0% found this document useful (0 votes)
53 views14 pages

A Complex Variable Method For The Floating-Body Boundary-Value Problem

This document summarizes a paper that reviews a complex variable method for solving the floating-body boundary-value problem in fluid dynamics. The method uses Cauchy's theorem to formulate the problem as a boundary-value problem on the free surface boundary rather than solving over the entire fluid domain. This allows for efficient numerical schemes to study problems involving nonlinear water waves and floating or submerged bodies. The paper outlines outstanding challenges with the method and provides examples of its application to problems like water exiting a wedge and horizontal motion of a submerged cylinder.
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
53 views14 pages

A Complex Variable Method For The Floating-Body Boundary-Value Problem

This document summarizes a paper that reviews a complex variable method for solving the floating-body boundary-value problem in fluid dynamics. The method uses Cauchy's theorem to formulate the problem as a boundary-value problem on the free surface boundary rather than solving over the entire fluid domain. This allows for efficient numerical schemes to study problems involving nonlinear water waves and floating or submerged bodies. The paper outlines outstanding challenges with the method and provides examples of its application to problems like water exiting a wedge and horizontal motion of a submerged cylinder.
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 14

Journal of Computational and Applied Mathematics 46 (1993) 115-128 115

North-Holland

CAM 1331

A complex variable method for the


floating-body boundary-value problem
Martin Greenhow
Department of Mathematics and Statistics, Brunel University, Uxbridge, United Kingdom

Received 24 October 1991


Revised 23 March 1992

Abstract

Greenhow, M., A complex variable method for the floating-body boundary-value problem, Journal of
Computational and Applied Mathematics 46 (1993) 115-128.

This paper reviews the use of a method using Cauchy’s theorem, which has recently formed the basis of very
successful numerical schemes for considering the nonlinear water-wave problem, both with and without
floating or submerged bodies. Outstanding problems, concerning the intersection point problem (where two
boundary conditions are to be applied) and the treatment of radiation conditions, are outlined. Notwithstand-
ing these difficulties, useful results for transient problems may be obtained; examples presented include new
results for water exit of a wedge and horizontal motion of a submerged cylinder.

Keywords: Floating body; free-surface flows; water exit; horizontal motion of submerged cylinder.

1. Introduction

The development of complex variable methods is intimately connected with that of inviscid
fluid dynamics, normally used in the study of water waves, and their interaction with floating
bodies. Under the usual assumptions of irrotationality, we can define a velocity potential @
with the fluid velocity vector v = (u, u, W) = V@. If the fluid is also incompressible, this implies
the field equation for @ is Laplace’s equation V2@ = 0. An alternative viewpoint, valid for
two-dimensional flow Y = (u, u) in the v-plane, is to introduce the stream function !P with the
property that
a@ a!P a@
u=-=- ,!=-=-a”y
(1)
ax aY ’ aY ax *

These are the Cauchy-Riemann equations for the real and imaginary parts of the complex
potential p(z, t) = @(x, y, t) + iP(x, y, t). Here z =x + iy and t is time. Consequently, any

Correspondence to: Dr. M. Greenhow, Department of Mathematics and Statistics, Brunel University, Uxbridge,
Middlesex, UB8 3PH United Kingdom. e-mail: [email protected].

0377-0427/93/$06.00 0 1993 - Elsevier Science Publishers B.V. All rights reserved


116 M. Greenhow / Floating-body problem

a@
+f (d)2+m
x2
= o

radiated z = q(x,t)
I
- waves \
radiated
aa
- = vn
WQ = 0 an waves -+

Fig. 1. The nonlinear boundary-value problem.

