⎛ ⎞
0 zG −yG s̈1
⎜ ⎟⎜ ⎟
= ⎝−zG 0 xG ⎠⎝s̈2 ⎠ = s̈ × X G . (3.8)
yG −xG 0 s̈3

Thus the translational equation of motion of the centre of mass involves the vectors q̈ and s̈.
The second equation of motion governs the rotational motion about the centre of mass G. Casey
gives it in a convenient form, namely ([50], eqn (2.35)2 ),

2 M G = −ε M∧ , (3.9)

where M G is the applied moment about G, ε is the alternator, M∧ is a tensor defined by Casey
([50], eqn (2.44)),
T T
M∧ = Q̈EQT − QEQ̈ = S̈E(I + ST ) − (I + S)ES̈
T
 S̈E − ES̈  S̈E + ES̈, (3.10)

to first order, and E is the Euler tensor in the reference configuration, (3.2). The tensor E is related
to the inertia tensor relative to G in the reference configuration, JG , by (3.1). In detail, we have
G G G
J11 = E22 + E33 , J22 = E11 + E33 , J33 = E11 + E22 (3.11a)

and
G G G G G G
J12 = J21 = −E12 , J23 = J32 = −E23 , J13 = J31 = −E13 . (3.11b)

Returning to (3.10), noting that E is symmetric and S is skew-symmetric, it follows that M∧ is


1 ∧
also skew-symmetric. Then, using components in (3.9), MG i = − 2 εijk Mjk , we obtain

1 ∧ 1 ∧ ∧
MG
1 = − ε123 M23 − ε132 M32 = M32 ,
2 2
∧ ∧
2 = M13 and M3 = M21 . Hence, from (3.10) and (3.7),
MG G

MG
1 = S̈3j Ej2 + E3j S̈j2 = S̈31 E12 + S̈32 (E22 + E33 ) + E31 S̈12
G G G
= J11 s̈1 + J12 s̈2 + J13 s̈3
and
7
MG G G G
2 = J12 s̈1 + J22 s̈2 + J23 s̈3 , MG G G G
3 = J13 s̈1 + J23 s̈2 + J33 s̈3 .

royalsocietypublishing.org/journal/rspa Proc. R. Soc. A 479: 20220463


..........................................................
Collecting the two equations of motion, we have

F = ms̈ × X G + mq̈ and M G = JG s̈. (3.12)

The second of these can be solved for s̈ given M G , and then the first equation gives q̈.
For some purposes, it may be convenient to express both the applied moment and the inertia
tensor with respect to O instead of G; denote these quantities by M ◦ and J◦ , respectively. We have
(see [50], eqn (2.23)),

M ◦ −M G = xG × F  X G × (ms̈ × X G + mq̈)
Downloaded from https://royalsocietypublishing.org/ on 04 March 2023

= m|X G |2 s̈ − m(X G · s̈)X G + mX G × q̈

= m(|X G |2 I − X G ⊗ X G )s̈ + mX G × q̈,

where we have used (3.12)1 . For the inertia tensors, we have the parallel axis theorem (see [3], eqn
(10.147) and [4], eqn (7.25)), J◦ − JG = m(|X G |2 I − X G ⊗ X G ). Hence (3.12)2 becomes

M ◦ = J◦ s̈ + mX G × q̈. (3.13)

When combined with (3.12)1 , we have






F mI G q̈
= , (3.14)
M◦ GT J◦ s̈

where G is a 3 × 3 matrix,
⎛ ⎞
0 mzG −myG
⎜ ⎟
G = ⎝−mzG 0 mxG ⎠ . (3.15)
myG −mxG 0

Later, we denote the 6 × 6 matrix on the right-hand side of (3.14) by L; see (3.18).
One feature of the calculations above is that we did not introduce the angular velocity vector
(often denoted by ω(t)). One reason is we want to extract the displacement at each point in the
body, not the velocity of those points (although we shall require the normal velocity of S in §4).

(c) Time-harmonic motions


We are interested in small time-harmonic oscillations of the body B about its equilibrium position;
the time dependence is e−iωt . Put q(t) = Re{q̃ e−iωt } and s(t) = Re{s̃ e−iωt }, where q̃ and s̃ are
constant vectors. Henceforth, we suppress the time dependence.
Rigid-body oscillations of B couple to the exterior via the displacement of the boundary S. This
displacement is obtained by substituting Q = I + S in (3.4)1 , giving x − X = q + SX = q + s × X,
after use of (3.8). The quantity x − X on the left is the small displacement of the point at X with
position vector X in the reference configuration. In the time-harmonic setting, we denote this
displacement by
u(X) = q̃ + s̃ × X; (3.16)

its value at all points X ∈ S will be needed. As expected, there are six DOF in (3.16), the three
components of the two constant vectors, q̃ and s̃.
The motion is governed by (3.14), which becomes




F̃ mI G q̃
= −ω2 , (3.17)
M̃ GT J ◦ s̃

where F = Re{F̃ e−iωt } and M ◦ = Re{M̃ e−iωt }. Henceforth, we drop the tildes to simplify notation.
The 6 × 6 matrix on the right-hand side of (3.17) is denoted by L; explicitly, it is
8
⎛ ⎞
m 0 0 0 mzG −myG

royalsocietypublishing.org/journal/rspa Proc. R. Soc. A 479: 20220463


⎜ ⎟

..........................................................
⎜ 0 m 0 −mzG 0 mxG ⎟
⎜ −mxG ⎟
⎜ 0 0 m myG 0 ⎟
L=⎜ ◦ ◦ ◦ ⎟. (3.18)
⎜ 0 −mzG myG J11 J12 J13 ⎟
⎜ ◦ ◦ ◦

⎝ mzG 0 −mxG J12 J22 J23 ⎠
−myG mxG 0 ◦
J13 ◦
J23 ◦
J33

We note that L is symmetric and it does not depend on ω. It is essentially that given by Olsson
([34], eqn (31)). See also ([22], p. 156, eqn (141)) and ([52], eqn (3.48)).
Downloaded from https://royalsocietypublishing.org/ on 04 March 2023

(d) Summary
At this stage, we have obtained the time-harmonic equations of motion of the rigid body B in
terms of the forces and moments applied to its surface S, F and M. The body can move with six
DOF. Its displacement is given by (3.16), which we write as


6
u(x) = ξI bI (x) inside B, (3.19)
I=1

where ξi = qi , ξi+3 = si ,

bi = ei and bi+3 = ei × x; (3.20)

here, and in what follows, lower-case indices i, j, . . . take values 1, 2, 3, and upper-case indices
I, J, . . . take values 1, 2, . . . 6. The quantities ξI are constants to be determined by solving the
equations of motion (3.17).

