Dynam Ics of Watem Waves: Selected Papers of Michael Longuet-Higgins
Dynam Ics of Watem Waves: Selected Papers of Michael Longuet-Higgins
DYNAM ICS OF
WATEM WAVES
Selected Papers of
Michael Longuet-Higgins
Volume 3
jT;--
S G Sajjadi
editor
World Scientific
DYNAMICS OF
WATER WAVES
Selected Papers of
Michael Longuet-Higgins
V olum e 3
ADVANCED SERIES ON OCEAN ENGINEERING
Series Editor-in-Chief
Philip L- F Liu (Cornell University)
Published
Vol. 24 Introduction to Nearshore Hydrodynamics
by lb A. Svendsen (Univ. of Delaware, USA)
Vol. 25 Dynamics of Coastal Systems
by Job Dronkers (Rijkswaterstaat, The Netherlands)
Vol. 26 Hydrodynamics Around Cylindrical Structures (Revised Edition)
by B. Mutlu Sumer and Jorgen Fredsoe (Technical Univ. of Denmark,
Denmark)
Vol. 27 Nonlinear Waves and Offshore Structures
by Cheung Hun Kim (Texas A&M Univ., USA)
Vol. 28 Coastal Processes: Concepts in Coastal Engineering and Their
Applications to Multifarious Environments
by Tomoya Shibayama (Yokohama National Univ., Japan)
Vol. 29 Coastal and Estuarine Processes
by Peter Nielsen (The Univ. of Queensland, Australia)
Vol. 30 Introduction to Coastal Engineering and Management (2nd Edition)
by J. William Kamphuis (Queen’s Univ., Canada)
Vol. 31 Japan’s Beach Erosion: Reality and Future Measures
by Takaaki Uda (Public Works Research Center, Japan)
Vol. 32 Tsunami: To Survive from Tsunami
by Susumu Murata (Coastal Development Inst, of Technology, Japan),
Fumihiko Imamura (Tohoku Univ., Japan),
Kazumasa Katoh (Musashi Inst, of Technology, Japan),
Yoshiaki Kawata (Kyoto Univ., Japan),
Shigeo Takahashi (Port and Airport Research Inst., Japan) and
Tomotsuka Takayama (Kyoto Univ., Japan)
Vol. 33 Random Seas and Design of Maritime Structures, 3rd Edition
by Yoshimi Goda (Yokohama National University, Japan)
Vol. 34 Coastal Dynamics
by Willem T. Bakker (Delft Hydraulics, Netherlands)
Vol. 35 Dynamics of Water Waves: Selected Papers of Michael Longuet-Higgins
Volumes 1-3
edited by S. G. Sajjadi (Embry-Riddle Aeronautical University, USA)
Vol. 36 Ocean Surface Waves: Their Physics and Prediction
(Second Edition)
by Stanisfaw R. Massel (Institute of Oceanology of the Polish Academy of
Sciences, Sopot, Poland)
*F or the co m p lete list o f titles in this se ries, p lea se w rite to the P ub lisher.
Advanced Series on Ocean Engineering — Volume 35
DYNAMICS OF
WATER WAVES
Selected Papers of
Michael Longuet-Higgins
Volume 3
editor
S G Sajjadi
Embry-Riddle Aeronautical University, USA
№ World Scientific
NE W J E R S E Y • L ONDON • S I N G A P OR E • B E I J I N G • S H A N G H A I • HONG KONG • TAI PEI • C HE NNA I
Published by
All rights reserved. This book, o r parts thereof, may not be reproduced in any form o r by any means, electronic
o r mechanical, including photocopying, recording o r any information storage and retrieval system now known
o r to be invented, without written permission from the Publisher.
For photocopying o f material in this volume, please pay a copying fee through the Copyright Clearance Center,
Inc., 222 Rosewood Drive, Danvers, MA 01923, USA. In this case permission to photocopy is not required from
the publisher.
It is a pleasure and honour to write this brief preface to introduce this three-
volume selection of papers by a close colleague, Professor Michael Longuet-
Higgins, FRS.
As a biographical sketch written by him follows this preface I will limit
myself to some thoughts of mine having to do with his influence on our
community. From the recognition which Professor Longuet-Higgins has received
in appreciation of his work I would like to mention his work on microseisms.
Scientists have long known about microseisms, but, until recently, no one
could determine where they came from. Microseisms were first recorded as a
strange, continuous buzz on the earliest seismometers, devices that measure
earthquakes. Every year, the cumulative energy released by these small vibrations
equals the amount of energy released globally from earthquakes. Records of
microseismic activity give scientists a history of wave interaction in Earth’s
oceans since the early 20th century. They are also used to examine the history
of storms over the ocean. Scientists are interested in learning where these
microseisms originate because the information can help them monitor stress in
Earth’s crust.
The theory of the origin of microseisms was first introduced in 1950 by
Michael Longuet-Higgins whilst at the University of Cambridge in England.
Longuet-Higgins suggested that the vibrations originated in places where ocean
waves were traveling in opposite directions toward each other at the same
frequency and at a certain ocean depth. According to his theory, when these
waves collide, they combine to form stationary waves that remain in a constant
position over large areas of the ocean. These waves create tall, pulsing columns
of pressure that repeatedly beat down on the ocean floor, causing it to vibrate.
The vibrations generate seismic surface waves, which spread out thousands of
miles and are detected by seismometers as noise. This new study on microseisms,
which appeared in the March 2007 issue of the Proceedings of the Royal Society,
Series A was part of interdisciplinary collaboration, which included Longuet-
Higgins and researchers from the California Institute of Technology, Pasadena
and the Hydrologic Research Center in San Diego.
This three volume set presents selected original research papers of Professor
Longuet-Higgins in the various areas of water waves, ocean dynamics, and
fluid mechanics. I believe they will serve as a milestone and beacon for future
generations.
Apart from the subject area covered in these three volumes, Longuet-Higgins
has published papers in other fields such as electromagnetic measurement of
tidal streams and ocean currents, time-varying currents depending on the Earth’s
rotation, projective geometry, etc. These papers have not been included in the
present selection, but it is hoped that they will be published in a separate volume
in the near future.
Finally, my thanks are due to a number of publishers for their permissions for
photographic reproduction of the original papers. In particular I wish to thank the
Royal Society of London, the Oxford University Press, the Cambridge University
Press, MacMillan Publishing Co., the Journal of Physical Oceanography, the
Journal of Marine Research, the American Physical Society, the American
Geophysical Union, Prentice-Hall, Pergamon Press and Elsevier Publishing Co.
S G Sajjadi
ERAU, Florida
December 2008
Preface by Michael Longuet-Higgins
The practical consequences of wave motion for both the coastal engineer and the
geophysicist are many and varied, and new applications constantly arise to keep
the subject alive. At the same time, surface waves challenge the fluid dynamist to
find an explanation for such spectacular phenomena as wave breaking, when the
upper surface overturns on itself.
The present collection of original papers is not intended to cover the whole
subject. It is simply a selection of basic contributions by a single author and
his collaborators. On reflection, the papers can be seen to have been guided by
certain points of view: (1) a preference for simple or “physical” explanations
where possible, and particularly for geometrical interpretations. (2) a desire
to solve problems by accurate analysis initially, but then to find simpler
approximate models, easier to apply in practice. (3) an acknowledgement of
the need to compare theoretical results with field observations or laboratory
experiments. Like many other subjects, fluid mechanics advances by putting one
foot forward and then the other.
For convenience, this collection is arranged according to subjects. Included
in this first Volume are the author’s principal papers on microseisms; on mass
transport by water waves; on stochastic processes, especially applied to wind-
waves; on various mechanisms of wave generation by wind; and on the theory
and consequences of radiation stresses (momentum flux due to waves).
Conference papers are generally omitted, since usually they consists of reviews
of the original papers, doubtless in a more readable form. Preceding each group
of papers, there is a brief introduction giving the circumstances under which the
papers were written, and providing further background information.
As an appendix to Volume III, a list is given of the author’s papers on other
subjects. Some, as for instance time-varying ocean currents, are closely related to
papers in this Collection, while other subjects are widely different.
My thanks are due to Professor SG Sajjadi for offering to undertake
the editing of this selection of my papers. I would also like to express my
indebtedness to some of my mentors, colleagues and collaborators, aside
from those named explicitly in the subsequent text. In particular I wish to
mention CV Durrell, the senior mathematics don at Winchester College, for his
thorough mathematical grounding, and for his encouragement of my first forays
into mathematical research; to Dr George Deacon, leader of Group W at the
Admiralty Research Laboratory at Teddington, England, from 1944 to 1949 and
subsequently the first Director of the UK National Institute of Oceanography
at Witley, for his consistent support; to Norman Barber, a senior physicist in
group W; he was my first research supervisor who introduced me to “physical”
ways of looking at mathematical problems; to Sir Harold Jeffreys, for his
inspiring lectures on probability and statistics at Cambridge University; and to
Sir Geoffrey Taylor, for his inimitable style of research. Walter Munk has been
a long-time friend and advocate. It is be appropriate to mention those many
anonymous referees who have helped me on my scientific journey.
Perhaps the greatest credit of all should go to the subject of water waves
for affording the student such great pleasure and interest, and to the waves
themselves, for following so obediently the laws of mathematical physics.
A B rief B iography
In 1943, at the age of seventeen, the author entered Trinity College, Cambridge,
with a scholarship in mathematics from Winchester College, the school where he
had spent the first four years of World War II. Already he had acquired a taste for
research in geometry. By the summer of 1945, he had qualified for a Cambridge
BA in mathematics. He was then required to do three years “work of national
importance.” Fortunately for him, he was assigned to “Group W” (for waves) at
the Admiralty Research Laboratory, Teddington. This Group had been formed in
June 1944 to study the long-distance propagation of ocean waves in preparation
for projected military operations in the Pacific Ocean. The Group had been
spectacularly successful, and its lease of life had been extended, with wider terms
of reference. During three years at Teddington (1945-8), he worked not only
on the theory of wind waves but also on the geomagnetic induction of voltages
by tidal streams, and on the generation of oceanic microseisms. In September
1945, he returned to Cambridge to read for a PhD. There was, however, no
break in his research; he continued to develop the same subjects of interest,
reporting at the end of each term to Sir Harold Jeffreys, and later to Dr Robert
Stoneley. In 1951, he was awarded a 4-year research Fellowship (Title A) at
Trinity College. The first year (1951-2) he spent in the USA as a Commonwealth
Fund Fellow, staying first at the Woods Hole Oceanographic Institute on Cape
Code, Massachusetts, and then at the Scripps Institution of Oceanography at La
Jolla, California, with Walter Munk. At Scripps he became interested in wave
generation by wind, in the statistical properties of sea states, and in several other
topics.
On his return to England in 1952, he spent two years of his Research
Fellowship in Cambridge after which he accepted Dr George Deacon’s invitation
to join the newly formed N10, the UK’s National Institute of Oceanography,
at Witley in Surrey. There he was to spend thirteen happy and fruitful years,
working (mainly) on ocean waves. After 1963, he concentrated more on time-
varying ocean currents, especially those which depend essentially on the rotation
of the Earth. This period also included visiting appointments in the Mathematics
Department at MIT (1957-8), the University of Adelaide, Australia (1964) and
at the University of California, San Diego (1961—2 and 1966-7). In 1963, he
was elected to the Royal Society of London. From 1967 to 1969, he spent two
years assisting in the expansion of the Department of Oceanography at Oregon
State University, in Corvallis, Oregon, but in 1969, he was appointed to a Royal
Society Research Professorship, to be held jointly at Cambridge University, in
the Department of Applied Mathematics and Theoretical Physics, and at The
National Institute of Oceanography at Witley. Once there, he decided on a multi
faceted attack on the problem of wave breaking, involving both theory and
innovative experiments in the field and laboratory. He was also free to pursue
research in other subjects.
In 1989, on reaching the age of formal retirement, he wished to continue
doing research and for two years joined the La Jolla Institute in San Diego,
California. In 1991, he was appointed to a position as a Senior Research Physicist
at the Institute for Nonlinear Science in the University of California, San Diego,
with an Adjunct Professorship at the Scripps Institution of Oceanography. There
he turned attention to the natural sources of underwater sound, particularly the
sound produced by the creation of bubbles in breaking waves (previously this
sound was call “wind noise”). Financial support came mainly from the US Office
of Naval Research and from the National Science Foundation. Since his second
“retirement” in 2001, he has used his freedom as a Research Physicist Emeritus
to indulge his interest in a variety of problems without the necessity of applying
for outside grants. These interests have included the damping of incoming swells
by sand ripples, and the construction of very simple but accurate approximations
to gravity waves of limiting steepness.
xi
Contents
Preface by S G Sajjadi v
A Brief Biography IX
Volume I
В. Oceanic Microseisms 3
C o n f e r e n c e papers (n ot in clu d ed ).
xii
C4. Mass Transport in the Boundary Layer at a Free Oscillating Surface, 146
J. Fluid Mech. 8 (1960) 293-306.
C5. Steady Currents Induced by Oscillations Round Islands, 160
J. Fluid Mech. 42 (1970) 701-720.
C6. Peristaltic Pumping in Water Waves, 181
J. Fluid Mech. 137 (1983) 393-407.
Volume II
H9. Wave Set-Up, Percolation and Undertow in the Surf Zone, 1033
Proc. R. Soc. Lond. A, 390 (1983) 283-291.
H10. On Wave Set-Up in Shoaling Water with a Rough Sea Bed, 1046
J. Fluid Mech. 527 (2005) 217-234.
Volume III
L5. Review of: ‘Breaking Waves’, (film) by G.B. Olsen and 1710
S.P. Kjeldsen,
J. Fluid Mech 57 (1973) 624.
L25. Bubbles, Breaking Waves and Hyperbolic Jets at a Free Surface, 1939
J. Fluid Mech. 127(1983) 103-121.
L32. A Stochastic Model for Sea Surface Roughness I. Wave Crests, 2024
Proc. R. Soc. Lond A 410 (1987) 19-33.
L37. Flow Separation Near the Crests of Short Gravity Waves, 2039
J. Phys. Oceanogr. 20 (1990) 595-599.
L45. The Crest Instability of Steep Gravity Waves or How Do Short 2126
Waves Break?, pp. 237-246 in The Air Sea Interface,
eds. M.A. Donelan, W.H. Hui and W.J. Plant, Toronto, Canada,
Univ. of Toronto Press (1996) 789 pp.