complex potential may be thought of as a flow (although solid boundaries may need to be
introduced, for example, along branch cuts) and, more pertinently, any fluid flow may be
described by an analytic function.
Two consequences immediately follow from this formulation. Firstly, many of the elementary
functions correspond to useful flows, examples being log(z), which is a source, sink or vortex
(depending on its coefficient), and exp( - ikz) which gives linearised water waves of wavenum-
ber k (oscillation in x and exponential decay in -y, i.e., depth below the free surface), see,
e.g., [30,36]. Secondly, any flow may be conformally mapped to another fluid flow. This
technique is particularly useful for finding the added mass of certain bodies moving in
unbounded fluid, if the flow around them can be mapped to uniform flow. The added mass can
then be deduced from the large z behaviour, being related to the dipole moment (coefficient of
the l/z term), so it is not necessary to solve for the more complicated flow local to the body,
see [33,41] for examples of simple body shapes. More complicated shapes may be treated by
discretisation and the Schwarz-Christoffel mapping.
These analytical techniques are, however, only successful for rather simple flow domains
satisfying simple conditions (usually zero normal flow) on the domain boundary. The water
wave problem, and the related problems of floating or submerged bodies outlined in Sections 2
and 3 respectively, are both interesting and hard because of the complicated nonlinear
free-surface boundary conditions, see Fig. 1. Although some progress had been made for rather
idealised problems, such as steady waves over a wavy seabed [22] or over a flat bottom [38], and
the zero-gravity entry of a wedge [21], the numerical solution of problems including both time
dependence and gravity were not practical until the work of Nichols and Hirt [34], who
discretised the entire fluid domain. Although some success was achieved for the early stages of
water entry of a cylinder, the method was computationally expensive and not suitable for more
general problems.
A breakthrough occurred when Longuet-Higgins and Cokelet [26] formulated the problem as
a boundary-value problem, requiring solution and evolution only on the free-surface boundary.
Since their interest was in periodic, but unsteady, flow of water waves as they break, they closed
the contour of integration by a conformal map (c = exp( - iz)) which wraps the contour around
joining the two vertical parts of the contour shown in Fig. 2; this then ensures periodicity, with
M. Greenhow / Floating-body problem 117

Fig. 2. Contours of integration. C, and C, denote parts of C on which either @ or T is known. Radiation or
periodic conditions are applied on the vertical boundaries.

large depths being mapped to the origin. A Green’s function technique was then applied in this
mapped plane, resulting in a Fredholm integral equation of the first kind for the unknown
normal fluid velocity at the free surface. The solution, when evolved in time according to the
(transformed) dynamic free-surface boundary conditions, see Section 2, developed a saw-tooth
instability which required smoothing. Nevertheless, the method, when mapped back to the
physical plane, gives spectacular results for the development of wave overturning, and the high
velocities and accelerations in the overturning and jet regions are of significant engineering
interest. Longuet-Higgins and Cokelet [27] also applied their method to the instability growth
rates of steep waves, obtaining very impressive agreement with earlier analytical theories.
Further developments of the Longuet-Higgins and Cokelet method include that of New et al.
[32] who considered waves in finite-depth water (the mapped region then becomes an annulus
with the Green’s function satisfying the bottom boundary condition a priori). Despite the
somewhat artificial nature of the initial conditions used in all the above calculations (i.e., a
sinusoidal wave of large amplitude, which may be further perturbed by applying a surface
pressure or modifying the water depth), the final stages of overturning are convincing. This
leads to the idea that wave breaking may be, in some sense, self-similar, and that the crest
region is largely uncoupled from the rest of the wave during overturning. Thus Longuet-Higgins
[24,25], New [31] and Greenhow [12] have proposed local models for the loop and jet regions.
Other authors have used finite-amplitude, steady wave theories, such as that of [38], as being
more realistic initial conditions. Tanaka et al. [40], Cooker et al. [S] and Cooker and Peregrine
[7] took solitons as initial conditions for waves in shallow water. They considered instabilities,
motion over semicircular obstructions on the seabed, and solitons in head-on collision. This last
situation represents a single soliton impacting on a vertical seawall, and was shown, in some
cases, to develop extremely high pressures, and accelerations of the order of hundreds, or even
thousands, of g in a very small impact region. Clearly very small time and space discretisations
need to be made for accurate calculations.
All of the above theories become invalid when the breaking wave jet falls onto the front
surface of the wave, and the numerical schemes quickly break down. This is, of course, to be
expected since the flow region is no longer simply connected, and within the loop region, the
analytic continuation of /?( z, t) fails to remain analytic. Thus the jet must have moved onto a
different Riemann sheet, and its space derivatives will no longer match those of the front face
of the wave: this results in splashing at the jet impact region, and the loop flow soon develops
118 M. Greenhow / Floating-body problem