4. Ship motions
A ship may be considered to be a rigid body floating in an ocean. (If deformations of the ship are
considered to be important, the theory described below can be extended by including vibrational
modes of the ship; this extension leads to what is known as ‘hydroelasticity theory’ [53].) Gravity
plays an essential role: a floating body is (usually) statically stable, so there will be a restoring
force when the body moves from its equilibrium position [54,55]. We are interested in deviations
from the static state caused by interactions with an incident surface gravity wave: the wave
will be scattered and the body will move. This is a classic problem in ship hydrodynamics. The
governing equations were derived by John ([56], §2) using a systematic perturbation procedure.
For a detailed exposition, see ([57], §7.2); see also [21], ([58], §19β) and ([52], ch. 3). For bodies that
are also translating with a constant mean velocity, see [53,59].
The theory starts by assuming that the ocean is an inviscid incompressible fluid, and that the
motion is irrotational. There will be surface gravity waves, and these will interact with the body.
The fluid motion can be determined by computing a (scalar) potential Φ(x, t), with ∇ 2 Φ = 0 in
the fluid and fluid velocity equal to grad Φ. The dynamic pressure is P(x, t) = −w ∂Φ/∂t, where
w is the fluid (water) density. The kinematic condition on the wetted surface of the body, S, is
that the normal velocity of the water ∂Φ/∂n should match the normal velocity of the body. If we
choose coordinates so that the mean free surface is at z = 0, with the z-axis pointing upwards,
the linearized free-surface boundary condition is ∂ 2 Φ/∂t2 + g ∂Φ/∂z = 0 at z = 0, where g is the
acceleration due to gravity. Finally, for simplicity, we consider deep water, so that there is no
motion as z → −∞.
incident
DIFFRACTION PROBLEM 9
waves

royalsocietypublishing.org/journal/rspa Proc. R. Soc. A 479: 20220463


..........................................................
outgoing fixed outgoing
waves waves

HEAVE PROBLEM

outgoing outgoing
waves waves
Downloaded from https://royalsocietypublishing.org/ on 04 March 2023

SWAY PROBLEM

outgoing outgoing
waves waves

ROLL PROBLEM

outgoing outgoing
waves waves

Figure 2. The solution of the two-dimensional problem is decomposed into four basic subproblems.

For time-harmonic motions, put

Φ = Re{φ(x) e−iωt } and P = Re{p(x) e−iωt }, (4.1)

where p = iωw φ. The displacement of B is given by (3.19), so the kinematic boundary condition
becomes
∂φ 6
= −iω ξI n · bI on S, (4.2)
∂n
I=1

where the unit normal vector n on S points into the water and the vectors bI are defined by (3.20).

(a) Decomposition into seven problems


There are seven unknowns, the amplitudes ξI in (3.19) and the velocity potential φ. To proceed,
we decompose φ into the sum of seven potentials,


6
φ = φd + ξI φ I (4.3)
I=1

where φd = φinc + φsc is known as the diffraction potential, φinc is the velocity potential of the
(given) incident wave, φsc represents scattered waves, ∂φd /∂n = 0 on S,
∂φ I
= −iω n · bI on S, I = 1, 2, . . . , 6, (4.4)
∂n
and φ I and φsc must satisfy the radiation condition (ensuring outgoing waves at infinity).
Thus the decomposition in (4.3) ensures that the kinematic condition, (4.2), is satisfied. For
two-dimensional problems, we decompose φ into the sum of four potentials, as shown in figure 2.
The diffraction potential φd solves the scattering problem, in which the incident wave interacts
10
with B when it is held fixed; the scattered wave is given by φsc , and it satisfies
∂φsc ∂φinc

royalsocietypublishing.org/journal/rspa Proc. R. Soc. A 479: 20220463


..........................................................
=− on S. (4.5)
∂n ∂n
The other six potentials φ I correspond to forced motion of B in each of the six modes characterized
by the vectors bI but in the absence of the incident wave. This kind of decomposition will be used
again later for other physical problems.
Each of the seven potentials, φ I and φsc , solves a boundary value problem with prescribed
normal derivative on S (Neumann problem); see (4.4) and (4.5). They can be computed by solving
the same boundary integral equation over S apart from different right-hand side functions ([60],
§6.6.3), [61]. If a ‘direct formulation’ ([60], §6.6.2) is used, the output will be the boundary values
Downloaded from https://royalsocietypublishing.org/ on 04 March 2023

of the potentials, and these are what will be needed in the next step: computing the forces and
moments exerted on the body so that we can solve the equations of motion for B to find the
amplitudes ξI in (4.3).
It is worth comparing with the simpler problem of potential flow generated by a rigid
body moving in an unbounded perfect fluid. For that problem, the corresponding potential is
decomposed into the sum of six potentials (as in (4.3) but without φd ), a decomposition that goes
back to Kirchhoff in 1869; see ([12], §118).

(b) Equations of motion


Denote the pressure corresponding to φd by pd = iωw φd . The corresponding (time-harmonic)
force and moment on the rigid body are
 
F d = − pd n dS and M d = − pd x × n dS. (4.6)
S S

We put the components of these two vectors into a six-vector D, with Di = Fdi and Di+3 = Mdi .
Next, consider the contributions from the potentials φ I , I = 1, 2, . . . , 6. Denote the pressure
corresponding to φ I by pI = iωw φ I . The corresponding forces and moments on B are
 
F I = − pI n dS and M I = − pI x × n dS.
S S
Their components are
 
FIi = − pI ni dS and MIi = FI,i+3 = − pI (x × n)i dS. (4.7)
S S
These define a 6 × 6 matrix F with complex frequency-dependent entries FIJ (ω). The real and
imaginary parts of F define added-mass and damping coefficients ([22], §6.17), and these can be
used to assess the contributions of each rigid-body mode to the response of the structure and to
the waves radiated by the structure. Note that F does not depend on the incident wave.
The matrix F is symmetric, FIJ = FJI . To see this, we use the reciprocity relation for two outgoing
potentials, φ1 and φ2 ,  
∂φ2 ∂φ1
φ1 − φ2 dS = 0. (4.8)
S ∂n ∂n
Start by taking φ1 = φ i and φ2 = φ j .