XXV
N4. Vertical Jets from Standing Waves: The Bazooka Effect, 2244
pp. 195-204 in IUTAM Symp. on Free Surface Flows,
ed. Y.D. Shikhmurzaev, Kluwer Acad. Publ. (2001).
N8.* Standing Waves in the Ocean, pp. 201-218 in Wind over Waves II:
Forecasting and Fundamentals o f Applications,
eds. S.G. Sajjadi and J.C.R. Hunt, Horwood Publ.,
Chichester, U.K., (2003) 232 pp.
1653
In tro d u c to ry N otes fo r P a r t L
L . U nsteady F ree-S u rface Flow s; W ave B reaking
Papers L I to L50
calculates the distribution of the “roughness” (i.e. the short wave steepness) at
the crests of the longer waves. Paper L38 calculates the distribution of roughness
over the whole sea surface.
The profile of a steep Stokes wave close to (but not actually at) its limiting
steepness, was examined earlier in Papers K5 and K6 on the almost-highest
wave. Now, in Papers L40, L41, and L45 to L47 are examined the instabilities
of these very steep waves. In Paper L40 it is shown that the principle (linear)
instability takes the form of a bulge on the forward face of the wave. In L41
it is shown that this instability is analytically continuous with the instability
of the complete wave in deep water and that it becomes unstable at the first
energy maximum of the whole wave, as calculated by Tanaka. A similar result
is predicted for the second energy maximum. In Paper L47 the form of the
instabilities is calculated with great accuracy, and they are shown to obey a
scaling law predicted by the analysis. Similar calculations are performed for
the crest instabilities of solitary waves. In Paper L46 the growth of the lowest
instability is followed numerically to its nonlinear stage, and it is shown to lead
to overturning of the wave crest, as in a plunging breaker. Interestingly it is found
that if the sign of the initial perturbation is reversed, the nonlinear stage is quite
different; the perturbation leads at first to a transition of the wave to an almost
steady wave of lower amplitude, followed by a long-term recurrence to the
original wave, and so on. Paper L48 contains a useful summary and discussion of
the preceding work, and adds new results.
Two papers in this series, both on the instability of a shearing flow, are
closely related to breaking waves. Paper L49 concerns the instability of the
shearing flows seen on the forwards face of the crest of a gravity wave (Type 2
instability). This instability quickly grows and then collapses into turbulence. A
simplified model (Section 4) shows that the instability may be considered as a
train of “vortex waves”, i.e. a horizontal array of vortices. A second Paper A50
considers a more general model, in which a surface layer travelling with uniform
velocity overlies a deeper layer with a smooth velocity profile, tending to zero at
comparative large depths. There is rough agreement before the horizontal scales
of the instabilities and those observed.
O f the remaining papers in this section, Paper L42 describes some paradoxical
results resulting from consideration of the water surface as a fractal, and Paper
L44 discusses how the sheet of water in a plunging jet may become unstable to
three-dimensional perturbations, as is often observed.
L. Unsteady Free-Surface Flows; Wave Breaking
1659
By M. S. L o n g u e t -H ig o in s , F.R.S.
Oregon State University, Corvallis, Oregon
A theoretical calculation ia made of the loss of energy by wave breaking in a random sea state
in terms of the spectral density function. In the special case of the equilibrium spectrum
F(cr) = адха~ь the proportion ro of energy lost per mean wave cycle is found to be given by
m = e-1/Sa
irrespective of the low-frequency cut-off in the spectrum.
Assuming that in the equilibrium state the loss of energy by breaking is comparable to
that supplied by the wind, one can estimate the constant a in terms of the drag coefficient
of the wind on the sea surface. It is found that
a = -Ц Щ Ш О С Ц р ^/р ^)].
Taking a representative value of G one finds a = 1*3 x 10“*, which falls within the range of
observed values of a. The above equation for a is rather insensitive to the various assump
tions made in the analysis.
There is some evidence, derived from observation, that a may not in faot be quite con
stant, but may decrease slightly as the wave age (gtjU) or the non-dimensional fetch (gxjU*)
is increased. I t is suggested that the drag coefficient may behave similarly.
I n t r o d u c t io n
The ‘equilibrium law ’ for wind-generated surface waves, as first derived by Phillips
( 1958 ) is that over a certain range of wave frequencies cr the spectral density of the
surface elevation £ is given by
where g denotes the acceleration of gravity and a is an absolute constant. The fre-
quency-dependence of the spectrum ( 1-1) has been approximately verified by ob
servations over a wide range of frequencies (Phillips 1966). Some experimentally
determined values of a, as quoted by Phillips, are shown in the last column of table 1.
Although there appear to be significant differences among the entries in the last
column, there is at least an order-of-magnitude agreement among them. However,
no independent theoretical estimate of a has yet been given.
Now the basic idea underlying Phillips’s derivation of the equilibrium law is that
over this range of frequencies the spectrum is limited by wave breaking. It seems
natural then to inquire how much energy is indeed lost by wave breaking in a
spectrum of the form ( 1*1), and to relate this to the energy input by the wind, which
can be estimated independently. This is the purpose of the present note.
independently of the frequency range a 1 < <x < or2, provided cr1 < cr2. On the other
hand, if we assume that the energy lost in breaking is comparable with that supplied
by the wind, we are led to the conclusion (see § 2 ) that, to an order of magnitude
® ~ 1 6 0 0 C W /W ). (1-3)
where G is the ‘drag coefficient’ of the wind on the waves. On assuming the value
С = 1*6 x 10-s and equating ( 1*2 ) and (1-3) we find
a value within the range of observation and moreover extremely insensitive to the
rough assumptions made in the analysis.
The question is raised in §4 whether a is an absolute constant or whether, on the
other hand, it varies throughout the development of a wave spectrum. There is
some evidence (derived from table 1) to suggest that a may in fact be a decreasing
function of the ‘wave age’ as defined by Sverdrup & Munk ( 1947).
2. T h e l o s s o f e n e r g y b y w a v e b r e a k in g
A rough estimate of the energy lost to the breaking waves may be made in the
following way. It is well known that in a standing wave, the maximum downward
acceleration is equal to g (see Taylor 1953 ). It is less well known that in a progressive
wave, whose limiting form at the crest is the Stokes 120 ° angle, the acceleration is
equal to directed away from the crest in all directions (Longuet-Higgins 1963).
1661
® i ь
F ig u b e 1 . The limiting acceleration near the crest in a Stokes 120 ° angle.
I f for a moment we assume that the wave is adequately described by the linear
expression £ = a co b (kx —art) (cr2 = gk), (2 *1)
a<r2 = (2 -2 )
and that the ratio of the wave height to the wavelength would be given by
2a ak acr2 1
-p = — = ---- = ^ - = 0*159 ..., 2-5
L it ng 2л
Now let us suppose that all those waves of amplitude greater than the critical
amplitude a0 will break, where
O'0v 2 = i9 (2*8)
and or is a mean frequency which may be determined from the spectral density
* V )b y Г оо Лео
cr2J l^((7)dcr = J (t2F((t) dcr. (2-9)
I f determined in this way, a has the property that the curve £ = a cos a t has the
same mean number of zero-crossings per unit time as the record with spectrum
F{cr) (see Rice 1944 ). Let us further assume that if the wave amplitude a exceeds
a0, then wave breaking reduces the amplitude to a 0 exactly, so that the amount o f
energy lost is \рд(аг —a§). With these assumptions it follows that the mean loss of
energy density in one wave cycle T = Ял/сг is given by
= f
J а,
exp { —a21а2} А(\рдаг)
( 2 - 10 )
Г a^2o" 3 dcr
and ^ = 7 ^------------- = 2 o f . (2-17)
ocg2a~5dcr
J <Ti
Hence w = ex p {—l / 8a}, (2-18)
a result independent of <гх.
If, on the other hand, F(<r) has a smooth cut-off as suggested by Neumann & Pier-
son ( 1966): F(cr) = ag»<r-‘ exp{-fi(<Tla1)i}, (2-19)
where J3 is an absolute constant (== 0-74) and o’1 = g/U, U being a representative
wind-speed, we find then r<X3 2
j. * № -■ & > (2'20)
3. An e s t im a t e o f а
Some of the energy lost to a wave through breaking will doubtless be transferred
to other parts of the gravity wave spectrum. A small but significant part may also
be transferred to capillary waves (Longuet-Higgins 1963 ). However, the visual
1664
where E denotes the total wave energy density per unit area and T the mean period
as defined in § 2 .
On the other hand since the equilibrium spectrum is limited by wave breaking,
if we assume that the wind puts energy mainly into the longer wave components
that are about to be limited by breaking, it is arguable that the loss of energy through
breaking should be comparable with the rate of growth of the total wave energy:
даE /T ~ dEjdt. (3-2)
Of these two inferences, (3*1) appears the better founded. But if both inferences
are correct, then it follows that
W ~ dE /dt (3-3)
as suggested on other grounds by R. W. Stewart ( 1961). Of course it is always true
that the left-hand side of (3-3) is greater than the right.
To estimate the rate of working of the wind we may adopt more than one method.
Perhaps the simplest is to assume that
W ~ ти+у (3*4)
where r denotes the wind-stress and u* the friction-velocity, defined by
T= Pair *4 ■ (3-6)
Since it is known empirically that
7 —Cpa.lT U2, (3*6)
where U denotes the mean wind-speed at a height 10 m above the sea surface, and
С ~ 1*6 x 10-3, it follows that, on this basis, = CXJ2 and so
W~ Pair < = C}Palr UK (3-7)
Alternatively we may adopt the suggestion of Stewart & Grant ( 1962 ) and as
sume that, since energy and momentum are transferred to a wave in the ratio of the
phase velocity, then W = TCl = Gpalr U \ (3 -8 )
where cxdenotes the phase velocity of the waves that are receiving most of the energy.
The two estimates are the same if it is supposed that
cx ~ C iU ~ 0*04£7. (3*9)
Without making this assumption we shall simply adopt equation (3*7), which is
almost certainly correct to within an order of magnitude.
1665
On substituting into equation (3-10) from equations (3-11), (3-12) and (3-7) we find
that the terms in U and g cancel each other identically, giving
Ш~ 1600CipaIr/p„ter (3-13)
or, with the values С = 1*5 x lO-3 and p ^ lp^ ter 1*3 x 10-3,
1-0x10-* (314)
Having now estimated the numerical value of w we are in a positionto find a from
equation (2*18): .
“ * - 8 Е ^ ф 1 ’35* 10' г- (316)
We notice that this value of a falls within the range of the experimentally deter
mined values of a as shown in table 1. However, we may remark that if any of the
relations between W, E, T and U were to be altered by a factor of 2 this would change
the value of In m by only one part in 12 . Hence the corresponding value of a given
by (3*13) would only vary by this amount and so would still be within the range of
observation.
However, the replacement of equation (2*18) by equation (2 -22 ) would put a
just beyond the upper limit of the observed range.
4. Is CL B E A L L Y CONSTANT?
For the appropriate values of E and T we selected, in the above analysis, the
values given by Sverdrup & Munk (1947 ) for large values of the nondimensional
fetch (gxl U2), say {gxj U2) 104. In a sea state which is still growing under the action
of the wind, smaller values of (gxfU2) would be appropriate and hence smaller
values of both E and T. The variation of T under different wind conditions is gener
ally less marked than the variation of E. Hence w, which is proportional to TIE,
might be expected to decrease with (gx/U)23and so a, by (3*15), would also decrease
with (gx/U2).
This inference gains some support from the data in table 1 . In the second and
third columns are shown the logarithmic-mean values of the fetch x and the wind
1666
The most striking observational conclusion was that as the fetch x increased, so
the peak energy density in the spectrum (which occurs at the low-frequency end of
the equilibrium range) first rose to a value 1-5 to 3 times greater than its final,
equilibrium value. The authors suggest, in fact an equilibrium spectrum of the form
F{or) = ад2ог~6/((т1сг0)} (4*1)
where cr0is the radian frequency of the maximum spectral density. This frequency
diminishes as the fetch is increased. Clearly, if an attempt were made to make a best
fit of the equilibrium law ( 1 -1) to a typical spectrum, the constant a would have to
be adjusted differently according to the fetch, the highest values necessarily occur
ring at the shortest fetches. This is apparently consistent with our conclusions, f
Nevertheless an examination of the original spectra summarized in table 1 shows
that many of them do not exhibit such a spectral peak as found by Barnett &
Wilkerson, even at quite small values of (gx/ll2). We therefore prefer not yet to
adopt any particular form of/(cr/cr0) apart from that implied by equation ( 1 -1).
There is no reason in principal why the calculation of m in terms of F(cr) given in
§ 2 should not be adapted to a function/(cr/<r0) of any reasonable form.
R eferences
B arn ett, T. P. & Wilkerson, J . C. 1967 On the generation of ocean wind waves as inferred
from airborne measurements of fetch-limited spectra. J . Mar. Res. 25, 292-321.
Burling, R. W. 1959 The spectrum of waves a t short fetches. Dtsch. Hydr. Z. 12, 45-64,
96-117.
Cartwright, D. E. & Longuet-Higgins, M. S. 1956 On the distribution of the m axima of a
random function. Proc. Roy. Soc. A 237, 212-232.
Chase, J . et al. 1957 The directional spectrum of a wind generated sea as determined from
d ata obtained by the Stereo W ave Observation Project. New York Univ. CoU. Eng.
Rep. Ju ly 1967.
Hicks, B. L. i 960 The energy spectrum of small wind waves. Univ. Illinois, C.S.L. Rep.
M-91.
Kinsm an, B. i 960 Surface waves a t short fetches and low wind-speed—a field study.
Chesapeake Bay Inst. Rep. no. 19.
Lam b, H . 1932 Hydrodyanmics (6th ed.). Cambridge University Press.
Longuet-Higgins, M. S. 1952 On the statistical distribution of the heights of sea waves.
J . M ar. Res. 11, 245-266.
Longuet-Higgins, M. S. 1963 The generation of capillary waves by steep gravity waves.
J . FI. Mech. 16, 138-159.
Longuet-Higgins, M. S., Cartwright, D. E. & Smith, N. D. 1963 Observations of the direc
tional spectrum of sea waves using the motions of a floating buoy. Ocean wave spectra,
pp. 111-136, Englewood Cliffs, N .Y .: Prentice-Hall, Inc.
Michell, J . H . 1893 The highest waves in water. Phil. Mag. 36, 430-437.
Neumann, F. & Pierson, W. J. 1966 Principles of Physical Oceanography. Englewood Cliffs,
N. J ., Prentice-Hall, 545 pp.
Phillips, О. M. 1958 The equilibrium range in the spectrum of wind-generated waves. J .