into a large vortex which propagates into the fluid. The problem of continuing the calculations
after jet touch-down is therefore a formidable challenge. A possible way forward is to keep
track of the two layers of fluid from the wave front and jet respectively, and to apply a
boundary condition across the interface of a jump in potential which conserves the Kelvin
impulse. This method has been successfully to study the collapse beyond jet re-entry of
axisymmetric bubbles, see [3].
Although conformal mappings are used in the above numerical schemes, the problem is not
formulated directly in terms of the complex velocity potential P(z, t). One possibility is to
construct p from singularities exterior to the fluid region, choosing their coefficients in such a
way as to minimise the errors when satisfying the free-surface conditions, see [29]. Another is to
construct p from a vortex sheet placed at the free surface; the (unknown) strength distribution
and its evolution are determined by the free-surface conditions, see [2]. Since the governing
Fredholm equations are of the second kind with globally convergent Neumann series, the
resulting matrix equations may be solved iteratively - a significant advantage over direct
Gaussian elimination methods.
Since p is analytic, Cauchy’s theorem holds:

for any contour C drawn within the fluid or along its boundary, and with z0 outside C.
Evidently, any other analytic function will satisfy this equation; to evolve the solution, we will
need (2) with a@/& in place of p, see Section 2. Another possibility includes the velocity
potential gradient as a function of w = u + iu, which results in an integral equation for the
(unknown) normal derivative of @, see [9]. An interesting feature of that paper is that the
higher time derivatives of the analytic function may be easily calculated since the kernel of the
integral equation is the same. Knowing these enables much larger time steps to be taken for
any given accuracy.
Section 2 outlines the method of Vinje and Brevig [45]. The major difference is that Vinje
and Brevig do not build in periodicity of the flow into their method (by mapping the contour C,
or otherwise), but rather solve (2) directly in the physical plane. Thus, although periodic waves
and flows may be considered, their method has the advantage that other conditions may be
applied at the vertical boundaries, and more importantly, floating or submerged bodies may be
introduced. In these cases periodicity at the vertical boundaries is not appropriate and one
needs to apply a radiation condition at these boundaries to account for the outgoing diffracted
waves, and any waves radiated by a moving body. This problem has not been satisfactorily
resolved, see [5,46]. These authors attempt to match the interior nonlinear region with a linear
outer region satisfying the Orlanski condition [35]

where C is calculated from the rate of change of @ in the interior region. (Using C from the
phase speed of linearised waves the Somerfeldt condition was also tried.) The results, which
always gave some reflection, were disappointing. A further problem was that the nonlinear
(finite-amplitude) waves in the interior region do not decay in amplitude as they propagate
A4. Greenhow / Floating-body problem 119

Fig. 3. Wedge entry at high speed. Note that the spray may separate from the wedge side. The analytic results of
Appendix 1 are shown dotted.

outwards, and were therefore difficult to match to the outer domain which has boundary
conditions applied at the undisturbed free-surface level. This is in contrast to the three-dimen-
sional situation where wave amplitude decays like l/fi, and where the inner and outer
regions may be matched successfully, see [23].
More pragmatic approaches appear to work well in some situations. For example, for the
problem of water entry or water exit into otherwise calm water, it is possible to place the
vertical boundaries sufficiently far away, so that no significant disturbance reaches them during
the transient phase of interest, see Fig. 3. Another successful technique was used for waves
moving over a submerged cylinder [6]. For this case the transmitted wave propagated into an
“energy sponge” region where its amplitude was artificially reduced to zero at the end wall.
This apparently gave insignificant reflection back into the physical region of interest around the
cylinder. A similar technique has been applied in [l].

2. The method of Vinje and Brevig

The contour integral of Cauchy’s theorem above may be split into two parts: C, where @ is
known and C, where !P is known, see Fig. 2; these quantities are either specified initially or
known from the evolution equations below. Thus @ and its time derivative are known on the
free surface (W/at is given from the Bernoulli equation), while !P and its time derivative are
known on the bottom (both are zero here) and, if present, on a body surface (!P and W/at are
here specified by the body geometry and its velocity). Taking zO ( =x,, + iyO> on the contour C,
and using either the real or imaginary parts of (2) gives integral equations of the second kind:

(4)
120 M. Greenhow / Floating-body problem

for z,, on C,, and

r@(x,, yo, t) + Re [/cd=]


i =O,

for z0 on C,. Similar equations hold for the time derivative of p. We have further assumed
that C is smooth; at corners, either due to the presence of a solid body or due to discretisation
of C, we use the angle subtended by the collocation points Ed in place of r in these equations.
On the vertical boundaries in Fig. 2, we apply either periodicity, or an appropiate fixed wall or
wavemaker condition, making it part of C,.
To step forward in time we use the boundary conditions for the two-dimensional free-surface
problem as given in Fig. 1 (for a derivation see, e.g., [36]). Following [26], we write the
free-surface conditions following a free-surface particle (a Lagrangian description of the flow)
as
DZ aP D@ P,
-=
Dt
u+iu=w*=--, -=~~*-pJ7----,
Dt
(6)
aZ P
where g is the gravity, p the density, and the material derivative is given by

w
-
a(*)
= at + V@ V(e).
Dt
These equations are used to evolve the position and value of @ of the free-surface particles to
the next time step. Specifically a single-step Runge-Kutta method is used to calculate the first
three steps, after which we may use a fourth-order Hamming predictor/corrector method. A
numerical derivative of p is calculated. The forces on the body (if present) are calculated by
integrating the hydrodynamic pressure, given by Bernoulli’s equation, over its surface; this gives
its acceleration and hence its new position and velocity. For further details see [15,44].
As mentioned above, the collocation method is applied to (4) and (5). Assuming a linear
variation between the collocation points, Vinje and Brevig perform the integrations analytically.
This results in an N x N matrix equation Ax = B for the unknown part X of p at each of the
N collocation points. The elements of the N X N matrix A consist of logarithmic terms,
requiring typically 40% of the calculation time for their evaluation. However, the matrix is also
used at each time step for the calculation of the unknown part of ap/at, and could also be used
for the calculation of higher derivatives of p, as in the method of [9].
When no body is present, it may be advantageous to use an iterated scheme to solve the
matrix equation, since we have an excellent initial guess given by the solution at the previous
time step, see [2,23]. Against this, we have already noted that the calculation time of A is
dominant, and iteration becomes awkward if we wish to apply regridding of the collocation
points, see [lo]. That paper illustrates the power of this method: a comparison of experimental
and numerical wave tanks,. both given identical wavemaker motions, is made for the wave
profile at various positions along the tanks. With 550 unknowns and about 4000 time steps, the
results agree very well, even to the resolution of wave breaking. The entire simulation took
about 30 hours on a CRAY 1 supercomputer.
In the case of ship motion, and particularly for the water-entry and -exit cases presented
here, we note that the number of collocation points N continually changes as the body
submerges or emerges; here we use direct Gaussian elimination. A further problem for
M. Greenhow / Floating-body problem 121

surface-piercing bodies occurs at the intersections of the free and body surfaces. Except in
special cases, the complex velocity potential /3 or its time derivatives are known to be singular
here, see, e.g., [39,43]. The introduction of viscosity may, in principle, be needed. In the present
formulation, we have both @ and ?P specified at these points. Although no theoretical
justification exists, we remove the two intersection points from the calculations and solve an
(N-2)x(N-2) system of equations. Then, treating the intersection points as ordinary
free-surface points for the purposes of time stepping appears to give acceptable results, see [23]
for the wavemaker problem and [13,14] for wedge and cylinder entry. These cases may be
compared with theoretical results. For the impulsively-moving wavemaker case, the numerical
results lie very close to the theoretical free-surface profile (which is logarithmic, see [37]),
except near to the intersection point. For high speed wedge entry, the self-similarity of the flow
has the consequence that the arc length along the free surface between any two free-surface
particles is maintained, see [ll]. The results of [13] satisfy this stringent check, and other checks
such as mass conservation, accurately only for slender wedges. When the deadrise angle (the
angle between the body and free surface) becomes smaller than 45”, the calculations become
unreliable and quickly break down. Resolution problems arise because of the spray jets of fluid
which move up the side of the wedge at great speed, see Fig. 3. Furthermore, the initial
condition for this problem is itself singular: thus resolving the jets more accurately does not
remove the problem, but rather makes it worse. Zhao and Faltinsen [49] point out that the
pressure in these jets is almost atmospheric, and claim success for deadrise angles as small as
1.5” by simply truncating the jets with an artificial boundary on which the pressure is atmo-
spheric. A better procedure may be to match to an analytic model of the outer (jet) region as in
the analytical work of [20,47].
A similar problem arises for the case of wedge exit; we present new results in Fig. 4.
Although no jets occur, the intersection point particles still move very quickly for small deadrise
angles. Some free-surface profiles are shown in Fig. 4, together with a comparison with the
transient linear wavemaker theory of [28], which is expected to be valid for slender wedges