i ∂φ j ∂φ
j i
0 = w φ −φ dS = −iωw (φ i nj − φ j ni )dS = Fji − Fij
S ∂n ∂n S

using (4.4) and n · bi = ni . Next take φ1 = φ i and φ2 = φ j+3 and use


1 ∂φ j+3
= n · bj+3 = n · (ej × x) = ej · (x × n),
iω ∂n
to show that Fi,j+3 = Fj+3,i . Finally, take φ1 = φ i+3 and φ2 = φ j+3 to show that Fi+3,j+3 = Fj+3,i+3 .
Combining with the contribution from φd , the forces and moments acting on B have
11
components
6

royalsocietypublishing.org/journal/rspa Proc. R. Soc. A 479: 20220463


..........................................................
DI + FIJ ξJ , I = 1, 2, . . . , 6. (4.9)
J=1

These drive the motion of B, and so they appear on the left-hand side of (3.17). Hence


6
(FIJ (ω) + ω2 LIJ )ξJ = −DI , I = 1, 2, . . . , 6, (4.10)
J=1

which is a 6 × 6 system for ξI . The 6 × 6 symmetric matrix L with components LIJ is given
explicitly by (3.18).
Downloaded from https://royalsocietypublishing.org/ on 04 March 2023

(c) Haskind’s relation


The quantities DI on the right-hand side of (4.10) are components of F d and M d , and these are
defined in terms of certain integrals of the diffraction potential φd = φinc + φsc . It turns out that
these integrals can be evaluated without computing φd itself.
Let us start with the force F d , (4.6)1 . Its components are
  
∂φ i
Fdi = −iωw φd ni dS = −iωw φinc ni dS + w φsc dS, (4.11)
S S S ∂n
using (4.4). Now, using the reciprocity relation, (4.8), with φ1 = φsc and φ2 = φ i gives
  
∂φ i ∂φsc ∂φinc
φsc dS = φ i dS = − φ i dS,
S ∂n S ∂n S ∂n
where the boundary condition (4.5) has been used. Hence (4.11) becomes
 
∂φinc
Fdi = −w iωφinc ni + φ i dS. (4.12)
S ∂n
Similarly,
  
∂φ i+3
Mdi = −iωw φd ei · (x × n) dS = −iωw φinc n · bi+3 dS + w φsc dS
S S S ∂n
 
∂φinc
= −w iωφinc n · bi+3 + φ i+3 dS, (4.13)
S ∂n

using ei · (x × n) = n · (ei × x) = n · bi+3 . So, the six-vector D has components


 
I I ∂φinc
DI = −w iωφinc n · b + φ dS. (4.14)
S ∂n
This formula shows that the forces and moments induced by the diffraction potential φd can be
calculated without calculating φd itself. Using (4.4) again, we can write (4.14) as


∂φ I I ∂φinc
DI = w φinc −φ dS. (4.15)
S ∂n ∂n

Then, if φinc satisfies ∇ 2 φinc = 0 everywhere in the water, the integration over the wetted surface
S on the right-hand side of (4.15) can be replaced by an integration over a control surface S∞ in
the far field where far-field approximations to φ I can be used.
The formulae in (4.14) and (4.15), with integration over S or S∞ , are known as Haskind’s relation
[27]; the derivation given here follows one given by Newman [62]. See also ([21], eqn 31) and ([57],
§7.6.3). Chertock ([63], eqn (12)) found a similar formula in a slightly different context; see also
[64]. The decomposition in (4.3) is also attributed to Haskind; see ([21], p. 244), ([57], §7.3.1) and
([58], p. 566).
5. Acoustic scattering by a movable rigid body 12
The discussion in §4 extends readily to problems where a sound wave in an inviscid compressible

royalsocietypublishing.org/journal/rspa Proc. R. Soc. A 479: 20220463


..........................................................
fluid interacts with a movable rigid body; in fact, this extension was also made by Haskind [10].
For these acoustic problems, the velocity potential Φ(x, t) satisfies the wave equation so that the
corresponding potential φ(x) (see (4.1)) satisfies the Helmholtz equation (∇ 2 + k2 )φ = 0 in the
fluid, where k = ω/c and c is the speed of sound. The corresponding dynamic pressure is given by
p = iωw φ where w is the density of the fluid.
All the remaining calculations in §4 can be repeated without changes.
The relevant potentials φ I satisfy the Helmholtz equation outside S, the Sommerfeld radiation
condition, and a Neumann boundary condition on S, (4.4). Such problems are standard in
acoustics, and they can be solved using boundary integral equations ([60], ch. 5); software is
Downloaded from https://royalsocietypublishing.org/ on 04 March 2023

readily available [65].

(a) An example: plane-wave scattering by a sphere


Let us consider the scattering of a plane wave by a movable rigid homogeneous sphere. The
purpose is to show that the general formalism presented in §4 leads to known results for this very
simple problem.
The sphere B has radius a and is centred at the origin. The plane wave propagates in the z-
direction, so that


φinc (r, θ ) = P eikz = P in (2n + 1)jn (kr) Pn (cos θ),
n=0

where z = x3 = r cos θ , jn is a spherical Bessel function, Pn is a Legendre polynomial, P is a constant


(with the same dimensions as φinc ) and r and θ are spherical polar coordinates ([60], §4.6). For
φsc = φd − φinc , we write



φsc (r, θ ) = P in (2n + 1)cn hn (kr) Pn (cos θ),
n=0

(1)
where hn ≡ hn is a spherical Hankel function and cn is a constant; this representation ensures
that the Sommerfeld radiation condition is satisfied, recalling the time dependence of e−iωt . The
boundary condition, ∂φd /∂r = 0 at r = a, gives cn , and then, evaluating on the sphere,

∞ n+1
i (2n + 1)
φd (a, θ ) = P Pn (cos θ),
(ka)2 hn (ka)
n=0

having used a Wronskian relation from ([66], 10.50.1).


The only non-trivial component of the diffraction force is

Fd3 = D3 = −iωw φd (a, θ) cos θ dS
S

= −2π iωw a2 φd (a, θ)P1 (cos θ) sin θ dθ
0
Pw
= 4π iω ,
k2 h1 (ka)

using orthogonality of Legendre polynomials.


We also want φ 3 , a radiating solution of the Helmholtz equation satisfying ∂φ 3 /∂r = −iω cos θ
on r = a; see (4.4). Evidently, φ 3 (r, θ ) = Ah1 (kr) cos θ with A determined by the boundary condition.
Then, evaluating on r = a,
13
h1 (ka)
φ 3 (a, θ ) = −iω cos θ.
k h1 (ka)

royalsocietypublishing.org/journal/rspa Proc. R. Soc. A 479: 20220463


..........................................................
Hence, the only non-trivial entry in the matrix F is

ω2 h1 (ka)
F33 = −iωw φ 3 (a, θ ) cos θ dS = − mw ,
S ka h1 (ka)

where mw = (4/3)π a3 w is the fluid mass displaced by B.