F I. Mech. 4, 426-434.
Phillips, О. M. 1966 The dynamics of the upper ocean. Cambridge University Press.
Pierson, W. J . (ed.) i 960 The directional spectrum of a wind-generated sea as determined
from d ata obtained by the Stereo Wave Observation Project. NewYork Univ. CoU. Eng.,
Met. Pap. 2, no. 6.
Rice, S. O. 1944 The m athem atical analysis of random noise. Bell Syst. Tech. J . 23, 282-
332.
Stew art, R. W. 1961 . The wave drag of wind over water. J. FI. Mech. 10, 189-194.
Stew art, R. W. & Grant, H. L. 1962 D etermination of the rate of dissipation of tu rb u
lent energy near the sea surface in the presence of waves. J . Oeophys. Res. 67, 3177—
3180.
Stokes, G. G. 1847 On the theory of oscillatory waves. Trans. Camb. Phil. Soc. 8, 441-455.
Sverdrup, H . U. & Munk, W. H . 1947 Wind, sea and swell: theory of relations for forecasting.
U.S. N avy Hydrogr. Office Pub. no. 601.
Taylor, G. I. 1953 An experimental study of standing waves. Proc. Roy. Soc. A 218, 44-59.
W atters, J. K. 1953 D istribution of height in ocean waves. N .Z . J . Sci. Tech. В 34. 409-
422.
1668
1. Introduction
E xact solutions of the time-dependent equations of motion for a non-viscous
fluid with a free surface are quite rare. One class of motions known since the
last century is the ellipsoids of Dirichlet (1860), in which a rotating, gravitating
mass of fluid, of constant density and vorticity, is contained within a closed
ellipsoidal surface, which deforms continuously with time. The solution of the
three-dimensional problem has been derived concisely by Lamb (1932, §382).
A two-dimensional form of these motions was rediscovered by Taylor (1960)
in the course o f an experimental and theoretical study of the form of a jet issuing
from an elliptical orifice.
More recently the present author (1969), who was interested in the form of
a standing gravity wave of maximum amplitude, inadvertently discovered
a hyperbolic form of the same solution. Unlike the elliptical case, however, the
hyperbolic solution shows the interesting feature of a pressure singularity, or
shock, at the instant when the angle between the asymptotes becomes equal
to a right-angle. Though not necessarily describing the limiting form of a standing
gravity wave, the hyperbolic solution may very well describe an instability near
the wave crest, or the time-dependent behaviour of a jet impinging on a plane
wall.
It is the description of the hyperbolic flow which is the main purpose of the
present paper. Section 2 outlines a general approach to the problem of time-
34 FLM 55
1669
630 M. S. Longuet-Higgins
dependent free-surface flows. The most general two-dimensional solution with
constant rate of strain is derived in § 3. It is seen to depend upon two dimension-
less parameters P and Q, governing respectively the time and length scales of
the motion. The elliptical case (P < 0 ) is described in § 4 , in a form rather
different from that given by Taylor (1960). The description of the hyperbolic
case is given in § 5, followed by a discussion of its applicability to the standing
wave.
The intermediate case (P = 0 ) is also of interest (§ 6 ). This describes a flow
bounded by two parallel plane surfaces, which approach one another. The
intervening fluid is ejected in such a way that any given particle travels outwards
with constant velocity. The flow is related to a well-known open-channel flow
but is exact, not being dependent on the shallow-water approximation or the
hydrostatic assumption. The flow is verified in a simple experiment, described
in § 8 .
^ = - ^ - i W ) 2- F , (2.2)
g , = ! +V № = 0 (2.4)
on the same surface. An arbitrary function of the time, which is sometimes added
to (2 .2 ), is considered as being absorbed into ф.
p (3.7)
A8 A *~ '
a constant, so that A 2 = Л*(1 + РЛ*). (3.8)
3+-2
1671
632 М. 8. Longuet-Higgins
leads to t = A f[ — ^ . — т. (4.4)
Jo cos20 ( 2 - sm 2 d)*
This integral may be expressed in terms of the Legendre elliptic integrals
Лас ^
E (a ,k) = I
(4.5)
F(a k) = Г ____
’ J о (1 —A sin26 )i’
—&;22sii
tabulated, for example, by Byrd & Friedman (1954). In fact
t = 2 W [ t a n a ( l - |s i n 2a ) i - ^ ( a ,2 i) + J P (a ,2 i)]. (4.6)
Using a as a parameter, A may now be plotted as a function of t as in figure 2 ,
for fixed values of M . When t < 0 the solution found by reflexion in the line
t = 0 is analytically continuous with the solution for t > 0 . The limiting case
M = 0 , however, requires special treatment (see below). In general A attains
a maximum value M~x at time t — 0 , as can be seen also from (4.2). When on the
other hand t -> oo then A -+ 0 , and from (4.2) A ~ t~x.
1672
t
. The function A{t) when P < 0 , giving the velocity a t a fixed point as a function
F ig u r e 2
of the tim e t. Only the curves for A > 0 are shown. Those for A < 0 are obtained by
reflexion in the t axis.
634 M. S. Longuet-Higgins
The free surface p = 0 may be written in the usual form for an ellipse:
#2 У2
^2 + £2= 1 ’ (4-8)
where a and b are the lengths of the semi-axes, given by
. QA* , QA*
~ A^+A’ Ь-кгА' <49>
Since A 2 > \A | by (3.13) it is clear that for a real free surface to exist Q must be
positive. Also from (4.9) we have
by the differential equation (3.8). Hence the product of the semi-axes ol the
ellipse is equal to ( —Q /P)iy a constant, as we would expect from continuity,
nab being equal to the cross-sectional area.
We also have
Q A 2- A A2 Q A 2+ A
M* A* ’ b ~ M * A* > ( )
and so if A is expressed in terms of A,
Q 1± (1 - M W ) \ Q 1T(1
M2 M 2A 2 ’ M2 M 2A 2 { >
according a s ^ O . When t = 0, then a and 6 are both equal to Qb/M, and the
cross-section is circular. When t > 0 the ellipse is elongated in the x direction
and when t < 0 it is elongated in the у direction, as shown in figure 3.
It will be seen from (3 . 12) that the pressure p is a maximum at the centre of
the ellipse, at which point the velocity is zero, and that in general the pressure
increases towards the centre. This suggests that the flow may be stable provided
that the fluid is contained within the ellipse. The corresponding flow for an
elliptical cavity, when the fluid is outside the ellipse, is probably unstable.
A further property of the flow is that all surfaces of constant pressure {p = p 0)
are similar to the free surface p = 0 , but that these do not move with the fluid
unless we assume that p 0 is not constant but varies in time like A*.
The pressure at the centre of the ellipse (x, y) = (0 , 0 ) is given by
This was compared by Taylor (I960) with the pressure measured at the centre
of a jet issuing from an elliptical orifice in a vertical wall, with very good agree
ment (see Taylor 1960, figure 5).
Taylor’s derivation of (4.13) differs from the one given here, the time dependence
being expressed in the form of an integral, which is then expanded as an infinite
series. I t may be noted that the infinite series
__ 1 _ 1 1
2. 3 + 8. 7 16.11 + ‘ ” *
1674
2 B № ~ K fa ),
where E and К are complete elliptic integrals.
N A = c o t (P ) (5.3)
F igube 4. The function A when P > 0, giving the velocity a t a fixed point as a function
of t. The curves for A < 0 and t < 0 may be obtained by reflexion in the axes.
1675
536 M. S. Longuet-Higgins
As before, A may now be plotted as a function of t (see figure 4 ). The curves
for A > 0, t > 0 may be extended by reflexion in either the t axis or the A axis.
A new feature, which did not appear in figure 2 , is the presence of a singularity
at t = 0 . In the neighbourhood of this point we have from (6 .2 )
t ~ ± (3N*A*y-\ N A ~ ± (N fit)i. (6 .6)
Thus the velocity (except at x —у = 0 ) becomes infinite like and the pressure
(except at the free surface) like tri. The displacement near t = 0 is like the integral
of the velocity and remains finite.
As we saw earlier, the cross-section of the free surface is a hyperbola, given by
{A + A 2)x2- ( A - A 2)y2 = QA*. (5 .7 )
Since in this case A 2 > A* it follows that the coefficients of x2 and y 2 have the
same signs as A and —A respectively. We may distinguish three cases as follows.
Case 1 : Q > 0
Then if A > 0 the form of the free surface is as shown in figure 5 (a), the shaded
area indicating the region where the pressure is positive. The angle which the
asymptotes make with the x axis, namely
a = arc tan
Case 2 : Q < 0
Then if A > 0 the form of the free surface is as in figure 6 (a), the angle a being
greater than 45 °. The shading shows the region of positive pressure. If A < 0
the configuration is as in figure 6 (b).
Case 3: Q = 0
In this case the free surface reduces to two planes making equal angles +^.m th-.
the x axis. If A > 0 the configuration is as in figure 7 (a), and if
figure 7 (b). j.
1676
* у
(a) №)
F ig u r e 5. The form of th e free surface when P > 0 and Q > 0. (a) A > 0, (6) A < 0.
F i g u r e 6. The form of the free surface when P > 0 and Q < 0. (a) A > 0, (b) A < 0 .
In the hyperbolic solutions just found, both pressure and velocity become
large at infinity. Hence these are essentially local solutions, which may never
theless describe the local behaviour of some realizable flows. For example the
solution of figure 5(a) may describe the time-dependent behaviour of a jet
impinging on a plane surface (y — 0 ). The flow of figure 6 may partially describe
the collapse of a mound of fluid under gravity. In each case the existence of
a singularity at a = 45° is very interesting. It appears that as this angle is
approached a weak shock will occur and that the pressure gradient will change
sign. Thus the configuration of figure 5 (a) will go over into the configuration of
figure 6 (6 ) (and that of figure 6 (a) into that of figure 5(b)). Where the pressure
1677
638 M. S. Longuet-Higgins
F ig u r e 7. T h e form o f the free surface when P > 0 and Q = 0. (a) A > 0, (6) Ac 0.
was previously positive it now becomes negative, and a flow that was stable
becomes suddenly unstable.
It is natural to consider how these solutions are related to the problem of the
limiting form of a standing gravity wave of maximum amplitude. Penney &
Price (1952) showed that at the instant of maximum elevation the acceleration
at the crest of a standing wave is equal to g, directed downwards. Thus the fluid
near the crest is in a state of almost free fall. Penney & Price suggested on
theoretical grounds that the limiting slope at the crest should be 45°. These
authors' arguments were not accepted by Taylor (1953), who nevertheless showed
experimentally that the maximum slope was indeed very close to 45°. However,
the occurrence of instabilities near the crest made the precise observation of the
maximum angle rather difficult.
Since the hyperbolic solution described above has a singularity at t = 0, at
which the velocity becomes infinite, it evidently does not describe the motion in
a smooth standing wave. Nevertheless, since the fluid near the crest is in a state
of almost free fall, the flows of figures 5 (b) and 7 (6) (in the region# < 0) may well
represent possible instabilities that could occur near the instant of maximum
elevation.
(2IQ)*y
I 2 1
/ / / / / yV/ ' /
1 I
/ У / / / Х / / / У / / / sty S / / / / 7
. , L T 2
7 / / / / у / / 7 У г77~
< 0
У / у / у ^ //уу / / / / \ /////'у _
T
S //S //*///, ///A s/ / / / / /
1 —l 1 1
/////A s///, / / / / / / / / / /
1 -2 1 i
The flow is contained between two planes parallel to the x axis, which come
together with velocity (2Q)U~Z (see figure 8 (b)). The pressure on the x axis is
proportional to (r*. Hence as t -> 0 the pressure becomes very large, as a greater
and greater thickness of fluid has to be accelerated outwards.
The second possibility, when t = —1/A > 0 , is essentially the reverse of the
first; the fluid is contained between two lines parallel to the у axis (figure 8 (a)).
A paradoxical feature of the flow described by (6.2)-(6.4) is that although
the horizontal pressure gradient is zero the horizontal velocity и given by ( 3 .2 )
appears at first sight to be decelerating (since А ос 2-1). The paradox is resolved
by recognizing the difference between the Eulerian acceleration du/dt and the
Lagrangian acceleration Du/Dt. In fact any marked particle moves outwards
with constant horizontal velocity и (though with steadily diminishing vertical
1679
640 M. S. Longuet-Higgine
velocity v), so that the horizontal component of the Lagrangian acceleration
vanishes everywhere.
The flow is such that when t is small all particles start from the neighbourhood
of the line x = 0 . As t increases, an observer at a fixed point sees first the swiftest
particles, since they arrive first, then particles with gradually slower and slower
velocities. Hence to a fixed observer the velocity appears to be decreasing in
time. The flow is analogous to an expanding universe.
An experiment to test the reality of this flow will be described in § 8 .
The flow will be recognized as being related to a well-known channel flow,
based on the nonlinear shallow-water theory in which vertical accelerations are
neglected (see, for example, Forchheimer 1930). The present solution is however
exact, and is independent of the shallow-water approximation.
7. Particle trajectories
The Lagrangian description of the class of flows described in this paper is
almost as simple as the Eulerian. Thus let (X, Y) denote the co-ordinates of
a fixed particle, functions of the time t and of the initial positions (X 0,70) of the
particle at time t0. Then corresponding to the flow described by (3.1) we have
dXjdt = u = дф/дх = Ax = A X , |
(7.1)
dY jdt = v = дф/ду = - A y = - A Y , )
and so (7.2)
т п в — А ®-
Hence Х = Х „Щ , T = (7.3)
'1
~
, ft,
a>
И
(7.4)
II
where
г
f
It can be shown directly that an expression of the general form (7.3) satisfies the
Lagrangian equations of continuity and of motion, and that the pressure is
constant on a free surface of elliptical cross-section provided that
FJP = X F -4 \F -')ld P , (7.6)
Л being a constant related to the axes of the ellipse.
The special flow described in § 6 corresponds to
A = 1Jt, F = t j t . (7.6)
The pressure in this case is given by
8. Experimental verification
In order to neutralize the effects of gravity on these free-surface flows there
are at least three experimental alternatives.
( 1) The experiments may be conducted under conditions of free fall. This was
in effect the method adopted by Taylor (I960), who replaced the time variation
by a gradual space variation, in a free nearly two-dimensional jet.
1680
и = VxfxB, (8 . 2 )
where x is the distance from one end of the tank and xB = Ut is the distance
through which the block has travelled. Thus we have и = A x as in (3.2), since
A = 1It.