Fig. 4. Wedge exit from initially calm water. On the left part the analytic results of Appendix 1 are shown dotted,
while on the right part the dotted line shows the contribution of buoyancy according to linear theory.
122 M. Greenhow / Floating-body problem

(deadrise angles approaching 90”). Analytical results for the free-surface profile for water entry
and water exit are given in Appendix 1. For smaller deadrise angles, no simple analytical theory
is yet available. (Unlike the high-speed water-entry case, we must here include both gravity and
the length scale of the initial submergence depth; thus the problem is Froude number F,.
dependent.)
For submerged bodies, we do not, of course, have the above intersection point problems.
However, a difficulty does arise with the integration around the contour, which now involves a
branch cut. We account for this when integrating around inner and outer contours in Fig. 2
(right part). Furthermore, for a fixed body, we do not know the constant value of ly on its
surface. However, referring to (5), we see that integrating around the cylinder gives
iYf dz
-dz=iP - =o, (8)
#cylZ -zo
#cylZ -20

when .zo is outside the cylinder, and

Re[itY,&dz] = Re[ri!P] = 0,

when z,, is on the cylinder. Thus the unknown constant value of !P is immaterial, but may be
calculated around the cylinder from (4) as a check. For the moving cylinders considered here,
we do know the variation of !P around the cylinder from the body boundary condition to within
this unknown constant value, see [4] for details. More general initial conditions, where the
circulation around the cylinder and hence the value of Y’ are specified, can also be considered,
see [42].
The other new results presented here concern the impulsive horizontal motion of a
submerged circular cylinder. The corresponding steady motion problem has been extensively
studied by Havelock [17], using the linearised free-surface condition. When the cylinder is
sufficiently deeply submerged and moving sufficiently slowly, his results agree well with the
numerical schemes of [1,16]. The present results concentrate on transient motion of the
cylinder very near the free surface. Appendix 2 presents and extends Havelock’s [I81 theory for
this situation. We see in Fig. 5, taken from [19], that the assumptions of linearity, as well as the
approximate nature of the satisfaction of the body boundary condition, preclude the use of the
analytical results in Appendix 2, unless the body is very deeply submerged. However, the
numerical scheme is capable of dealing with high speeds at quite small submergences.

3. Conclusions

Over the past fifteen years, some fast and reliable programs have been developed for the
nonlinear water wave problem. We can now calculate the modulation of a wavetrain, interac-
tions between waves or solitons, and overturning of a wave crest; such work has further
stimulated analytical work. For two dimensions, the main outstanding problems appear to be
(i) the successful implementation of radiation conditions, so that the computational domain
may be kept small;
(ii) a satisfactory way of treating the intersection points for surface-piercing bodies, espe-
cially when the deadrise angles are small; and
M. Greenhow / Floating-body problem 123

,w I I I t

75-

. l
dr Kev
Force Magnitude

--- V&B Horizootal


Compownt FH

---X0 hW\ v&E3 vertical


Component F,
>r x ~e~~avelock‘s
‘\ XXX
y* xx c
_.
-.._
-- . .
_
.._

Fig. 5. Impulsive motion of a horizontal submerged cylinder. The analytic results for the forces are shown dotted.
This figure is taken from [19].

(iii> a theoretical framework for treating waves after breaking, and its numerical implemen-
tation.
When bodies are introduced, or waves are generated by a wavemaker, the most successful
approach to date appears to be that of [44,45], described in Section 2. Although the method
124 M. Greenhow / Floating-body problem

may be applied to the large computational domain of a “numerical wave tank”, as in [lo], many
interesting calculations may be made at modest cost by considering transient problems where
the initial conditions may be given exactly. Examples of water entry and water exit, and the
motion of submerged bodies have been presented; the results are compared with those from
linear theories given in the appendices.

Appendix 1

For comparison with the numerical results calculated by the Vinje and Brevig method we
here extend the transient wavemaker theory of Mackie [28]. Mackie considers the entry of a
solid wedge of half-angle E; if the wedge is slender (E is small), it is appropriate to use the
linearised free-surface condition, see, e.g., [36],
a2@ a@
--g--o, applied on y = 0. (LO)
at2 ay