Finally, the only non-trivial component of (4.10), I = 3, gives
 
h1 (ka) 4π Pw
m− mw ξ3 = . (5.1)
ka h1 (ka) iωk2 h1 (ka)
Downloaded from https://royalsocietypublishing.org/ on 04 March 2023

This agrees with ([7], eqn (6)). Some calculation, using h1 (x) = −eix (i + x)/x2 , gives

3(k/ω)P e−ika
ξ3 = − (5.2)
(2γ + 1)(1 − ika) − γ (ka)2
where γ = m/mw . This agrees with ([13], p. 101) and ([14], eqn (11)). In particular, in the long-wave
limit, ka → 0, (5.2) gives
3(k/ω)P
ξ3 ∼ − as ka → 0, (5.3)
2γ + 1
in agreement with Lamb ([12], §298, eqn (9)).
As a check, suppose  = w (γ = 1) in (5.3), giving ξ3 ∼ −(k/ω)P as ka → 0. In this limit, we
should recover the incident wave. Its velocity is grad φinc = ikP eikz e3 , with z-component ikP at
the origin. On the other hand, from (3.19), this same component is given by −iωξ3 . Comparing
these two expressions gives ξ3 = −(k/ω)P, in agreement with the asymptotic approximation.
It is worth noting that the sphere problem is not typical, in the following sense: the sphere can
rotate in any manner without affecting the fluid motion. This non-uniqueness was avoided above
by restricting to one-dimensional motion from the outset. Similarly, for a rigid spheroid (or any
other axisymmetric body), rotations about its axis do not affect the fluid motion.

6. Soil–structure interaction
In this section, we consider scattering of an incident time-harmonic elastic wave Re{uinc e−iωt } by
a movable rigid body B. The problem is to compute the scattered waves and the motion of B. As
in §4, this scattering problem will be decomposed into seven (or more) simpler problems. This
can be done in several different ways, as we shall see.
We start by assuming that B is embedded in an unbounded solid. This implies that the scattered
waves must satisfy an elastodynamic form of the Sommerfeld radiation condition.
In the SSI context, it is usual to embed B in a half-space, with a flat traction-free boundary. (For
two-dimensional problems, see figure 2, where the incident waves shown could be body waves
or surface waves.) We discuss this situation later, in §6c.

(a) Thau’s decomposition


In our opinion, the best decomposition mimics what is done for ship motions (§4). However, it is
not the one used in most of the SSI literature; for an exception, see [67].
We start by defining the diffraction field ud = uinc + usc , where ud = 0 on S. Thus usc satisfies
a radiation condition and the boundary condition usc = −uinc on S. Denote the corresponding
traction vector on S by td .
The other six problems are associated with the six DOF in the motion of the rigid body; its
displacement is given by (3.19), involving the six unknowns ξI , I = 1, 2, . . . , 6. Thus, solve six
exterior problems for uI with boundary condition uI = bI on S (with bI defined by (3.20)), together
with a radiation condition; there is no incident wave in these problems. Denote the corresponding
14
traction vectors on S by tI .
The seven problems for usc and uI are standard exterior Dirichlet problems in elastodynamics.

royalsocietypublishing.org/journal/rspa Proc. R. Soc. A 479: 20220463


..........................................................
They can be solved using the same boundary integral equation over S with different right-hand
sides. If a direct formulation is used, the output will be the traction vectors, and (most of) these
will be needed later. For some details, see ([60], §6.5) and [49].
Then, we write the displacement as


6
u(x) = ud (x) + ξI uI (x) outside B. (6.1)
I=1

This ensures continuity of u across S; see (3.19), recalling that ud = 0 and uI = bI on S.


Downloaded from https://royalsocietypublishing.org/ on 04 March 2023

The particular decomposition of u into the solution of seven uncoupled problems as (6.1) was
introduced by Thau [26] in 1967. Apparently, the advantages of this decomposition were not
noticed at the time.

(i) Equations of motion


In order to determine ξI in (6.1), we calculate the (time-harmonic) forces and moments on the
body. From the diffraction field ud , we have
 
F d = td dS and M d = x × td dS. (6.2)
S S

We put the components of these two vectors into a six-vector D, with Di = Fdi and Di+3 = Mdi .
From uI , the corresponding forces and moments are
 
F I = tI dS and M I = x × tI dS. (6.3)
S S
Their components are
 
FIi = tIi dS and MIi = FI,i+3 = (x × tI )i dS. (6.4)
S S
These define a 6 × 6 matrix F with complex frequency-dependent entries FIJ (ω). As noted below
(4.7), the real and imaginary parts of F can be used to assess the contributions of each rigid-body
mode to the response of the structure and to the waves radiated by the structure. Note again that
F does not depend on the incident wave.
The matrix F is symmetric, FIJ = FJI . To see this, we use the reciprocal theorem, relating two
outgoing elastodynamic states, 
J J
(uIi ti − ui tIi ) dS = 0, (6.5)
S

with the usual summation convention. Start by taking I = j and J = k:


 
j j j
0 = (ui tki − uki ti ) dS = (δij tki − δik ti ) dS = Fkj − Fjk ,
S S
j J j
using ui = δij on S. Next, take I = j and J = k + 3: on S, we have uIi ti = ui tk+3
i = tk+3
j and
J
ui tIi = uk+3
i tIi = (ek × x)i tIi = εipq δpk xq tIi = εikq xq tIi = εkqi xq tIi = (x × tI )k .

Hence reciprocity gives



0= (tk+3
j − (x × tj )k ) dS = Fk+3,j − Fj,k+3 .
S
Finally, take I = j + 3 and J = k + 3: on S, we have

0 = ((x × tk+3 )j − (x × tj+3 )k ) dS = Fk+3,j+3 − Fj+3,k+3 .
S
We conclude that FIJ = FJI .
15
Combining (6.2) and (6.3), we find the forces and moments acting on B have components given
by (4.9), leading to the 6 × 6 system for ξI , (4.10).

royalsocietypublishing.org/journal/rspa Proc. R. Soc. A 479: 20220463


..........................................................
(ii) Haskind-type relations
At this stage, the problem has been solved in terms of seven elastodynamic vectors ud and uI ,
I = 1, 2, . . . , 6, and these can be computed by solving appropriate boundary integral equations. It
is not surprising, perhaps, that we can avoid computing ud , much as we avoided computing φd
in §4 because of Haskind’s relation.
Let tinc denote the tractions on S corresponding to uinc with components tinc i and uinc
i ,
sc sc
respectively. Define tsc , ti and ui similarly. Then, as td = tinc + tsc , (6.2)1 gives
Downloaded from https://royalsocietypublishing.org/ on 04 March 2023