Any other motion of the plunger will cause a tilting of the free surface-for
example if the plunger is accelerated or decelerated, or if it is suddenly removed
altogether as in the collapse of a dam. Such flows will invite specifically gravita
tional effects.
Figure 10 (plates 1 and 2 ) shows the realization of such an experiment in
a rectangular tank of height 18 in., width lOJin. and total length 12 ft 2 in. The
block, which moves from right to left, is in the form of a trolley running on
rubber wheels. Attached to the rear of the trolley is a vertical sheet of plywood,
1681
642 M. S. Longuet-Higgine
sealed to the floor and side walls of the tank by a thin rubber sheet. Attached to the
front of the trolley is a sponge-rubber buffer and a horizontal steel cable, which
in turn is attached to an electric motor, as in figure 9. When the trolley reaches
the far end of the tank the clutch is automatically released and the trolley comes
to an abrupt halt. The trolley also carried two bricks whose weight helped to
seal the lower edge of the rubber sheet to the floor of the tank.
At the start of the experiment the trolley was held in position by a long pole
(whose shadow on the rear wall is visible to the left of the trolley in figure 10)
while the space to the right of the trolley was filled with dyed water, to a height
of 14in. At the word ‘go’ the pole was released and the clutch of the motor was
engaged. The trolley quickly accelerated to an almost uniform speed of about
1*6 ft/s. The sequence of photographs in figure 10 was taken with a cine camera
running at 64 frames/s. A selection of frames is shown, the serial numbers of
each frame being given on the right.
It can be seen that, at the start of the run, the initial acceleration of the trolley
produces a tilt of the free surface, and hence an incipient wave (figures 10 (a)-(c),
plate 1). But as the motion of the trolley becomes more uniform the wave is
damped out and the surface becomes level to a remarkable degree (figures 10(A)-(j),
plate 2 ).
On reaching the far end of the tank the trolley comes to a halt and water begins
to pile up against the rear face of the trolley (figures 10 (&) and (I), plate 2 ).
Fluid is theD thrown back (figures 10 (m) and (n), plate 2) on top of the horizontal
stream, which continues to flow from right to left and to diminish in depth.
At this stage, and indeed after about figure 10 (e) (plate 1), the horizontal flow is
supercritical, that is to say the velocity и exceeds {gh)b, and hence the oncoming
stream remains unaffected by the return flow. At a later stage (not shown in
figure 10 ) a bore is formed travelling from left to right. Ultimately, viscosity and
turbulence damp the flow, but in the short time corresponding to figures 10 (a)-(n),
the thickness of the boundary layer, given by (vt)i, is of order 1 mm, and so can
be neglected.
In principle it is possible to reverse the experiment and to compress the
fluid by forcing the plunger in the opposite direction (towards the fixed end)
still maintaining a horizontal free surface. But in practice the required initial
flow would be difficult to establish. Any irregularity of the free surface would
cause wave motion, and the horizontal contraction of the fluid by the plunger
would then tend to increase the wave amplitude through work done against the
radiation stresses (see Longuet-Higgins & Stewart 1961; Taylor 1962).
When, on the other hand, the plunger moves away from the fixed end, any
unwanted disturbance of the free surface tends to be reduced.
F io u r e s 10 (a) to (g)
F ig u b e s 10 (Л) to (n)
F igure 10. A model experiment designed to realize the expanding flow described in §6.
Each number on the right indicates the serial number of the corresponding frame. Tho
film was shot at 64 frames per second.
LONGUET-HIGGINS
A class of exact, time-dependent,free-surface flows 643
REFERENCES
B yrd, P . F . & F r i e d m a n , M. D. 1964 Handbook of Elliptic Integrals for Engineers and
Physicists. Springer.
D ir i c h l e t , G. L. I860 Untersuchungen uber ein Problem der H ydrodynam ik. Abh. Kon.
Oest. Wiss. Gottingen, 8, 3-42.
F o r c h h e e m e r , P. 1930 Hydravlik. Leipzig: Teubuer.
L a m b , H . 1932 Hydrodynamics, 6th edn. Cambridge University Press.
L o n g u e t -H i g g i n s , M. S. 1969 The exact hydrodynamical description of a slosh. Proc.
N A T O Advanced Study Institute on Topics in Geophys. Fluid D yn., p. 44. Univ. Coll.
N. Wales, Bangor.
L o n g u e t -H i g g in s , M. S . & S t e w a r t , R. W. 1961 The changes in am plitude of short
gravity waves on steady non-uniform currents. J . Fluid Mech. 10, 529-549.
P e n n e y , W. G. & P r i c e , A. T. 1952 Finite periodic stationary gravity waves in a perfect
fluid. Phil. Trans. Roy. Soc. A 244, 254-284.
T a y l o r , G. I. 1953 An experimental study of standing waves. Proc. Roy. Soc. A 218,
44-59.
T a y l o r , G. I. 1960 Form ation of thin flat sheets of water. Proc. Roy. Soc. A 259, 1-17.
T a y l o r , G. I. 1962 Standing waves on a contracting or expanding current. J . F luid
Mech. 13, 182-192.
1685
Periodicity in Whitecaps
W h il e flying low over the ocean between Adelaide and Kan
garoo Island (36° S, 137° E) one of us (J. S. T.) observed the
following interesting phenomenon.
In a moderate wind of 6-9 m/s from the north, whitecaps
were being formed sporadically, each appearing for about a
second or less and then disappearing, leaving a streak of foam
that would persist for a few seconds. Strikingly, the whitecaps
would generally appear in succession, each making its appear
ance down-wave of the previous one. The interval between the
appearance of successive whitecaps was from about 8 to 10 s.
Another of us (M. D.) has made similar observations from
the deck of a ship steaming westwards in the N. Atlantic
(53° 20' N, 37° 54' W) in a cross-wind of 30 knots and of
nearly constant direction of 350°. Again, whitecaps were
observed to appear in succession, each down-wave of the last,
with a periodicity of about 12.2 s. The period of the larger
waves, on the other hand, was substantially less, being about
8.7 s, as determined visually.
We offer the following explanation. Sea waves are not of
uniform amplitude, but occur in groups of high and low
waves1. The wave envelope progresses with the group velocity
Cf, which for waves in deep water is about half2 the phase
velocity C. Individual wave crests will appear therefore at the
rear of a group, move forward through the group and diminish
again towards the front (Fig. 1). The maximum wave amplitude
will be attained near the centre of the group, at which point
breaking may occur. Now progressive waves tend to break at
a certain limiting value of the wave steepness (=wave ampli
tude/wavelength) for which the downward acceleration of the
crest reaches 1/2^ (see ref. 3). Provided that the wavelength
of the dominant waves is nearly constant, this limiting con
dition will occur at a certain critical value of the wave amplitude,
as indicated-in Fig. 1. At this critical amplitude a whitecap
will appear and will persist until the wave amplitude falls
below the critical amplitude.
1686
-------- -С
Q
2
1687
M. D o n e la n
1. Introduction
The breaking of surface waves has important dynamical consequences in the
ocean, both in deep and in shallow water. In deep water, white-caps may be
responsible for a high proportion of wave energy dissipation (see Stewart &
Grant 1962), and hence for the conversion of wave momentum to larger scales of
motion. Thus if D denotes the mean rate of energy dissipation per unit horizontal
area, and if с is the phase velocity of the breakers, an amount of wave momentum
9 FLU 57
1691
130 M. S. Longuet-Higgins
equal to Djc is available for conversion to larger scales of motion, including mean
currents.
The turbulence due to breaking waves is probably also the chief agent of
vertical mixing and for vertical transfer of heat and momentum in the upper
most layers of the ocean, while the entrainment of air by breaking waves must
influence the transfer of dissolved gases between atmosphere and ocean.
Likewise in shallow water it is known that longshore currents, which transport
large quantities of sand and sediment parallel to the coastline, are generated with
in the surf zone by waves striking the coast at an oblique angle (Galvin 1967).
The mean tangential stress r exerted by the waves is equal to {Djc) sin Q, where
D and с have the same meanings as before and в denotes the angle of incidence
(see Longuet-Higgins 1970a, b). Though most of the dissipation in the surf zone
is due to breaking waves, neither the form of the breaking waves, nor the dis
tribution of turbulence with depth are well known.
Indeed, while the theory of surface waves of small amplitude has been highly
developed (see, for example, Lamb 1932) and much progress has been made
with the theory of weakly nonlinear interactions (Phillips 1966), our know
ledge of breaking waves is surprisingly scanty.
It was shown by Stokes (1880) that a steady progressive wave of limiting
amplitude must have a sharp crest with a discontinuity of 60° in surface slope.
Using this criterion, the limiting form of waves in deep water was determined
by Michell (1893) and in shallow water by McCowan (1894), Davies (1952) and
others. Numerical calculations of unsteady waves have also been successfully
carried out by, for example, Street (1972) up to the point of breaking. But
little, if any work has been done on the analytical description of a gravity
wave after it has broken. The reason is no doubt connected with the fact that
such flows are essentially turbulent, and potential theory is no longer applicable
to the whole flow.
Laboratory studies of breaking waves in shallow water have been published
by Mason (1952), Iverson (1952), Ippen & Kulin (1955), Divoky, Le Мё1ти^ &
Lin (1970), and others. Mason (1952) distinguished two types of breaking wave:
on the one hand ‘ plunging breakers’, in which the wave crest topples forwards
and falls violently onto itself; and on the other hand *spilling breakers’, in
which the free surface becomes unstable near the wave crest and forms a quasi
steady white-cap on the forward slope of the wave. Illustrations of these two types
are given in figures 1 and 2 of Mason (1952).
In contrast, a considerable amount of experimental work has been done on
hydraulic ‘ jumps’ in channel flows; for a comprehensive review see Rajaratnam
(1967). These are essentially shallow-water flows. The flow in the jump has been
explored in its dependence on the Froude number. Evidently some experimental
difficulties remain to be overcome, in particular those arising from the entrain
ment of air at the free surface. Quantitative measurements by Rajaratnam
(1962) indicate that the mean density in a vertical section can fall as low as
0 -8 g/cm3, and possibly lower.
It must be emphasized, however, that breakers are by no means a shallow-
water phenomenon, as is shown by the presence of white-caps in deep water
1692
(Monahan 1971). Generally they seem to occur when the vertical acceleration
approaches the limiting value \g (Longuet-Higgins 1969 a). It is also certain that
the extreme shearing motion near the crests of large-scale waves will sometimes
give rise to a shearing instability.
In deep water there are at least two ways in which waves may attain their
limiting amplitude. The first occurs when the frequency spectrum is narrow. Since
the phase velocity of deep-water waves is twice that of their group velocity,
the amplitude of waves will grow to a maximum and then decay as they pass
through their wave envelope. Near the maximum amplitude they may break.
The second mechanism applies when the shorter waves riding on the backs
of longer waves are forced by the latter to steepen and break near the crests of
the longer waves (Longuet-Higgins 1969 b).
In deep water, white-caps appear generally to last for a shorter time than on
gently sloping beaches, but from visual observation white-caps may be better
classed as ‘ spilling’ than as ‘ plunging’. Some observations of white-caps are
given in a recent note by Donelan, Longuet-Higgins & Turner (1972).
The complete analytical description of breaking waves, and of hydraulic
jumps, presents a challenging task. In the present paper we do not attempt a
complete description of either of these phenomena, but we suggest on the other
hand a possible model for one local feature of such flows, namely the flow near the
forward edge of a spilling breaker or hydraulic jump, where the turbulent flow
meets the more tranquil water (see figure 1). In this region the flow is evidently
turbulent, as elsewhere in the breaker. We suggest that it may be treated
as a turbulent wedge with a certain eddy viscosity whose magnitude is deter
mined by conditions outside the local flow (see figure 2 ). Such a local solution
is developed in the present paper.
In § 2 we show that the only type of smooth irrotational flow which can sustain
a discontinuity of surface gradient is one with an angle of 120° - in fact, a general
ization of the Stokes 120° angle flow (see figure 2). But to sustain this the fluid on
9-2
1693
132 M. S. Longuet-Higgins
the left of the discontinuity must be at rest, and the free surface on the left must
be horizontal. There is no way to support a rise in water level on the left, except
by frictional forces, or (in turbulent motion) by Reynolds stresses.
We therefore assume (in § 3) that the fluid on the left of the discontinuity is
turbulent, and is dominated by Reynolds stresses. The latter are assumed to be
represented by a coefficient N of eddy viscosity. The mean velocity in the turbu
lent zone is assumed small compared to the flow on the right (in a frame of refer
ence travelling with the phase velocity). Across the boundary of separation, both
normal and tangential stresses are balanced, the tangential stresses on the left
being given by a constant coefficient С times the square of the velocity on the
right. At the free surface, both components of stress vanish. Under these con
ditions it is shown that a local two-dimensional solution does exist. Moreover
it has the remarkable property that the inclination of the free surface on the two
sides of the discontinuity depends only upon О, and is independent of both N and
g. The angle of elevation on the left is shown to lie between 10° 54' and 30°; the
angle of depression on the right is between 10° 54' and zero. Various special
cases are discussed in § 4.
In §5 we discuss the effect of a possible difference in density on the two sides of
the interface, caused by the entrainment of air at the point of separation. The
flows described above are generalized so as to include an arbitrary ratio of the
densities, and the effect on the inclination of the free surface is determined.
A comparison with some observations of the flow in hydraulic jumps is made
in§ 6 .
In the discussion in §7 it is suggested that a way may be opened up to treat
the hitherto intractable problem of breaking waves, and more generally, of mixed
turbulent and laminar free-surface flows. In such analyses the present local
1694
2. A frictionless flow
By choosing a reference frame moving with a suitable horizontal velocity we
may reduce the motion to a steady flow. Let polar co-ordinates be taken as in
figure 2 , with the origin at the point of discontinuity and with the line 0 = 0
directed vertically downwards. Let the boundary of the fluid on the right be
given by в = a, that on the left by в = —у, with a possible interface at в = —/?.
Let p and g denote the pressure and gravitational acceleration, and p and p'
the densities on the right and left respectively.
Let us first suppose the flow to be irrotational. Then, in an incompressible fluid,
we seek a stream function \jr such that
V2^ = 0 (2.1)
with boundary conditions
Ф = 0, p = 0, when в = a,
^ = 0, ^continuous, when в = —fi, ■ (2.2)
rjr = 0, p = 0, when в = —у..