The free-surface elevation is given by

(11)
The body boundary condition is not applied on the wedge surface, but rather on an “equivalent
wavemaker” at x = 0, satisfying
aa
- =u(y, t), on x=0. (12)
ax
Here u( y, t) is chosen to simulate the motion of the wedge, i.e.,
forOGy<ut,
u(y, t)=
L i” fory>ut

for wedge entry at constant velocity u, and


(13)

forO<y<d+ut,
otherwise,
(14)
for wedge exit at constant velocity u, with the vertex initially at depth d.
Mackie solves the general problem by Fourier and Laplace transforms; the Fourier trans-
form of the free-surface elevation

H(h, t) = Lm+c, t) cos(Ax) dh (15)

is given by

H(A, t) = -j-)&i, t) cos[iG(t -T)] dr, (16)


where U(A, t) is the Fourier transform of u(x, t). This expression may be interpreted as a
M. Greenhow / Floating-body problem 125

convolution over the previous history of wavemaker motion U(h, 7) from r = 0 to the present
time t. Substituting the Fourier transforms U(A, t) for the entry and exit problems gives

fif)
1
)
H(A, t) = -aA2 e --Aut- cos( + u/psin(fit
h2u2 + Ag
(17)

and

EU e -hd Au[cos(lihgl) - e-AUr] e Ad


H(A,t)= A + sin&t&
A2u2 + Ag Ag
-11

(18)
respectively. We then take the inverse Fourier transform

V(X, t> = ;imH(A, t) cos(Ax) dh (19)


to give the profiles shown dotted in Figs. 3 and 4. For slender wedges there is quite good
agreement with the numerical results. In these figures, we have started drawing the free-surface
profiles not at x = 0, but at the wedge surface. Although this is sensible, we note that the
necessity of making such a choice arises from the inadequacy of the above theory to account
properly for the actual wedge profile. This inadequacy is likely to result in serious error for
wedges which are not slender. Indeed, we see immediately that, according to this (linear)
theory, the above profiles are linear in E (the half-angle of the wedge). Another source of
disagreement may be that for nonslender wedges the free-surface deformations are so steep
that the linearised boundary condition (10) is likely to give serious errors. Nevertheless, when
applied to appropiate cases, the above analytic results are capable of providing a simple
explanation of the numerical results, and, since gravity is included, the analytic solutions evolve
to give realistic wavetrains (not shown in Figs. 3 and 4).

Appendix 2

We here present and extend the dipole theory of Havelock [18] for the impulsive horizontal
motion of a horizontal cylinder with axis at depth d. In unbounded fluid (i.e., no free surface
present) the flow around a cylinder of radius r, moving horizontally with velocity u, is given by
a dipole /3(z, t) = ur2/z. The effect of free-surface proximity is to introduce an image of this
dipole plus a memory term representing radiated waves:

p(=, q =; - 2!L
z - 2id

+ Ur2~~fd~~m[eik(~~(p/k)‘~iXt-~) _ e-ik(~-(p/k)1’zXr~T)]e-ik-2kd~dk.
0 0

(20)
126 M. Greenhow / Floating-body problem

In this case the force Z =X + iY may be obtained from the Blasius theorem (the time-depen-
dent terms are nonsingular and therefore do not contribute to the contour integral). Thus

Z= -ipi (21)
This gives Havelock’s result for the horizontal force, or resistance X:

sin(ay(y - 1)) sin(ay(y + 1))


x = 4Trgpr4K; y5e-6y* dy, (22)
y-l - Y+l I

where

(Y = KOUf, 6 = 2K,d. (23)


The vertical force Y is given by

1y5e-syz
1 - cos(ay(y + 1))

1
dy . (24)
Y+l

Following [18], we use the method of stationary phase to give the large time asymptotics of
these expressions. Thus

and

+ 1 + 6 - 62e-SEi(S) + +2e-S/4--- r
1/KoUt
COS(+(7T. - K,+t))
1
. (26)

The first terms, which arise from a singularity on the contour of integration, represent the
steady-state forces and correspond to the forces calculated from the steady (nonimpulsive)
motion problem, see [48], after correcting a sign in their expressions. The last terms are
oscillatory. The numerical calculations are qualitatively similar for the very deeply submerged
cylinder case, although for initial stages of the motion the numerical and analytic results
disagree significantly. For cylinders near the surface, as shown in Fig. 5, the theories disagree
significantly, as we might expect in such a nonlinear situation, and a comparison of the large
time results is not possible because the numerical results produce a large wave which breaks,
terminating the calculations.
M. Greenhow / Floating-body problem 127