 
Fdj = tinc
j dS + tsc
j dS. (6.6)
S S
j
Using the reciprocity relation, (6.5), with I = j and J = sc, together with ui = δij and usc
i = −ui
inc

on S, we can rewrite the second integral in (6.6), giving



Fdi = (tinc i
i − uinc · t )dS. (6.7)
S
For the moment, (6.2)2 gives
 
Mdj = ej · (x × tinc ) dS + ej · (x × tsc ) dS. (6.8)
S S
Using the reciprocity relation with I = j + 3 and J = sc,
  
j+3 sc j+3 j+3
ui tsci dS = u t
i i dS = − uinc
i ti dS.
S S S
The integrand on the left-hand side is (ej × x) · tsc = ej · (x × tsc ), as in the second integral in (6.8).
Hence 
Mdi = (tinc · (ei × x) − uinc · ti+3 ) dS. (6.9)
S

Putting the components of F d and M d into a six-vector D, we obtain



DI = (tinc · bI − uinc · tI ) dS (6.10a)
S

= (tinc · uI − uinc · tI ) dS. (6.10b)
S
These are used on the right-hand side of the equations of motion, (4.10).
We can regard (6.10) as the elastodynamic versions of Haskind’s relation, as derived in §4c for
ship motions. We note that, in (6.10b), the integration surface can be moved to infinity if uinc is a
valid elastodynamic field everywhere outside S.
Let us summarize. As a consequence of (6.10), we have to solve six exterior problems with
uI = bI on S and then compute the corresponding tractions tI on S, I = 1, 2, . . . , 6. The motion of B
is then determined by solving a 6 × 6 system of linear algebraic equations. We shall compare this
workload with alternative formulations below.

(b) Alternative decompositions


(i) Replace the rigid body by a cavity
Instead of constructing ud with ud = 0 on S (thus replacing B by a fixed rigid body), suppose we
introduce u∗d defined by u∗d = uinc + u∗sc , where u∗sc satisfies a radiation condition and t∗d = 0 on S
(thus replacing B by an empty cavity). This field exerts no forces on B but the displacement u∗d on
S must be calculated and taken into account.
We write the displacement inside B using (3.19). Outside B, we write (cf. (6.1))
16

6
u(x) = u∗d (x) + ur (x) + ξI uI (x), (6.11)

royalsocietypublishing.org/journal/rspa Proc. R. Soc. A 479: 20220463


..........................................................
I=1

where ur is a radiating field that satisfies ur = −u∗d on S. This ensures continuity of u across S. We
can then proceed as before. Note that we have to solve an additional radiation problem (to find ur ),
and we have to calculate the forces and moments associated with ur . Also, the construction of u∗d
requires solving a problem with traction (Neumann) data (t∗sc = −tinc on S); all other problems
have displacement (Dirichlet) data. Thus, we conclude that this approach is not attractive.

(ii) Luco’s decomposition


Downloaded from https://royalsocietypublishing.org/ on 04 March 2023

Another option is to introduce an exterior problem that has zero forces and moments on S, but not
zero tractions. Here, we use notation similar to that used by Luco [68]. He splits the displacement
inside B, writing u = uR ∗
o + uo ([68], eqn (3c)); each of these can be written as in (3.19), using

coefficients ξI and ξI , respectively. A radiating exterior field uR is constructed, with uR = uR
R
o on S
([68], eqn (4a)); evidently,

6
uR (x) = ξIR uI (x) outside B,
I=1

where the coefficients ξIR are to be found. A second radiating exterior field us is constructed; it
takes account of the incident field uinc , and satisfies us = u∗o − uinc on S, ([68], eqn (5a)). Thus


6
us (x) = usc (x) + ξI∗ uI (x) outside B,
I=1

where usc was defined above: usc = ud − uinc with ud = 0 on S. In addition, the forces and
moments induced by us are required to cancel those induced by the incident field ([68], eqn (5c)).
Enforcing this condition, using the notation introduced earlier (see (6.2) and (6.3)), leads to


6
FIJ ξJ∗ = −DI , I = 1, 2, . . . , 6.
J=1

Solving this 6 × 6 system gives ξI∗ and hence u∗o .


Next, write the exterior field as in ([68], eqn (3a)),

u(x) = uR (x) + us (x) + uinc (x) outside B.

This matches with the interior field on S. The associated forces and moments come from uR alone
(because of the condition imposed on us ). These drive the motion of B. Then, proceeding as before,
we can obtain equations for the remaining coefficients ξIR .
This approach seems much more cumbersome than Thau’s approach because one needs to
compute nine exterior fields and two sets of coefficients, ξI∗ and ξIR .

(iii) The ‘substructure theorem’


A standard approach for solving elastodynamic inclusion (transmission) problems proceeds as
follows. Let uS denote the (unknown) total displacement of S. Solve the exterior problem, taking
account of the incident field and the radiation condition, and compute the corresponding traction
on S, text ; formally, we can write text = Zext uS , for some operator Zext , where the subscript ext
denotes exterior. (The Dirichlet-to-Neumann operator Zext could be built using boundary integral
equations, for example.) If B is an elastic inclusion, we can do exactly the same in B, giving
tint = Zint uS , where the subscript int denotes interior. Then, if the interface S is ‘welded’, we have
text = tint on S; enforcing this gives an equation for uS .
For the exterior part of this calculation, let u be the total displacement outside S and write
17
u = uinc + u◦sc where u◦sc = u − uinc is an outgoing field. Then text = tinc + Z u◦sc , where the operator
Z incorporates the radiation condition but does not depend on uinc . Hence

royalsocietypublishing.org/journal/rspa Proc. R. Soc. A 479: 20220463


..........................................................
(Z − Zint ) uS = Z uinc − tinc , (6.12)

an equation for uS that holds on S; it is the continuous analogue of the ‘substructure theorem’ in
Kausel’s book ([45], eqn (7.150)), where the focus is on discrete formulations.
Let us simplify the right-hand side of (6.12), noting first that tinc = Z uinc ; for the latter, we
want the outgoing field uout out
inc , say, satisfying uinc = uinc on S, and then Z uinc gives the tractions on
S corresponding to uinc . However, comparison with §6a shows that uout
out
inc = −usc whence Z uinc −
tinc = −td .
Downloaded from https://royalsocietypublishing.org/ on 04 March 2023