We may, or may not, have a discontinuity in дг/г/дв at в ——/3. Following Stokes
(1880) we try f = .4 г” вши(0 - а ) ( - / ? < 0 < a ). (2.3)
This satisfies the first condition at the surface в = ос. To satisfy the second con-
dion we note that in steady frictionless flow
pip —gr cos в + \q2 = constant (2.4)
by Bernoulli's theorem. Since from (2.3)
g2 = n2A*r2n~2 (2.5)
we have p = p(gr cos в — -j- constant. (2.6)
The vanishing of p on в = a then implies that
n= h (2.7)
so that the angle between radial streamlines must be 120°, and further
Л2 ==|^соза, (2.8)
so that if А Ф 0, then cos a > 0 ; in other words, the streamlines must slope
downwards from the origin. This is also clear from the fact that the velocity has
to vanish at the origin, and a particle on a surface of constant pressure can gain
kinetic energy only by going downhill.
If now we try to satisfy the free-surface conditions at 6 = —y by the expres
sion -ф- = Вгп&шп{6+у)} —у < в < —Д we find, by a similar argument that
either у < \тт, which is obviously impossible when 0 < a < £я, or else В = 0,
that is to say the fluid on the left is stationary. Then the free surface on the left
is horizontal (y = ^n).
1695
134 M. S. Longuet-Higgins
Can we satisfy the boundary conditions on the interface в = -ft\ On the left
the pressure is simply the hydrostatic pressure p = p'grcos ft. On the right the
pressure is given by equation (2 .6), which, in view of (2.8), may be written as
p = pgr (cos ft - cos a). (2.9)
Equating p and p' on the two sides of в = —ft we have
p' cos ft = p(cos ft—cos a), (2 .10)
or, since a-f ft = ftf,
{p—p') cos ft = p cos (§7Г - ft). (2 . 11)
This leads immediately to
(p-p')lp = y st& n ft-b (2.12)
and hence tan/? = (Ър-2р')ЦЗр. (2.13)
The two extreme cases are as follows. First, when p' = 0, then
a = ft = £tt, A = ±$gi. (2.14)
This is the Stokes 120° angle (Stokes 1880). Secondly, when p' = p then
a= h P = i”, A = 0. (2.15)
The free surface on the right is then horizontal and no flow takes place. Inter
mediate values of the density ratio p'Jp give flows which have non-zero values of
A and which are in effect generalizations of the Stokes 120° angle flow. All these
have a discontinuity in velocity on the surface 6 = —ft, and all have a horizontal
free surface on the left.
We remark that if a non-zero vorticity (without friction) is allowed in the
flow on either side of the discontinuity, it will not affect the angle of the free sur
face. For ijr is then the sum of a harmonic function plus a solution ijrQof the equa-
tion (2.16)
VVo = -"о>
where o)0is the limiting vorticity near the comer. But since in radial co-ordinates
(2.17)
it follows that the relevant solution of (2.16) will be of order o)0r2, which tends to
zero with r more rapidly than the harmonic function (2.3). Hence the presence of
vorticity may affect the curvature of the free surface near the crest, but not the
limiting angle. For the Stokes 120° angle this was pointed out by Miche (1944).
3. A turbulent flow
The observation that the free surface generally slopes upwards to the left
of the discontinuity means that the flow there must be effected either by friotional
forces or by Reynolds stresses, for without these it would be impossible for a
particle at the free surface to be brought to rest at the origin (see equation (2.4)).
In this section, therefore, we shall seek a solution representing laminar flow
on the right but highly turbulent flow on the left (see figure 2 ). For simplicity we
1696
shall first assume that the densities on the left and on the right are equal, i.e.
Most turbulent flows are difficult to represent satisfactorily, the present being
no exception. Moreover there are some outstanding differences between this and
other types of flow, in that the source of the turbulence may be largely external
to the region considered - in a breaking wave, for example, most of the turbulence
may originate nearer to the wave crest. Secondly, the mean flow near the stagna-
tion point must be relatively small to an observer moving with the phase velo
city of the wave.
These considerations lead us to suggest a very simple model as follows. Let us
suppose that the p« are given in terms of the mean flow by expressions analo
gous to those for ordinary viscous stresses, that is to say
(3.1)
(3.2)
V^ = 0 (3.3)
and we see also that Pfp + gy is the harmonic conjugate of NVhJr, the expression
(PIp+ gy) -f iNVhjr (3.4)
136 M. S. Longuet-Higgins
On в = —у we have the boundary conditions
f = 0, Pm = 0, Pr0 = 0, (3.5)
where now, in radial co-ordinates,
(3.6)
jj,—
(32I)
From the first and third of equations (2.19) it follows that
E = - 30^/sin 38, F = ЗСд/sin 8 (3.22)
and so on substituting in the second of (3.19) we find, if Q Ф 0 ,
cot 3ft—cot 8 = 1/3G. (3.23)
Similarly on substituting in (3.18) and using (3.21) we find
138
F ig u r e 3. Graph of the function f(d) — —£ sin 36 sec 8, giving the drag coefficient С as a
function of the angle S at the apex of the turbulent wedge.
The angles of inclination a ', fi' and y \ and the angle S' of the discontinuity in
F ig u r e 4 .
slope at the separation point, shown as functions of the drag coefficient C.
1700
Finally when 8 = 60° we have the same flow as in figure 5 (a) except that the direc
tion is reversed.
We note that in deriving (3 .22) and (3.23) it was necessary to divide by Q,
so that the solution Q = 0 , hence E = F — 0 was excluded. This is the trivial solu
tion in which the fluid is at rest on both sides of the interface. But we see that this
solution is not continuous with either of the ranges examined above. This remark
serves to emphasize the nonlinearity of the present flows. To establish them we
must pass through some configuration other than a state of rest. Once established,
however, observation suggests that they are at least securlarly stable.
Е = ~ sm
Ш 3o
й оp- ’ J, = sm
S ^оp (5Л)
t It is possible that this flow approximates the flow near the crest of a spilling breaker.
1703
142 М. 8. Longuet-Higgins
С
The angle 8 shown as a function of O, for given values of the density ratio
F ig u h e 7.
p'lp. The dashed line corresponds to p'jp — 0-5. The maximum value of С on this curve is
indicated by a solid circle (•).
i ' 1 ^ co sa * J
II
to
F ig u r e 8. Values of af and y ' for the quasi-static flows (£' < 0 ) , giving the inclination of
the free surface in the laminar and the turbulent zones respectively.
In fact we are interested only in positive values of C, and this limits the relevant
values of 8 to two ranges, (1) when 30° < 8 < 60° and (2) when 60° < 8 < 120°.
In range 1, С can tend to zero for values of 8 given by
sin (£ - J7r)+ 77sin£ = 0. (5.6)
At such values both E and F vanish, by (5.1), so that the fluid on the left is static.
From (5.5) we may verify that in this case у — 90°, hence 8 — 90° —p and (5.6)
reduces to (2 . 11). Thus in the limit as С 0 we recover the flows described in § 2 .
As О is increased, keeping 8 within the range 1, we obtain a set of flows which are
generalizations of the flows in § 2. The tangential stress at the interface no longer
vanishes. It can be seen from figure 7 that these more general flows are continuous
with those corresponding to points on the line 8 — 60° (8' = 0 ) as R ->• 1 (ij -+ 0 )
which correspond to a state of rest in both the laminar and turbulent zones.
For this reason we may call the flows in range 1 ‘ quasi-static’ flows. For
1705
144 М. 8. Longuet-Higgins
such flows we see that 8' (= £ -6 0 °) is negative, so that the free surface is
convex.
The values of a' and y' corresponding to the quasi-static flows are shown in
figures 8 (a) and (b). It will be seen that as R is diminished from unity, keeping С
constant, so the angle of depression a' on the right generally increases, as we
might expect, though the behaviour of y' is more complicated. In the limit
as jR->0 and С -> 0 , we see that a! 30°, so that the flow on the right is the
Stokes 120° angle flow, as we should expect. However, the flow on the left
depends on the way in which С and R each tend to zero. Such non-uniform
convergence is to be expected, since the ratio of the tangential stress to normal
stress in the turbulent fluid depends on the ratio Сpip1; that is, C/R.
Consider now the flows in range 2 (60° < 8 < 120°). From figure 9 it can be seen
that these are continuous with the flows described in §§ 3 and 4. We shall call the
flows in this range ‘ dynamic’. For dynamic flows С may not tend to 0 unless
either 8 -> 60° or 120°, or unless R -> 0 . There is a critical ratio of the densities:
R = 0-5. When R > 0-5, then С becomes arbitrarily large as 8 approaches a
certain value. On the other hand, when R ^ 0-5, С cannot exceed a certain maxi
mum value, depending on R.
When R = 0-5 the maximum value of С occurs at 8 = 116° 30', which is inside
the range 2 . Hence there exists a small range of the density ratio, say
0*5 < R < R0,
for which С has two stationary values. Thus we see that when 0 < R < 0-5 there
are generally two dynamic flows or none, for a given value of C; when 0*5 < R < R0
there are either three dynamic flows or one; and when R0 < R < 1 there is always
just one flow for a given value of C.
An approximate method (not described here) shows that the value of R0 is
roughly £(l + £), where £ = (2/135)2 = 0-00022. Hence R0 = 0-50011. By direct
computation using (5.4) it was found that in fact R0 = 0-500117.... The corre
sponding values of a', /?', y' and 8' are given by
a' = 3°23', ft’ = 56° 37',1
y '= l ° 5 4 ', 8' = 57°42',J
and С has the value 0-26146.
The general values of a! and y' for the dynamic flows are shown in figures
9 (a) and (b). At moderate values of С it appears that, as R is diminished from
unity, so a' and y' both initially increase. For example, if С is of order 0 -1, then
a change of 10 % in the ratio R, from 1-0 to 0-9, will increase a' by about 5° and
will increase y' by about 8°. At very low densities (small values of R) if С also
is sufficiently small, a' must tend to 30°, so that we retrieve the Stokes 120°
angle. At the same time y' lies somewhere between 30° and 60°.
F ig u r e 9. Values of (a) a' and (6) y ' for the dynamic flows (£' > 0), giving the inclination
of the free surface in the laminar and the turbulent zones respectively.
extensively studied (for a review see Rajaratnam 1967). In these, however, the
flow is strongly influenced by the presence of the bottom, especially at high
Froude numbers. Nevertheless, some comparisons may be made. Bakhmeteff &
Matzke (1936) showed that the profile of the hump was dependent on the Froude
number Ft = UJfahJi, where XJ1 and \ denote the mean velocity and water
depth below the jump. From figure 2 of Rajaratnam (1968), based on their data,
zo F L M 57
1707
146 M . S. Longuet-Higgins
tany' У'
1-98 0-59 29°
2-92 0*44 24°
4-09 0-39 21 °
5-53 0-26 15°
8-63 0-23 13°
T able 1. Inclination of the free surface above the toe of a hydraulic jump
in open channel flow.
the inclination of the free surface just above the toe of the jump is given by
table 1. Comparison with figure 9(b) shows that these inclinations, in the range
13-29°, might be expected on the basis of an entrainment coefficient С of order
0*1 and a density ratio lying in the range 0*7 < p’jp < 1-0 .
Air entrainment in hydraulic jumps has been studied by Rajaratnam (1962).
The volume concentration c, as defined by him, is approximately equal to 1 —p'lp.
Over a comparable range of Froude numbers (2-42 < Fx ^ 8 -12) he found that,
for points near the upper surface, с lay generally between 0-5 and 0-20. No mea
surements very close to the toe of the jump are reported, but the trend of the
observations suggested that the air concentration there was even higher. Thus,
the density ratio p'/p was certainly sometimes as low as 0*8 and may well have
been less. The range of densities inferred earlier is therefore reasonable.
There appear to be no careful observations of the lower interface of the air-
entrained flow. Rajaratnam (1967, p. 218) states that the angle of depression of
the entrained bubbles increases with the Froude number. This may be due
partly to the increase in the mean flow velocity, relative to the rate at which
bubbles rise to the surface. Generally, interfacial angles indicated by air-bubble
entrainment will tend to underestimate the actual angle of depression. Again,
the presence of the bottom will tend to diminish the apparent angle of depression,
except perhaps very close to the toe, and to reduce the range of validity of the
local solution.
Immediately upstream of the toe it has been generally assumed, on the basis
of nonlinear hydrostatic theory, that the free surface is exactly horizontal.
However, in the neighbourhood of a discontinuity of surface elevation the hydro
static assumption does not necessarily apply. Figure 9 (a) suggests that close to
the toe the free surface will be inclined to the horizontal at an angle at! lying
between about 10° and 20°. Such small angles may have been overlooked.
REFERENCES
B akhm eteff, B. A. & M a t z k e , A. E. 1936 The hydraulic jump in terms of dynamic
similarity. Trane. A.S.G.E. 101, 630-680.
D a v i e s , Т. V. 1962 Symmetrical, finite amplitude gravity waves. In Gravity Waves,
pp. 65-60. U.S. Nat. Bur. Standards. Circular no. 521.
Divoky, D .,L e M £ h a u t £ , B. & L i n , A. 1970 Breaking waves on gentle slopes. J. Geophys.
Res. 75,1681-1692.
D o n e l a n , М., L o n g u e t - H i g g i n s , M. S. & T u r n e r , J. S. 1972 Periodicity in white-caps.
Nature, 239, 449-451.
E , Т. H. & T u r n e r , J. S.
l l is o n 1959 Turbulent entrainment in stratified flows. J.
Fluid Mech. 6 , 423-448.
G a l v in ,C. J. 1967 Longshore current velocity: A review of theory and data. Rev. Geo-
phys. 5, 287—304.
Ipfen , A. T. & K u l i n , G. 1965 Shoaling and breaking characteristics of the solitary
wave. M .I.T . Hydrodynamics Lab. Rep. no. 15.
I v e r s o n , H. W . 1952 Laboratory study of breakers. In Gravity Waves, pp. 9-32. U.S. Nat.
Bur. Standards. Circular no. 521.
L a m b , H . 1932 Hydrodynamics, 6th edn. Cambridge University Press.
L o n g u e t - H i g g i n s , M. S. 1969a On wave breaking and the equilibrium spectrum of wind-
generated waves. Proc. Roy. Soc. A 310, 151-159.
L o n g u e t -H i g g i n s , M. S. 19696 A non-linear mechanism for the generation of sea waves.
Proc. Roy. Soc. A 311, 371-309.