References

[l] G.R. Baker, D.I. Merion and S.A. Orszag, Application of a generalised vortex method to nonlinear free surface
flows, in: J.H. McCarthy, Ed., Proc. Third Con& Ship Hydrodynamics, Paris (Office of Naval Research, 1981)
111-3-1-111-3-13.
[2] G.R. Baker, D.I. Merion and S.A. Orszag, Generalised vortex methods for free-surface flow problems, J. Fluid
Mech. 123 (1982) 477-501.
[3] J.P. Best, The dynamics of underwater explosions, Ph.D. Thesis, Univ. Wollongong, Australia, 1991.
[4] P. Brevig, M. Greenhow and T. Vinje, Extreme wave forces on submerged cylinders in: Proc. 2nd BHRA
Internat. Symp. on Wave and Tidal Energy, Cambridge (British Hydraulics Research Assoc., 1981) 143-166.
[5] T. Christiansen, An investigation of non-linear wave motion, Report No. 2.9, NTNF Research Program, A.S.
Veritec, Oslo, 1986.
[6] R. Cointe, Waves travelling over a submerged cylinder: nonlinear vs. second-order theory, in: J. Grue, Ed.,
Proc. 4th Internat. Workshop on Water Waves and Floating Bodies, @stese, Norway, 1989, 39-43.
[7] M. Cooker and D.H. Peregrine, Computation of violent motion due to waves breaking against a wall, Proc.
Coastal Engrg. 1 (1990) 164-176.
[8] M. Cooker, D.H. Peregrine, C. Vidal and J.W. Dold, The interaction between a solitary wave and a submerged
semicircular cylinder, J. Fluid Mech. 215 (1990) l-22.
[9] J.W. Dold and D.H. Peregrine, Steep unsteady water waves: an efficient computational scheme, Report No.
AM-84-04, School Math., Univ. Bristol, 1984.
[lo] D.G. Dommermuth, D.K. Yue, W.-M. Lin, R.J. Rapp, ES. Chan and W.K. Melville, Deep-water plunging
breakers: a comparison between potential theory and experiments, J. Fluid Mech. 189 (1988) 423-442.
[ll] P.R. Garabedian, Oblique water entry of a wedge, Comm. Pure Appl. Math. 6 (1953) 157-165.
[12] M. Greenhow, Free-surface flows related to breaking waves, J. Fluid Mech. 134 (1983) 259-275.
[13] M. Greenhow, Wedge entry into initially calm water, Appl. Ocean Res. 9 (4) (1987) 214-223.
[14] M. Greenhow, Water-entry and -exit of a horizontal circular cylinder, Appl. Ocean Res. 10 (4) (1988) 191-198.
[15] M. Greenhow, T. Vinje, P. Brevig and J. Taylor, A theoretical and experimental study of the capsize of Salter’s
duck in extreme waves, J. Fluid Mech. 118 (1982) 221-239.
[16] H.J. Haussling and R.M. Coleman, Nonlinear water waves generated by an accelerating circular cylinder, J.
Fluid Mech. 92 (1979) 767-781.
[17] T.H. Havelock, The forces on a circular cylinder submerged in a uniform stream, Proc. Roy. Sot. London Ser. A
157 (1936) 526-534.
[18] T.H. Havelock, The wave resistance of a cylinder started from rest, Quart. J. Mech. Appl. Math. 22 (2) (1949)
401-414.
[19] T. Hepworth, An investigation of the motion of a submerged cylinder moving below a free surface with constant
velocity, 4th year project report, Dept. Math. Statist., Brunel Univ., 1991.
[20] S.D. Howison, J.R. Ockendon and SK. Wilson, Incompressible water-entry problems at small deadrise angles,
J. Fluid Mech. 222 (1991) 215-230.
[21] O.F. Hughes, Solution of the wedge entry problem by numerical conformal mapping, J. Fluid Mech. 56 (1)
(1972) 173-192.
[22] F. John, Two-dimensional flows with a free boundary, Comm. Pure Appl. Math. 6 (1953) 497-503.
[23] W.-M. Lin, J.N. Newman and D.K. Yue, Nonlinear forced motions of floating bodies, in: W.C. Webster, Ed.,
15th Symp. Naval Hydrodynamics, Hamburg (National Academy Press, Washington, DC, 1984) 33-49.
[24] M.S. Longuet-Higgins, Parametric solutions for breaking waves, J. Fluid Mech. 121 (1982) 403-424.
1251 M.S. Longuet-Higgins, Rotating hyperbolic flow: particle trajectories and parametric representation, Quart. J.
Mech. Appl. Math. 36 (1983) 247-270.
1261 M.S. Longuet-Higgins and E.D. Cokelet, The deformation of steep surface waves on water. I. A numerical
method of computation, Proc. Roy. Sot. London Ser. A. 350 (1976) l-26.
[27] M.S. Longuet-Higgins and E.D. Cokelet, The deformation of steep surface waves on water. II. Growth of
normal mode instabilities, Proc. Roy. Sot. London Ser. A 364 (1978) l-28.
[28] A.G. Mackie, Gravity effects in the water entry problem, J. Austral. Math. Sot. 5 (1965) 427-433.
1291 P. McIver and D.H. Peregrine, The computation of unsteady and steady free-surface motions using a small
number of singularities in the exterior flow field, Report No. AM-81-13, School Math., Univ. Bristol, 1981.
128 M. Greenhow / Floating-body problem