Suppose now that the elastic inclusion is replaced by a rigid body B: how does this affect (6.12)
and the interpretation of Zint ? The ‘welded’ condition, text = tint on S is replaced by conditions
involving integrals of text over S (giving the forces and moments acting on B) and the equations
of motion of B. In addition, we know that we can write uS as (3.19), whence (6.12) becomes


6
ξJ {tJ (x) − (Zint bJ )(x)} = −td (x), x ∈ S, (6.13)
J=1

where the six traction vectors tJ are defined by solving six exterior problems, as in §6a. equation
(6.13) must be projected onto a six-dimensional space in order to determine the six coefficients
ξI ; all the other formulations in §6 do not require projections because they deal directly with
forces and moments. Denote the projection operator by P; given a traction vector t(x) defined for
x ∈ S, define P t to be the six-vector containing the forces and moments acting on B. In particular,
(P td )I = DI and (P tJ )I = FIJ ; see (6.2) and (6.3). Hence, applying P to (6.13) gives


6
{FIJ − (PZint bJ )I }ξJ = −DI , I = 1, 2, . . . , 6. (6.14)
J=1

Comparison with (4.10) gives (PZint bJ )I = −ω2 LIJ , which may be used to interpret Zint , if such an
interpretation is desired.
We conclude that the apparent simplicity of (6.12) is not easy to exploit when applied to wave-
induced motions of rigid bodies. Nevertheless, discrete formulations based on (6.12) are popular,
perhaps because of the ubiquity of finite-element methods.

(c) Half-space problems


Consider a homogeneous elastic half-space, z > 0, with a traction-free boundary at z = 0. We
suppose that there is a plane wave propagating towards the flat boundary; this is the incident
field uinc . There will be reflected plane waves, both P-waves and S-waves in general. These are
readily calculated in the absence of the body ([69], ch. 6), and then the sum of the incident and
reflected plane waves is known as the free-field displacement uff ; in the context of SSI, uff gives
‘the motion that the ground would have experienced if neither the soil had been excavated nor
the structure erected’ ([45], p. 514).
For the analogue of Thau’s decomposition (§6a), we replace uinc by uff , defining the diffraction
field by ud = uff + usc with ud = 0 on S. The field usc is required to be outgoing and to satisfy
the zero-tractions condition on z = 0 (outside B if the body intersects z = 0); these conditions are
also imposed on the fields uI . Similar alterations are required in the alternative decompositions
outlined in §6b.
Construction of the fields uI is more complicated than for full-space problems because they
have to satisfy the zero-traction condition on the flat boundary. Appropriate boundary integral
equations can be derived, using a ‘half-space Green function’ (one that satisfies the zero-traction
boundary condition at z = 0), but there are alternatives; for references and discussion, see ([60],
18
§6.5.5) and [70,71].

royalsocietypublishing.org/journal/rspa Proc. R. Soc. A 479: 20220463


..........................................................
7. Conclusion
We reviewed several physical situations where a three-dimensional rigid body interacts with a
time-harmonic incident wave. After reviewing the classical theory of rigid-body motions, we
discussed problems involving floating bodies (ships) and other offshore structures: a surface
gravity wave is incident on the structure, causing it to move and affecting how the incident wave
is scattered. These problems have a rich literature, dating back to the 1940s. They require the
determination of a scalar potential φ, and the method of choice is to decompose φ into seven
potentials, one for each of the six rigid-body modes (three translations and three rotations), and
Downloaded from https://royalsocietypublishing.org/ on 04 March 2023

one to take account of the incident wave. The six potentials corresponding to the rigid-body
modes are independent; they depend on the geometry and the frequency, but they do not involve
the incident wave. The contribution from the seventh potential can be computed in terms of the
other six, using Haskind’s relation. Exactly the same approach can be used for acoustic problems.
The same basic approach can be used for elastodynamic problems, where an incident elastic
wave interacts with a rigid body. For full-space problems, Thau’s decomposition [26] into
seven subproblems is straightforward: of course, they are vector problems but they can be
solved numerically using standard methods; and a form of Haskind’s relation in available. The
decomposition into subproblems is advantageous because it breaks the original problem into
seven problems of the same basic type (exterior problems with different specified displacements
of the body’s surface), allowing the contributions from the incident waves and the rigid-body
modes to be separated: there is also a close analogy with the standard introduction of added-mass
and damping matrices in hydrodynamics.
The extension to half-space problems (as usually encountered in problems of SSI) is also
straightforward, but now it is not so easy to solve the six subproblems because the traction-free
condition on the flat boundary of the half-space has to be imposed. Nevertheless, this approach
does provide a rigorous way to analyse such problems, and we believe that further investigations
are warranted.
To conclude, we should not lose sight of the underlying assumptions behind the analysis. First,
we used linear theory: the motions are assumed to be small. This is standard in many situations
involving wave-induced motions and, moreover, the resulting predictions are often better than
one might expect. (It is difficult to quantify how small the motions should be.) Large-amplitude
motions are certainly of interest (especially those of floating bodies), but nonlinear theories
are expensive to deploy because they inevitably require solving partial differential equations
numerically in an unknown moving domain coupled to the motion of the rigid body. We assumed
time-harmonic motions; motions that are truly time dependent can be studied, in principle, by
Fourier analysis, or by other methods [16]. We also assumed that we are dealing with rigid bodies.
This could be relaxed (e.g. buildings can deform), but then the six rigid-body modes would have
to be augmented with the vibrational modes of the body [53].
Data accessibility. This article has no additional data.
Authors’ contributions. P.A.M.: conceptualization, formal analysis, investigation, methodology, validation,
writing—original draft, writing—review and editing.
Conflict of interest declaration. I declare I have no competing interests.
Funding. No funding has been received for this article.
Acknowledgements. The author thanks Harry Dankowicz, Chris Garrett and three anonymous referees for
constructive comments on earlier drafts of the paper.