L o n g u e t - H i g g i n s , M. S. 1970a Longshore currents generated by obliquely incident sea
waves, 1 . J. Geophys. Res. 75, 6778-6789.
L o n g u e t - H i g g i n s , M. S. 19706 Longshore currents generated by obliquely incident
sea waves, 2. J. Geophys. Res. 75, 6790-6801.
M c C o w a n , J. 1894 On the highest wave of permanent type. Phil. Mag. 39, 361-359.
xo-a
148 M . S. Longuet-Higgins
M ason, M. A. 1952 Some observations of breaking waves. In Gravity Waves, pp. 215-220.
U.S. Nat. Bur. Standards. Circular No. 521.
Miche, R. 1944 Mouvements ondulatories de la mer en profondeur constante ou d6crois-
sante. Ann. Ponts et Chaussees, 114, 25-87, 131-164, 270-292, 396-406.
M i c h e l l , J. H. 1893 The highest waves in water. Phil. Mag. 36, 430-437.
M o n a h a n , E. C. 1971 Oceanic whitecaps. J. Phys. Oceanogr. 1, 139-144.
P h i l l i p s , О. M. 1966 The Dynamics of the Upper Ocean. Cambridge University Press.
R a j a r a t n a m , N. 1962 An experimental study o f the air entrainment characteristics o f
hydraulic jump. J. Inst. Engng India, 42, 247-273.
R a j a r a t n a m , N. 1 9 6 7 Hydraulic jumps. Advances in Hydroscience, vol. 4 , pp. 197-280.
Academic.
R a j a r a t n a m , N. 1968 Profile of the hydraulic jump. Proc. A.S.C.E. H Y 3, 663-673.
S t e w a r t , R . W . & G r a n t , H. L. 1962 Determination of the rate of dissipation of turbu
lent energy near the sea surface in the presence of waves. J. Geophys. Res. 67, 3177-
3180.
S t r e e t , R. L. 1972 Shoaling of finite-amplitude waves on plane beaches. Proc. 12th
Conf. on Coastal Engng, pp. 345-361. New York: A.S.C.E.
S t o k e s , G. G. 1880 On t h e t h e o r y o f o s c i ll a t o r y w a v e s , Appendix B. Math. & Phys.
Papers, 1, pp. 225-228. Cambridge University Press.
T u r n e r , J. S. 1969 Buoyant plumes and thermals. Ann. Rev. Fluid Dyn. 1, 29-44.
1710
F ILM REVIEW
1. Introduction
The physical importance of the phenomenon o f wave breaking, both in shallow
and in deep water, has been discussed in a recent review (Longuet-Higgins 19736).
I PLм 63
1712
F i g u r e 1 . Sketch showing the features of a spilling breaker which are incorporated in the
theoretical model. The wave is moving from right to left and has a whitecap on its forward
face. The velocities in both the wave and whitecap are measured relative to the wave
crest, with positive direction downwards.
properties of the flow some distance behind the front, using the balance of forces
at a fixed cross-section in the manner proposed for turbulent gravity currents
by Ellison & Turner (1959). Later (in § 6) it will be shown how the advance of
the front can be described in a way which is consistent with this steady plume
model. Common to both Longuet-Higgins’s and Ellison & Turner’s theories,
however, is the assumption that there is a tangential stress at the boundary
between the turbulent and laminar flows, due to the entrainment across this
boundary
The several elements to be incorporated in the theory of §2 are shown in
figure 1. In co-ordinates moving with the phase velocity the flow at the surface
of the waves is directed upwards towards the crest, with velocity decreasing to
zero at the corner. On top of this is the whitecap, with gravity tending to drive
it down the slope. Entrainment into this turbulent flow will add both water
from below and air from above (or at the front). Mixing into the whitecap will
be driven by turbulence produced by shear across the interface, but it will be
inhibited by the fact that the turbulent air-water mixture in the whitecap is
lighter than the water below which it entrains. The rate at which water is in
corporated will depend on the velocity difference u0 = и' —и between the white
cap and the wave, on the density difference p —p' (wherep' is the mean density
o f the air-water mixture) and on the local length scale, say the depth 8. Following
Ellison & Turner (1959) we suppose that the entrainment rate is a function of
the overall Richardson number
Bi0 = g'8 cos вlu%, ( 1)
where g' = g(p—p')lp and в is the slope. Rather less is known about the entrain
ment of air, but in § 3 we discuss laboratory measurements of air concentrations
in hydraulic jumps (see, for example, Rajaratnam 1962) and in self-aerated flows
on steep slopes (e.g. Straub & Anderson 1958) which allow us to put some limits
on p'.
A similarity solution describing the development of the whitecap with dis
tance from the crest is obtained in § 4. This must be an oversimplification of the
actual unsteady flow, but it exhibits very clearly the most important physical
features. The problem is unusual in that entrainment is both the driving and
retarding mechanism: driving because it provides an increasing flux of water to
accelerate the whitecap down the slope of the wave, and retarding because the
i«a
1714
There are certain differences between the present flow and the turbulent
gravity currents considered previously which should be pointed out immediately.
The density difference which drives the flow is now not the same as that which
inhibits the entrainment. The full density difference between air and the air-
water mixture in the breaker provides the driving buoyancy force, whereas it
is the smaller density difference between the air-water mixture and the water
below it which must be used when calculating entrainment. The entrainment
assumption is summarized in the equation of continuity
d(Su')lds = Е\ий\, (5)
which states that the downslope mass flux is increased by entrainment at a rate
proportional to the velocity difference between the whitecap and the wave
slope below. The factor E is a function E(Ri0) of the overall Richardson number
defined by (1), whose form is known from laboratory experiments. Implied by
the use of Ri0 as the only non-dimensional parameter involving the density
difference is the Boussinesq approximation, which is used in all that follows.
1715
where the Boussinesq approximation is used, and g rather than g' appears in the
driving terms. On the left is the total rate of change of momentum flux. The
first term on the right arises from the entrainment of fluid with velocity и at
a rate Eu0, the second represents the component of gravity accelerating fluid
down the slope, and the last expresses the pressure force on the layer due to
its changing depth, the pressure gradient being assumed quasi-hydrostatic, equal
to p'g cos в normal to the surface. There is no solid boundary, and hence no other
drag term. and S2are profile constants, which are required to allow for arbitrary
density profiles. They are assumed to be independent of s and are defined as in
Ellison & Turner (1959) by
SlP'S* = j^2p^ndn (7)
which is the form used in the subsequent analysis. This differs from the momentum
equation used by Ellison & Turner through the addition of the term in du/ds,
which was zero in the previous application, where the external flow was uniform.
1716
0 2 4 6
и (m/s)
Slope (deg)
F i g u r e 3. The mean density of self-aerated flows measured as a function of slope, (a) shows
the result of averaging over a lower region consisting of air bubbles suspended in turbulent
water, and (6) includes a more diffuse region of spray droplets above this. (After Straub &
Anderson 1968, figure 6.) (c) represents our theoretical estimate of the minimum density
differences required to produce a self-sustaining whitecap at various slopes.
1718
where h is the depth, U the mean velocity, и* the friction velocity, cr the surface
tension and v the kinematic viscosity. Thus the ‘ inception number’ I must
exceed some critical value Jc (of order 50 when the turbulent boundary layer
has reached the surface) before air can be trapped at the surface in this way.
For a breaker, this means that the flow must be deep enough and fast enough,
or in terms o f the theory in § 4 (which implies that I oc si), must have progressed
far enough from the crest, before this mechanism can operate. In the early stages
of all breakers, over-running at the front will be the only way in which air is
trapped, but it is possible that at a later stage whitecaps on large waves can be
self-sustaining while those on small waves cannot.
making the same quasi-hydrostatic assumption for the pressure as before. When
the last term, which allows for the effect of the whitecap on the surface pressure,
is neglected, then
и2 = 2gs sin в, (11л)
and we obtain an explicit value for the constant of proportionality in the velocity
relation.
1719
where к, U and U' are dimensionless functions. (It is worth noting that these
imply an ‘ eddy viscosity’ proportional to Succ ghi. Also if all velocities are
proportional to one another, then the ‘ inception number’ I defined by (10)
behaves like si.)
Substituting (12) in (5) gives
d(gh*kU')lds = E g h i(U '-U )
, 2 U '-U
k=- E. (13)
U'
From the definition (1) we obtain
g(Ap/p) ks cos в
BL =
0_ (U'-Ufgs
D. 2/Д,p\ E совв
ОГ 0 3 V/э / £/)(' ( *
Thus for particular values of U' and U (which specify the magnitudes of the
velocities in the whitecap and in the fluid immediately below it), the turbulent
layer thickens linearly with s, and the Richardson number remains constant.
We are therefore dealing with a kind of ‘ normal ’ flow, whose gross properties
can be described in the same terms at all s, in spite of the fact that the flow is
accelerating uniformly down the slope. Note that the linear spread implies (from
(11)) that the same form of velocity variation at the surface of the wave will
hold even when the pressure effect of the whitecap is taken into account.
The momentum equation (9) becomes, using (12) and (13) and simplifying,
Calculations have been made for three slopes, 10°, 20° and 30°, and for three
density differences at each slope. The values of the profile functions Sx and S2
1720
0 04 0-8 1-2
(о) -U
(Ъ) -U (с) -U
-U
F ig u r e 5 . T h e r a t io U'I\U\, a s a fu n c t io n o f U, f o r t h r e e s lo p e s a n d s e v e r a l d e n s i t y
d i f f e r e n c e s , c a l c u l a t e d f r o m f ig u r e 4 .
—U — XJ
which will be able to propagate more easily than it could initially. The values of
the non-dimensional surface velocity |J7|, corrected in this way, are plotted
on figures 4(a), (6) and (c) at the values of U' derived in the earlier calculations.
The crossing points (marked by bold dots) of these and the original curves thus
represent possible solutions for a downslope flow, including its effect on the velocity
of the opposing flow which it is entraining. On a slope of 30° (figure 4 a) the effect
is to increase the downslope velocity to 0-21 times the undisturbed surface
velocity (this ratio depends very little on Apjp). For 10° (figure 4c) the effect
is more dramatic, since solutions now become possible where none existed
before (with a downslope velocity 0 *16 times that at the surface of the undisturbed
wave). A possible implication is that a finite disturbance, due perhaps to a breaker
spilling down from the higher slope above, may allow the flow to go further than
would be expected from the argument based on the marginal stability of an
undisturbed flow.
deeper penetration
at some fraction o f the layer velocity at the same position. The flow in the layer
behind the front is just the same as that in a completely steady plume extending
to infinity in the s direction (which has been implied in all the previous discussion).
In a ‘ starting plume’ , however, this flow feeds momentum and buoyancy to
the front at just the rate required to maintain similarity; but because of the
accumulation of fluid at the front, the velocity o f advance is reduced.
For both axisymmetric and two-dimensional starting plumes in neutral sur
roundings (i.e. with a constant buoyancy flux) the above picture has been shown
to be consistent with the detailed equations of motion of the front. That is, if
the region is regarded as a turbulent ‘ thermal’ , which is mixing with its sur
roundings and whose properties are also being changed by the injection o f
buoyant fluid from behind, then the deduced behaviour of the velocity and
radius does indeed match that in the plume. What we must show now is that the
same matching can be carried out for a two-dimensional plume whose properties
are changing according to the similarity solution of § 4.
The model to be discussed is sketched in figure 7. The steady, established part
o f the whitecap is entraining water from below and air from above, at such a rate
that the density of the air-water mixture remains constant. This also implies,
according to the similarity solution, that the mass and buoyancy fluxes are
increasing down the slope, with 8 oc s and u' oc ght. Thus the whitecap is
accelerating uniformly down the slope, and the position of a fluid particle as
a function of time t is given by s oc t2. At the front the layer will feed fluid into
a circulating region which is like one half of a vortex pair (cf. the two-dimensional
starting plume). Extra entrainment of both air and water will certainly occur
here, and the penetration of the surface of the wave below may well be greater
than it is in the steady part of the flow.
The evidence from experiments on hydraulic jumps (§3) has suggested that
the net effect will be to keep the density of the mixture near the front constant
(at a given Froude number); in fact this result applies more particularly to the
front, rather than to the flow behind. If we now assume that g' (or g) is the major
1725
In the case of the similarity solution with constant acceleration this becomes
dPjdtcct4. (18)
A second contribution comes from the flux o f momentum from the layer behind
i l f o c ^ o c i 4. (20)
Since (18), (19) and (20 ) have the same t dependence, the momentum equation
expressing the balance between these terms will give a consistent description of
the motion of the front.
The starting plume model, therefore, provides some justification for the kind
of theoretical model we have used to describe a whitecap. If the similarity
solution is a good approximation to the real, accelerating flow developing from
a small disturbance at wave crest, then we would expect the front also to advance
initially with constant acceleration. The velocity of the front could, however,
be less than that predicted in figure 4.
F ig u r e 9. The measured horizontal position of the crest X , and the horizontal distance £
between the toe of the whitecap and the smoothed position of the crest, plotted for the
same film sequence as figure 8. The time intervals marked along the bottom are those used
to obtain the average values given in the table.
the toe relative to the wave crest. We also measured the maximum thickness $
o f whitecap on each frame, and occasionally the angle of the wedge, approxi
mately $/l.
In figure 8 are plotted the quantities Z, в and rj = I sin0, as functions of the
frame number. It can be seen at once that, apart from a general increase in I
and q with time, there is also a marked intermittency. The observations showed
that after initiation of the whitecap there was a preliminary period of growth,
after which the wave crest would become more rounded. A part of the whitecap
would then be dragged over the crest, which then became sharp again. The reasons
for this behaviour are discussed in § 8 . Sometimes two wave ‘ crests’ could be
identified on a single frame and in this case the measurements of both are
plotted.
However, most of the intermittency in I shown in figure 8 is due to the oscilla
tion in the horizontal position of the crest, not the toe. This is shown in figure 9,
where we have plotted the horizontal co-ordinate | of the toe relative to the
smoothed position of the crest. It will be seen that, after an initial jump, | in
creases much more smoothly than f (note the change in scale). A parabola,
representing a uniform acceleration, can be drawn to fit the points reasonably
2 FLM 63
1728
t
(frames
from
initiation H V
Interval of spill) R в (cm) (cm) (d lfd tyu
1 13 0078 35° 29*7 5-3 016
2 31 0187 215° 27-8 5-8 0-31
3 64 0-326 17-5° 24-6 8-2 0-42
T able 1. Summarizing the mean properties of a spilling breaker as measured from the
film at three times after the initiation of a whitecap.