[30] L.M. Mime-Thompson, Theoretical Hydrodynamics (Macmillan, New York, 5th ed., 1986).
[31] A.L. New, A class of elliptical free-surface flows, J. Fluid Mech. 130 (1983) 219-239.
[32] A.L. New, P. McIver and D.H. Peregrine, Computations of overturning waves, J. Fluid Mech. 150 (1985)
233-251.
[33] J.N. Newman, Marine Hydrodynamics (MIT Press, Cambridge, MA, 1977).
[34] B.D. Nichols and C.W. Hirt, Nonlinear hydrodynamic forces on floating bodies, in: J.H. McCarthy, Ed., 2nd
Internat. Conf on Numer. Ship Hydrodynamics, Berkeley (Office of Naval Research, 1977) 382-394.
[35] I. Orlanksi, A simple boundary condition for unbounded hyperbolic flows, J. Comput. Phys. 121(1976) 251-269.
[36] A.R. Paterson, A First Course in Fluid Dynamics (Cambridge Univ. Press, Cambridge, 1983).
[37] D.H. Peregrine, Flow due to a vertical plate moving in a channel, Unpublished note.
[38] M.M. Rienecker and J.D. Fenton, A Fourier method for steady water waves, J. Fluid Mech. 104 (1981) 119-137.
[39] A.J. Roberts, Transient free-surface flows generated by a moving vertical plate, Quart. .Z.Mech. Appl. Math. 40
(1) (1987) 129-158.
[40] M. Tanaka, J.W. Dold, M. Lewy and D.H. Peregrine, Instability and breaking of a solitary wave, J. Fluid Mech.
185 (1987) 235-248.
[41] J.L. Taylor, Some hydrodynamic inertia coefficients, Philos. Mug. 9 (7) (1930) 161-183.
[42] A.F. Teles da Silva and D.H. Peregrine, Nonlinear perturbations of a free surface induced by a submerged
body: a boundary integral approach, Engrg. Analysis with Boundary Elements 7 (1990) 214-222.
[43] T. Vinje, On the small time expansion of nonlinear free surface problems, in: J. Grue, Ed., Proc. 4th Znternat.
Workshop on Water Waves and Floating Bodies, Qstese, Norway, 1989, 245-250.
[44] T. Vinje and P. Brevig, Nonlinear, two-dimensional ship motions, Ship Res. Inst. Norway R-112-81, 1981.
[45] T. Vinje and P. Brevig, Breaking waves on finite depth water; a numerical study, Ship Res. Inst. Norway
R-118-81, 1981.
[46] T. Vinje, P. Brevig and M. Xie, A numerical approach to nonlinear ship motion, in: 14th Symp. Naval
Hydrodynamics, Ann Arbor (Office of Naval Research, 1982) l-32.
[47] H. Wagner, Uber St&s- und Gleitvorgange an der Oberflache von Fliissigkeiten, Z. Angew. Math. Mech. 12
(1932) 193-215; see also: The Phenomena of Impact and Planning on Water (N.A.C.A., 19361 Translation 1366,
l-57.
[48] J.V. Wehausen and E.V. Laitone, Surface waves, in: S. Flugge, Ed., Handbook of Physics 9 (Springer, Berlin,
1960) 446-778.
[49] R. Zhao and 0. Faltinsen, Water entry of a two-dimensional body, in: Proc. 6th Znternat. Workshop on Water
Waves and Floating Bodies, Woods Hole, 1991.

You might also like