References
1. Whittaker ET. 1937 A treatise on the analytical dynamics of particles and rigid bodies, 4th edn.
Cambridge, UK: Cambridge University Press.
2. Goldstein H. 1950 Classical mechanics. Reading, PA: Addison-Wesley.
19
3. Symon KR. 1971 Mechanics, 3rd edn. Reading, PA: Addison-Wesley.
4. O’Reilly OM. 2008 Intermediate dynamics for engineers: a unified treatment of Newton–Euler and

royalsocietypublishing.org/journal/rspa Proc. R. Soc. A 479: 20220463


..........................................................
Lagrangian mechanics. Cambridge, UK: Cambridge University Press.
5. Rudinger G. 1973 Wave propagation in suspensions of solid particles in gas flow. Appl. Mech.
Rev. 26, 273–279.
6. Imbert D, McNamara S, Le Gonidec Y. 2015 Fictitious domain method for acoustic waves
through a granular suspension of movable rigid spheres. J. Comp. Phys. 280, 676–691.
(doi:10.1016/j.jcp.2014.10.006)
7. Hickling R, Wang NM. 1966 Scattering of sound by a rigid movable sphere. J. Acoust. Soc. Am.
39, 276–279. (doi:10.1121/1.1909887)
8. Krysl P, Hawkins AD, Schilt C, Cranford TW. 2012 Angular oscillation of solid scatterers in
response to progressive planar acoustic waves: do fish otoliths rock? PLoS ONE 7, e42591.
Downloaded from https://royalsocietypublishing.org/ on 04 March 2023

(doi:10.1371/journal.pone.0042591)
9. Popper AN, Hawkins AD. 2018 The importance of particle motion to fishes and invertebrates.
J. Acoust. Soc. Am. 143, 470–488. (doi:10.1121/1.5021594)
10. Khaskind MD. 1957 Diffraction and radiation of acoustic waves in fluids and gases. Part I.
Sov. Phys. Acoust. 3, 371–384.
11. Olsson P. 1985 Acoustic scattering by a rigid movable body immersed in a fluid. J. Acoust. Soc.
Am. 78, 2132–2138. (doi:10.1121/1.392673)
12. Lamb H. 1932 Hydrodynamics, 6th edn. Cambridge, UK: Cambridge University Press.
13. Wolf A. 1945 Motion of a rigid sphere in an acoustic wave field. Geophys. 10, 91–109.
(doi:10.1190/1.1437153)
14. Tang S. 1968 Transient response of a rigid sphere to a plane acoustic step wave. J. Acoust. Soc.
Am. 44, 289–291. (doi:10.1121/1.1911071)
15. Temkin S, Leung CM. 1976 On the velocity of a rigid sphere in a sound wave. J. Sound Vib. 49,
75–92. (doi:10.1016/0022-460X(76)90758-6)
16. Martin PA. 2021 Time-domain scattering. Cambridge, UK: Cambridge University Press.
17. Novozhilov VV. 1959 On the displacement of an absolutely rigid body under the action
of an acoustic pressure wave. J. Appl. Math. Mech. 23, 1138–1142. (doi:10.1016/0021-
8928(59)90050-4)
18. Grigolyuk ÉI, Gorshkov AG. 1968 Motion of a rigid sphere under the influence of an acoustic
pressure wave. Sov. Phys. Dokl. 12, 1009–1011.
19. Temkin S. 1972 On the response of a sphere to an acoustic pulse. J. Fluid Mech. 54, 339–349.
(doi:10.1017/S0022112072000710)
20. Gorshkov AG, Tarlakovsky DV. 2001 Transient aerohydroelasticity of spherical bodies. Berlin,
Germany: Springer.
21. Wehausen JV. 1971 The motion of floating bodies. Annu. Rev. Fluid Mech. 3, 237–268.
(doi:10.1146/annurev.fl.03.010171.001321)
22. Newman JN. 2017 Marine hydrodynamics, 40th anniversary edn. Cambridge, MA: MIT Press.
23. Jonkman JM. 2009 Dynamics of offshore floating wind turbines—model development and
verification. Wind Energy 12, 459–492. (doi:10.1002/we.347)
24. Borg M, Collu M. 2015 Offshore floating vertical axis wind turbines, dynamics modelling state
of the art. Part III: hydrodynamics and coupled modelling approaches. Renew. Sustain. Energy
Rev. 46, 296–310. (doi:10.1016/j.rser.2014.10.100)
25. Borg M, Collu M. 2015 A comparison between the dynamics of horizontal and vertical
axis offshore floating wind turbines. Phil. Trans. R. Soc. A 373, 20140076 (doi:10.1098/
rsta.2014.0076)
26. Thau SA. 1967 Radiation and scattering from a rigid inclusion in an elastic medium. J. Appl.
Mech. 34, 509–511. (doi:10.1115/1.3607720)
27. Haskind (Khaskind) MD. 1957 The exciting forces and wetting of ships in waves. Izvestia
Akademii Nauk SSSR, Otdelenie Tekhnicheskikh Nauk, no. 7, 65–79. Translated from the Russian
as David Taylor Model Basin Translation T-307, November 1962.
28. Lamb H. 1900 Problems relating to the impact of waves on a spherical obstacle in an elastic
medium. Proc. Lond. Math. Soc. 32, 120–150. (doi:10.1112/plms/s1-32.1.120)
29. Pao YH, Mow CC. 1963 Scattering of plane compressional waves by a spherical obstacle. J.
Appl. Phys. 34, 493–499. (doi:10.1063/1.1729301)
30. Iwashimizu Y. 1972 Scattering of elastic waves by a movable rigid sphere embedded in an
infinite elastic solid. J. Sound Vib. 21, 463–469. (doi:10.1016/0022-460X(72)90830-9)
31. Pao YH, Mow CC. 1973 Diffraction of elastic waves and dynamic stress concentrations. New York,
20
NY: Crane Russak.
32. Oien MA, Pao YH. 1973 Scattering of compressional waves by a rigid spheroidal inclusion. J.

royalsocietypublishing.org/journal/rspa Proc. R. Soc. A 479: 20220463


..........................................................
Appl. Mech. 40, 1073–1077. (doi:10.1115/1.3423128)
33. Datta SK, Sangster JD. 1974 Response of a rigid spheroidal inclusion to an incident plane
compressional elastic wave. SIAM J. Appl. Math. 26, 350–369. (doi:10.1137/0126033)
34. Olsson P. 1985 The rigid movable inclusion in elastostatics and elastodynamics. Wave Motion
7, 421–445. (doi:10.1016/0165-2125(85)90018-6)
35. Olsson P. 1986 Scattering of elastic waves by a smooth rigid movable inclusion. J. Acoust. Soc.
Am. 79, 1237–1247. (doi:10.1121/1.393703)
36. Datta SK. 1970 The diffraction of a plane compressional elastic wave by a rigid circular disc.
Q. Appl. Math. 28, 1–14. (doi:10.1090/qam/99806)
37. Mal AK. 1971 Motion of a rigid disc in an elastic solid. Bull. Seism. Soc. Am. 61, 1717–1729.
Downloaded from https://royalsocietypublishing.org/ on 04 March 2023