9. Conclusion
We have shown how several features of a spilling breaker can be described
using a model which regards the whitecap as a turbulent plume, running down
the forward slope of the wave and entraining the laminar fluid below it. In
particular, the predicted sensitive dependence on the slope, and the magnitude
of the downslope velocity, are in agreement with laboratory observations of
spilling breakers.
Our explicit similarity solutions, however, are based on the assumption of
2-2
1730
Reprinted from Tenth Symposium on Naval Hydrodynamics, Hydrodynamics for Safety Fundamental
Hydrodynamics, ACR-204, Office o f Naval Research, Department o f the Navy
M. S. Longuet-Higgins
Department o f Applied Mathematics and Theoretical Physics,
Silver Street, Cambridge, England and Institute o f
Oceanographic Sciences, Wormley, Surrey, England.
597
1732
Method 1 ; Storm-building. Рог then the wave energy will tend to con
small-amplitude waves in deep water we verge in the neighbourhood of a single
know that the group-velocity C* (the pro point OC , building up a "storm" in
pagation velocity for the wave energy) which breaking waves can occur. A deep-
is given by water breaking wave produced in this
manner is shown in Figure 3.
(1 )
л A»-
(where С is the phase velocity). The
longer waves (with lower O ' ) are there
fore progagated faster than the shorter
waves (with higher O ' ). Suppose that
at time f ; » the wave maker generates a
train of waves of frequency O ', , and Fig. 3 A breaking wave in relatively
that at time this is followed by a deep water produced in a channel 1 5 ш
train of waves of frequency сГ, > о -» long and 50 cm deep, at a point 7 m from
(see Figure 2). the wavemaker, by decelerating the wave
maker. (Method 1).
dt u>
598
1733
which eouala zero at the front itself; Figure 6 shows five records taken
and F (X j describes the wave envelope, at the same poljnt , with different wave
in fact amplitudes. As the amplitude is in
creased, so the arrival of the wave
front is advamced. The bottom record,
with the largest amplitude shows a sharp-
'•> pointed crest typical of a wave of maxi
mum amplitude.
(see for example ^3 )). This function is
related to the Fresnel integral ( k ) . A A film record of the advancing wave
compressed sketch of the surface eleva train shows the occurrence of breaking
tion С is shown in Figure U,. waves at the head of the wave train. 1
Very striking is the fact that the break
ers occur once every two wave cycles.
This is due to the relative motion of the
waves and their envelope. The waves tend
to break when a crest passes the position
of a maximum in the wave envelope. But
the speed of a crest relative to the
velope is only tJ-с (since the
itself travels with velocity С
JL denotes the wavelength, then
■T time interval between the occurrence of
two successive breakers is
599
1734
Fig. 6 Records of the wave amplitude at a fixed point X = 160 ft, for waves
of the same frequency but different amplitude, showing effects of nonlinearity.
600
1735
image ship moving parallel to the first For practical reasons, the experi
The wave pattern due to the image ship ment was transferred to the No. 2 towing
then reinforced the first so that the tank at the National Physical Laboratory,
wave amplitude near the embankment was where moreover the experiment could be
easily increased beyond the breaking repeated under controlled conditions.
point. Moreover the wave crests became Figure 8 shows the bow-wave produced by
almost normal to the embankment, making a model ship towed at a speed of 2 m/s
the flow nearly two-dimensional. parallel to the wall, and at a distance
Fig. 8 The bow wave produced by a model ship towed at a speed of 2.0 т/в in
the No. 2 towing tank at the National Physical Laboratory. Teddington.
601
1736
602
1737
of 1 m from it. The velocity field was window is essential. Figure 10 shows a
measured with ал electromagnetic flow
meter capable of recording two compon
ents of flow, in this case the vertical
component and the horizontal oomponent
in the direction of the motion of the
carriage. The response time of the
flowmeter was of order 0 . 0 5 sec so that
fluctuations with frequency less than
about 20 c/s were measurable. By sub
tracting out the constant velocity of
the carriage it was possible to con
struct a picture of the mean velocity
field (l*,\r) relative to a stationary
observer and the fluctuating components
{и/,-**') . Because of cavitation round
the instrument and its support, it was
not possible to measure the velocity
very close to the free surface.
The mean velocity field is shown
in Figure 9a. Evidently this resembles
the particle velocity field in a non
breaking, irrotational wave, that is to
say upwards on the forward face and
downwards at the rear. These velocities
are of course much less than the steady
carriage velocity, also shown in Figure
9a.
603
1738
605
(Page 606 blank)
1740
By M I C H A E L S. L O N G U E T -H IG G IN S
Department of Applied Mathematics and Theoretical Physics, Silver Street, Cambridge
and Institute of Oceanographic Sciences, Wormley, Godalming, England
A simple derivation is given o f the parabolic flow first described by John (1963)
in semi-Lagrangian form. It is shown that the scale of the flow decreases like
J-3, and the free surface contracts about a point which lies one-third of the way
from the vertex o f the parabola to the focus.
The flow is an exact limiting form of either a Dirichlet ellipse or hyperbola, as
the time t tends to infinity.
Two other self-similar flows, in three dimensions, are derived. In one, the free
surface is a paraboloid o f revolution, which contracts like t~z about a point lying
one-quarter the distance from the vertex to the focus. In the other, the flow is
non-axisymmetric, and the free surface contracts like t~b.
The parabolic flow is shown to be one of a general class o f self-similar flows in
the plane, described by rational functions of degree n. The parabola corresponds
to n — 2 . When n = 3 there are two new flows. In one of these the scale varies
as fV and the free surface has the appearance of a trough filling up. In the other,
the free surface resembles flow round the end of a rigid wall; the scale varies
as f-4'17.
1. Introduction
The review of free-surface flows by Gilbarg (1960) convincingly emphasizes
the scarcity of known, exact solutions to time-dependent flows with a free
surface, particularly when gravitational terms are included. Among the known
solutions are the accelerated cavities of von K&rman (1949) and Gilbarg (1962),
which have also been extended by Yih (1960); the similarity flows for the impact
of a cone on a free surface (Garabedian 1953), and some interesting examples o f
free-surface flows derived by an inverse method due to John (1953). Not men
tioned by Gilbarg (1960) are the long-standing exact solutions known as
Dirichlet ellipsoids (Lamb 1932, p. 382), which were first derived by Dirichlet in
1860, and partly rediscovered by John (1952) and Taylor (1960). A hyperbolic
form of this solution was recently discussed by Longuet-Higgins (1972) in relation
to the appearance of instabilities at the crest of a standing wave.
Among the very simplest of non-trivial solutions to the time-dependent
problem is the flow derived in §2 below. In this, the free surface takes the form
of a parabola, whose linear dimensions vary as irz (where t denotes the time) and
which therefore reduces to a thin sheet as t -> oo. This flow was in fact discovered
by John (1953), in whose treatment, however, the nature of the solution is some-
1741
604 М. 8 . Longuet-Higgins
what hidden by the inclusion of gravity in a non-essential way (see § 7 , below).
When viewed in the natural frame of reference it becomes clear that the flow is
self-similar, and that the parabolic surface contracts (or expands) about a fixed
point one-third of the distance from the vertex to the focus.
The purpose of the present paper is to investigate the possibility of other flows
o f a similar kind. In § 3 we derive two flows in three dimensions, in which the free
surface contracts like t~2and t~5respectively. The possible connexion of all these
flows with the Dirichlet ellipsoids is discussed in §4, and in §§6 and 6 it is shown
that the parabolic flow of § 2 is in fact an exact asymptotic form of the two-
dimensional Dirichlet hyperbola. The analysis is given in some detail because of
a possible future application to the theory of slender breaking waves (Longuet-
Higgins & Cokelet 1976).
In the second part of the paper (§§7-9) the parabolic solution is generalized
in another direction, namely to higher-order rational flows in a plane. For this
purpose, the semi-Lagrangian formulation of John (1953) proves particularly
useful. We show that for each positive integer n there exist two classes of self-
similar, time-dependent flows, with a free parameter Л. By an appropriate choice
o f Л one can generally exclude certain singularities in the flow. The case n = 2
yields the aforementioned parabolic flow and another self-similar flow confined
to the outside of a parabolic surface. The case n = 3 gives rise to a solution
representing a fluid filling up a trough in an otherwise plane surface, and another
solution representing a free-surface flow round the end of a solid wall.
АД the flows discussed in this paper are gravity-free. The paper is intended to
prepare the ground for a future study of time-dependent flows incorporating
gravity in an essential way.
ф - % ^ - У г) + Р ^ , (2-1)
where (x , y) are rectangular co-ordinates, t is the time and P and Л are constants
to be determined. Taking the density as unity, we find that the pressure p , from
Bernoulli's equation, is given by
- ? = |a+ ( l - A ) P ^ i + ^ + / , (2-2)
where / is a function of the time only. Hence the rate of change of p following
a fixed particle is given by
At the free surface both p and DpJDt must vanish. The vanishing of (2 .2 ) and
(2 .3 ) will represent the same surface provided the coefficients of corresponding
1742
dt t t* ’
whence we have
/ ~ ! ? 4 <»>
where Q is an arbitrary constant. The velocity potential is now
Ф= ^ г- У г)+ Р % ( 2 .6 )
and the free surface is
y2 = 3P b l v - % (2-6)
where P and Q are both arbitrary constants.
To fix the ideas, let us suppose P < 0 and t > 0 . Then setting
a = -fP /^ , с = Q/3P (2.7)
the equation o f the free surface becomes
y2 = —4a{x —\cc —ct). (2 .8 )
t Note that in equation (6.4) of that paper, the last term should be <-1.
1743
606 М. 8. Longuet-Higgine
following a particle. The surface always has the form of a parabola, parallel to
the surface corresponding to Q = 0 , but travelling to the right with constant
velocity c. To produce this flow we simply reduce the pressure at infinity by the
amount Q/t4.
The velocity normal to the median plane of the jet is
Фу = ~ у11>
which is always independent of x. It follows that any line of particles parallel to
the median plane always remains so, and that the flow may be realized by the
ejection of fluid from between two approaching parallel plates. A more interesting
realization is likely to be in the jet of water formed by a ‘ plunging breaker*
during the short time that the jet is thin and almost horizontal (see Longuet-
Higgins & Cokelet 1976).
1744
^ = i [ m ( m + l ) V + » ( * + l ) s*s] + A ( l —A ) ^ + ( 2 A - 1 ) ^ j — (3.4)
Fio x j b e 2 . Sketch of the free surface in the axisymmetric flow given by ( 3 .6 ) . The surface
contracts about the point 0 , lying one-quarter of the distance from V to the focus i^.The
dimensions vary as t~l.
1745
608 М. 8. Longuet-Higgins
and comparing coefficients of y2 and x we have
A = 2(m+1) = 3,
and from the terms independent of x and у
1 6 1* + t* •
Hence we obtain for the velocity potential
<*■»>
and for the free surface
у2 + 22 = - Щ Х- ± Р - а ) , (3.6)
F i g u r e 3. Sketch o f the free surface in the flow described by (3.8). Though the flow is
three-dimensional, the surface is a parabolic cylinder. The surface contracts about a line
through 0 , lying two-fifths of the distance from the vertex to the focal line. The dimensions
vary as t~*.
1746
Ф = ± ( * г+Уг- Ь * ) + Р% (3.8)
and the free surface is
z2 = - 4 y(x - % y - c t ) , (3.9)
where у = — с = Q/5P. (3 . 10 )
Thus the flow is the sum of an axisymmetric flow and o f a uniform translation
in the x direction. The free surface is a parabolic cylinder (see figure 3) whose
dimensions vary as J-5. The surface expands or contracts about a line which lies
| of the distance between the vertex and the focal line.
610 M .S . Longuet-Higgins
In the following sections we shall study a special two-dimensional case which
in fact corresponds to the limit c ^ c o with b2 replaced by —b2. This analysis is
given fully, since it may have later application to the theory of breaking waves
(Longuet-Higgins & Cokelet 1976).
f 00 dA
А * ( 1 + № А * ) 1 ’ ( 5 ,3 )
a ^ l + a1)*' (6 '61
f In equation (4.6) of that reference, the modulus of the elliptic integral should be 2“^
as in equation (5.6).
1748
F ig u r e 4 . Graphs of the functions F , dFjdr and d^F/dr*, which give the particle
displacement, velocity and accelerations as functions of the time.
Since (6.8)
J > -
we have also F = {(1 + a4)*+ !}*/«. (6.9)
The particle velocity x and acceleration x are proportional respectively to
dFjdr and d2Fjdr2. The quantities F, dFjdr and d2F/dr2 are plotted in figure 4
above.
From (5.2), the free surface may be expressed as
s 2/a2-$, 2/ 62 = 1, (6.10)
R a R w/4
where
(6 .И )
R a
b=
N {(1 + a4)i+ l}i
r
F ig u r e 6. Graph of the angle 2 y between the asymptotes of the hyperbola,
as a function of the dimensionless time r.
where = Ь* = Ъ Щ 2 ^ (fU 0)
614 М . 8. Longuet-Higgins
7. A semi-Lagrangian method
John (1963) has given a general method for finding two-dimensional, time-
dependent, irrotational flows with a free surface, as follows. Set x + iy = f, and
let us seek solutions in the form
S = £(M), (7.1)
where a) is a Lagrangian co-ordinate (constant following a given particle) and
£ is an analytic function of a). Thus the particle velocity and acceleration are
respectively and The pressure gradient is in the direction of the vector
(£tt + ig)> where g denotes gravity.
On the free surface oj is assumed to be real, so the tangent is in the direction
of the vector Hence the free-surface condition may be written as
£ t t '+ *9 = ir '(*>» 0 L * w r e a l> ( 7 -2 )
(7.6)
where n is any positive integer, and c0, ..., cn and A are constants to be determined.
On substituting into (7.2), taking g = 0 and r independent of (o, and equating
coefficients of wn_1, o)n~\ ..., we obtain
A(A + 1) 6! = nc0rtx+2t y
2 A(2 A + l)c a = (w—l)c 1riA+2,
3A(3A + 1) c3 = (n - 2 ) c2rf*+2, * ( 8 .2 )
To avoid a singularity, any zero of dZjdQ. in the domain of the fluid must also be
a zero of dW/dQ. This gives us in general a condition to determine A. Then the
co-ordinates (x, y) and the velocity components (u, v) are given parametrically by
x + iy = tr*xZ> u —iv = AHnA+1)W. (8.7)
In the special case n = 2 we have
Z - n , n f . Q - (A + .H 2A + i r <8-8’
л 4 1 а + (Ш )Щ Т Т )- “ 9)
Now dZjdO. vanishes at Q = —г/(A + l), but since dW/dCl has no zeros, the
branch-point cannot be annulled. Nevertheless the solution given by (8 .8 ) and
(8.9) can, for general A, still represent the flow outside a parabolic free surface,
whose scale, from (8.4) varies like ^_2Л.