(doi:10.1785/BSSA0610061717)
38. Cai LW, Hambric SA. 2016 Movable rigid scatterer model for flexural wave scattering on thin
plates. J. Vib. Acoust. 138, 031016. (doi:10.1115/1.4033060)
39. Luco JE. 1982 Linear soil–structure interaction: a review. In Earthquake ground motion and its
effects on structures (ed. SK Datta), pp. 41–57. New York, NY: ASME.
40. Karabalis DL, Beskos DE. 1987 Dynamic soil–structure interaction. In Boundary element methods
in mechanics (ed. DE Beskos), pp. 499–562. Amsterdam, The Netherlands: North-Holland.
41. Kausel E. 2010 Early history of soil–structure interaction. Soil Dyn. Earthq. Eng. 30, 822–832.
(doi:10.1016/j.soildyn.2009.11.001)
42. Roesset JM. 2013 Soil structure interaction: the early stages. J. Appl. Sci. Eng. 16, 1–8.
(doi:10.6180/jase.2013.16.1.01)
43. Anand V, Satish Kumar SR. 2018 Seismic soil-structure interaction: a state-of-the-art review.
Structures 16, 317–326. (doi:10.1016/j.istruc.2018.10.009)
44. Beskos DE. 1997 Boundary element methods in dynamic analysis: part II (1986–1996). Appl.
Mech. Rev. 50, 149–197. (doi:10.1115/1.3101695)
45. Kausel E. 2017 Advanced structural dynamics. Cambridge, UK: Cambridge University Press.
46. Richart FE, Hall JR, Woods RD. 1970 Vibration of soils and foundations. Englewood Cliffs, NJ:
Prentice-Hall.
47. Triantafyllidis T, Prange B. 1988 Rigid circular foundation: dynamic effects of coupling to the
half-space. Soil Dyn. Earthq. Eng. 7, 40–52. (doi:10.1016/S0267-7261(88)80014-9)
48. Clouteau D, Cottereau R, Lombaert G. 2013 Dynamics of structures coupled with elastic
media—a review of numerical models and methods. J. Sound Vib. 332, 2415–2436.
(doi:10.1016/j.jsv.2012.10.011)
49. Bouchon M, Sánchez-Sesma FJ. 2007 Boundary integral equations and boundary elements
methods in elastodynamics. Adv. Geophys. 48, 157–189. (doi:10.1016/S0065-2687(06)48003-1)
50. Casey J. 1995 On the advantages of a geometrical viewpoint in the derivation of Lagrange’s
equations for a rigid continuum. Zeit. Angew. Math. Phys. 46, Special Issue, S805–S847.
(doi:10.1007/978-3-0348-9229-2_41)
51. Chadwick P. 1976 Continuum mechanics. London, UK: Allen & Unwin.
52. Faltinsen OM. 1990 Sea loads on ships and offshore structures. Cambridge, UK: Cambridge
University Press.
53. Bishop RED, Price WG, Wu Y. 1986 A general linear hydroelasticity theory of floating
structures moving in a seaway. Phil. Trans. R. Soc. A 316, 375–426. (doi:10.1098/rsta.1986.0016)
54. Moseley H. 1850 On the dynamical stability and on the oscillations of floating bodies. Phil.
Trans. R. Soc. 140, 609–643. (doi:10.1098/rstl.1850.0031)
55. Spyrou KJ. 2022 The stability of floating regular solids. Ocean Eng. 257, 111615.
(doi:10.1016/j.oceaneng.2022.111615)
56. John F. 1949 On the motion of floating bodies. I. Commun. Pure Appl. Math. 2, 13–57.
(doi:10.1002/cpa.3160020102)
57. Mei CC. 1989 The applied dynamics of ocean surface waves, 2nd printing with corrections.
Singapore: World Scientific.
58. Wehausen JV, Laitone EV. 1960 Surface waves. In Encyclopedia of physics (eds S Flügge, C
Truesdell), vol. IX, pp. 446–778. Berlin, Germany: Springer. (doi:10.1007/978-3-642-45944-3_6).
59. Newman JN. 1979 The theory of ship motions. Adv. Appl. Mech. 18, 221–283. (doi:10.1016/
S0065-2156(08)70268-0)
60. Martin PA. 2006 Multiple scattering. Cambridge, UK: Cambridge University Press.
21
61. Sheng W, Tapoglou E, Ma X, Parsons DR, Aggidis G. 2022 Hydrodynamic studies of floating
structures: comparison of wave-structure interaction modelling. Ocean Eng. 249, 110878.

royalsocietypublishing.org/journal/rspa Proc. R. Soc. A 479: 20220463


..........................................................
(doi:10.1016/j.oceaneng.2022.110878)
62. Newman JN. 1962 The exciting forces on fixed bodies in waves. J. Ship Res. 6, no. 4, 10–17.
(doi:10.5957/jsr.1962.6.4.10)
63. Chertock G. 1953 The flexural response of a submerged solid to a pulsating gas bubble. J. Appl.
Phys. 24, 192–197. (doi:10.1063/1.1721237)
64. Chertock G. 1962 General reciprocity relation. J. Acoust. Soc. Am. 34, 989. (doi:10.1121/1.
1918239)
65. Śmigaj W, Betcke T, Arridge S, Phillips J, Schweiger M. 2015 Solving boundary integral
problems with BEM++. ACM Trans. Math. Softw. 41, article 6. (doi:10.1145/2590830)
66. NIST Digital Library of Mathematical Functions. (http://dlmf.nist.gov/).
Downloaded from https://royalsocietypublishing.org/ on 04 March 2023

67. Suárez M, Avilés J, Sánchez-Sesma FJ. 2002 Response of L-shaped rigid foundations
embedded in a uniform half-space to traveling seismic waves. Soil Dyn. Earthq. Eng. 22,
625–637. (doi:10.1016/S0267-7261(02)00058-1)
68. Luco JE. 1986 On the relation between radiation and scattering problems for foundations
embedded in an elastic half-space. Soil Dyn. Earthq. Eng. 5, 97–101. (doi:10.1016/0267-
7261(86)90003-5)
69. Graff KF. 1991 Wave motion in elastic solids. New York, NY: Dover.
70. Chaillat S, Bonnet M. 2013 Recent advances on the fast multipole accelerated boundary
element method for 3D time-harmonic elastodynamics. Wave Motion 50, 1090–1104.
(doi:10.1016/j.wavemoti.2013.03.008)
71. Bruno OP, Yin T. 2021 A windowed Green function method for elastic scattering problems on
a half-space. Comput. Methods Appl. Mech. Eng. 376, 113651. (doi:10.1016/j.cma.2020.113651)

You might also like