1753
616 M. S. Longuet-Higgins
У/73
Figure 6. The form of the free surface in the self-similar flow described by (8.14). The
points A and В are branch-points; A is annulled by choice of Л. The curve expands about
the origin with dimensions proportional to
1
A+ l (A + l) (2A+ 1)
(8. 11)
A+l (A + l) (2A+ 1)
7A + 4 = 0 ,
This is shown in figure 6 . At large distances the tangent becomes nearly parallel
to the X axis. From (8 .11) the branch-points are given by
П = - 7 i, (8.16)
15
li 16.73 .
and Z= (8.17)
3’ 15.9
The branch-point (8.16), which is on the —Y axis, is already annulled, the other
is not. So the solution is valid for a fluid filling the space on the convex side of the
free surface. The motion is like an expanding trough. From (8.7), the dimensions
of the trough vary like ^-лЛ, that is like By reversing t we have a description
of a trough ‘ filling up*, though it must be remembered that the solution is
essentially gravity-free.
where now
z = a n+id1a » - 1 - a»-* - w 3 a » -3+ ...,
(9.6)
W = £2" + (A - l)idxfl ” - 1 + (2 A- 1 ) d2П»-3- ....J
1755
618 M. S. Longuet-Higgins
We now proceed as before. In the case n = 2,
(9.7)
W = П2+ 2»П+т;-!--.
Л— 1
The only zero of dZ/dQ, is when П = —г/(Л—1), and on inserting this value in
dWjdO. = 0 we get A—1 = 1 so A = 2 . This gives the solution of §7.
In the next case, n = 3, we have
(9.10)
(9.11)
f The complex roots of (9.12) do not represent solutions, since A would have to be
replaced by A* in (9.10).
1756
F ig u r e 7. The form of the free surface in the self-similar flow described by (9.13). The
branch-point В is excluded from the flow. The free surface contracts like J-4"17 and on the
right becomes parallel to the X axis.
wall. As t increases to infinity the point of contact of the free surface moves out
to the end of the wall, and the free surface curls tightly round it, becoming
eventually almost plane and perpendicular to it.
Solutions for higher values of n may be investigated similarly.
It will be noticed that because the function r(o), t) is the same in (8 .2 ) and (9.2),
and (7.2) is linear in £, any linear combination of (8.1) and (9.1) will satisfy (7.2)
also, and indeed we may add expressions corresponding to different values of n.
However, (8.4) and (9.4) show that the relation between £ and Z is not the same
in the two cases, or for different values of n, so that the corresponding expressions
will not represent self-similar flows.
An exception might occur for two values of n, say n± and n2, for which
%A = w2A—1, X = ll(n2—n1).
But this possibility is excluded by the restrictions on A which are implicit in
(8.3) and (9.3).
1757
620 M. S. Longuet-Higgins
Hence the two classes of flows given by (8 .1) and (9.1) are possibly the only
self-similar flows of polynomial type that can be readily realized. (We can of
course replace o) by any rational function of a) alone, without essentially altering
the flow.)
REFERENCES
D irichlet , G. L. 1860 Untersuchungen iiber ein Problem de Hydrodynamik. Abh. Kon.
Ges. Wiss. Gottingen, 8, 3-42.
G ababedian , P. 1953 Oblique water entry of a wedge. Comm. Pure Appl. Math. 6 ,
167-167.
G ilbabg , D. 1952 Unsteady flow with free boundaries. Z. angew. Math. Phys. 3, 34-42.
G ilbabg , D. 1960 Jets and cavities. Handbuch der Physik, vol. rx (ed. S. Flugge), pp. 311-
445. Springer. (See especially pp. 15-18, 350-358.)
J ohn , F. 1952 An example of a transient three-dimensional flow with a free boundary.
Rev. Gen. Hydratd. 18, 230-232.
J ohn , F. 1953 Two-dimensional potential flows with a free boundary. Comm. Pure Appl.
Math. 6 , 497-503.
KAbm JLn , T. von 1949 Accelerated flow of an incompressible fluid with wake formation.
Ann. Mat. Pure Appl. 29 (4), 247-249.
L amb , H . 1932 Hydrodynamics, 6th edn. Cambridge University Press.
L onguet -H iggins , M. S. 1972 A class of exact, time-dependent, free-surface flows. J . Fluid
Mech. 55, 529-543.
L onguet -H iggins , M. S. & Cokelet , E. D . 1976 The deformation of steep surface
waves: I. A numerical method of computation. Proc. Roy. Soc. A (in press).
T aylo b , G. I. 1960 Formation of thin flat sheets of water. Proc. Roy. Soc. A 259, 1-17.
Y ih , C.-S. 1960 Finite, two-dimensional cavities. Proc. Roy. Soc. A 258, 90-100.
1758
Plunging breakers are beyond the reach of all known analytical approxima
tions. Previous numerical computations have succeeded only in integrating
the equations of motion up to the instant when the surface becomes
vertical. In this paper we present a new method for following the time-
history of space-periodic irrotational surface waves. The only independent
variables are the coordinates and velocity potential of marked particles at
the free surface. At each time-step an integral equation is solved for the new
normal component of velocity. The method is faster and more accurate
than previous methods based on a two dimensional grid. It has also the
advantage that the marked particles become concentrated near regions of
sharp curvature. Viscosity and surface tension are both neglected.
The method is tested on a free, steady wave of finite amplitude, and is
found to give excellent agreement with independent calculations based on
Stokes’s series. It is then applied to unsteady waves, produced by initially
applying an asymmetric distribution of pressure to a symmetric, progres
sive wave. The freely running wave then steepens and overturns. It is
demonstrated that the surface remains rounded till well after the over
turning takes place.
1. I n tr o d u c tio n
Breaking waves are the agent for many significant processes in the upper ocean
including the transfer of horizontal momentum from wind-waves to surface cur
rents. Yet remarkably, one of the'most familiar and spectacular properties of the
sea surface-its capacity to turn over on itself-is one of the least well understood.
All the usual theories for surface waves - the small-ampHtude approximations of
Airy and Stokes, the nonlinear shallow-water theory and the Korteweg-De Vries
equations for solitary and cnoidal waves - are essentially approximations, valid only
when the fluid acceleration is sufficiently small compared to gravity. These approxi
mate theories cease to be valid when the acceleration is comparable to g, or when
the surface elevation is a multivalued function of the horizontal displacement.
The small-amplitude theories can indeed be carried to higher approximations,
showing that the form of steady waves of large amplitude tends, in the limit, to the
simple comer-flow discovered by Stokes (1880 a) in which the free surface has a sharp
wave develops in time, ultimately turning over and plunging towards the forward
face of the wave.
Some of the implications of this straightforward calculation are discussed in § 11.
This paper is intended to be the first of a series in which the present method of
computation is employed as one tool in a systematic investigation of free surface
flows and breaking waves.
2. B asic eq u a tio n s
Let (x,y) denote rectangular coordinates with the ж-axis horizontal and the
у-axis vertical, as in figure 1. The motion is assumed to be periodic in the ж-direction,
with period
L = 2п/к. (2 .1)
The fluid is inviscid and incompressible, and the origin is taken to be in the mean
surface level. The motion may be assumed to be started from rest by conservative
forces, so that it is irrotational at all times; any slow diffusion of vorticity inwards
L = 2k. (2.4)
We have now the following equations for the velocity potential ф. Since the fluid
is irrotational and incompressible
where g = ! + v^.V, (2 .8 )
which denotes differentiation following a given particle. We have also the dynamical
condition derived from Bernoulli’s equation, namely that
Ъф1Ы= - р 3- у - \ ( Ч ф ) \ ( 2 .9 )
where padenotes the pressure applied at the surface. From this we can immediately
derive the rate of change of ф following the motion, namely
The difference between the right-hand sides of (2.9) and (2.10) is simply a change of
sign in the last term.
3. T ra n sfo rm a tio n o f co o rd in a te s
where f is a new complex variable, analytic and single-valued everywhere inside the
contour С which corresponds to the fluid surface (see figure 2 ). (ir,6) are polar
coordinates in the f-plane, and we have from (3.1)
г = е*, у = In r,
(3 .2 )
в = —x, X —— 6 .
All points at infinite depth in the (x>y) plane are transformed into the origin О in
the £-plane.
1763
if- plane
(3.6)
4. T he D ir ic h l e t problem
Suppose that at some initial instant t = t0 we are given the velocity potential ф
throughout the fluid, and hence the value of ф and its derivatives both inside and
on the contour C(t0). Let (г, в) denote the (Lagrangian) coordinates of a particle on
1764
C(t0). Then equations (3.5) will determine the position of the same particle a short
time dt later. Similarly (3.6) will determine the value of + df) on the new contour
C(£0 + df). By considering adjacent particles, and differentiating along the surface
we can then obtain the tangential component of velocity дф/дв. But this does not
immediately determine the normal component дф/дп, which is also needed for the
step afterwards.
The problem of determining the normal component of velocity at the boundary
is equivalent to the Dirichlet problem of finding the normal gradient of a function ф
whose values are given on a closed contour C, and which is harmonic (V20 = 0)
everywhere inside G. We may formulate the problem as an integral equation as
follows.
Let (s,n) be tangential and normal coordinates at a typical point P on the
boundary (see figure 3), and let (E, oc) be the polar coordinates of P with respect to
an arbitrary point Q{r0, в0) in the interior. Let
S= ^nB (4.1)
( i '2 )
J c S b -Bd4 - P j V a - ^ 0. (4.5)
Since the right-hand side involves only the values of ф on G, which are known, and
since JR, a are determined solely by the shape of C, equation (4.5) is an integral
equation for (дф/дп), with a singular kernel In J?.
It remains to express the time-derivatives of г, в and фin terms of the tangential
and normal derivatives of ф. We have in general
0 J . J
0; = c o s ^ +sm % ’
13 • *э в9
7 S0 = sm% - CO8% ’
dr ■ * = т—.
se
where sm/?
с03^ = э ? os
Dr 0дгдф двдф
d t = r a«a7+r aia^’
m ,г е м дгдб
(4.6)
d t. ~ Г dsds ГЪзЪп*
)'}■)
Equations (4 .5 ) and (4 .6 ), together with initial conditions, are the basis for the
following computations.
5. S o lu tio n of the in t e g r a l e q u a t io n
The values of ф, Ъф/ds and дф/дп are to be evaluated at a finite number N of points
on the boundary which will correspond to fixed particles. We label these with the
suffix j , where j = 1, 2 ,..., JV. The corresponding values of г, в are similarly labelled
rp 6j. We will suppose that Q lies in turn at (r{, where г = 1, 2 , ...,N also. Hence
It, a, s take the values JRij} ccip s{j. The integral on the left of (4.5) can be approxi
mated as the product of the {N x 1) matrix (дф/дп)] multiplied by an (N x N) matrix
1766
of coefficients, say Ay. The integral on the right can be approximated by a quantity
Hence we have N linear simultaneous equations of the form
Ау(Ъф1Щ = Bi (i = 1, ...,N) (5.1)
to be solved for the values (дф/дп)^
The success of the method will depend critically upon the accuracy with which
the integrals in equation (4.5) are approximated by linear sums. Let г be fixed, and
let us take Q(rit 6^ as origin of s, writing si}- —s(j_i)mo<1N. Since In В is singular both
at s = 0 and 5 = sN, but In (i2/s) or In [BI(sN—5)] is not, we write the integral on the
left of (4.5) as follows
Each term in the integrand can be evaluated exactly. To find the derivatives of
дф/дп with respect to s, we approximated (S^/dTi),- by a fourth-order Lagrangian
interpolation polynomial over the 5 points centred on j, and differentiated this.
The integral I6is handled in a similar way, except centred on the point (j -f 1). This
technique gives local errors of order (As)6In (As).
To calculate the arclengths sp the quantities and](fy —Znj/N) as functions of,?
(considered as a real variable) were approximated by cubic splines (see Ahlberg,
Nilson & Walsh 1967). From these can be calculated dr/dj, dd/dj and hence
g - [ ( D 4 $ 7 .
at each point j. Equation (5.4) was then integrated by'Simpson’s rule, giving s} at
1767
each point, with local errors of order (As)4. The tangential derivative dф|dj was
calculated similarly. Hence we found
(5.5)
© ,-© y ® /
Because the curve is a simple closed contour we have aN = с^+я. The right hand
side of (4.5) can now be written
6. T i m e -s t e p p i n g
dy/dt = f(t) (6 .1 )
1768
is as follows:
У\р = Уо+Y i (55/o - 59/_1+ 37/ _ 2 - 9/_3),l
д , (6 -2 )
2/ic = 2/о + 2^ ( 9 /ip + 1 9 / 0 — 5 / _ x + / _ 2) . J
Here A£ denotes the short time interval, f n denotes f(t+nAt) and ylp, ylc denote the
‘ predicted* and ‘ corrected * values of уг (see Acton 1970). The method is fourth-
order, local errors being 0(At)b, but requires only two evaluations of / at each time
step. Applied to equations (4.6) this means that we have to solve the integral
equation for дф/дп only twice for each time-step.
Since the A.B.M. method needs information from three previous time-steps, a
fourth order Runge-Kutta (R.K.) technique was used to make the first three time-
steps from the initial conditions (see Gerald 1970). This method uses no information
about previous time-steps, but takes four mini-steps forward from the current time.
A weighted average is then used to calculate the function at the new time. The U.K.
method requires four evaluations of the time-derivative at each step and is thus
twice as time-consuming as A.B.M.
7. C h ecks of accuracy
The accuracy of our numerical solution to the integral equation (4.6) was tested
by calculating, at each time-step, the value of
This represents the total ‘ outflow5across С in the £-plane, or across G' in the z-plane,
and it is clear from Green’s theorem, or from considerations of continuity, that
Q should vanish identically. The numerical value of Q was found by writing
Ч Ж *
and integrating by Simpson’s rule.
As further checks we calculated in a similar way the mean level
1 f 2-rr 1 /*2тг
= 2J 0 y^ = - T n ) a Ыгйв (73)
which also should vanish, and the potential and kinetic energies
“ d T ' i r f - . - s j / s 8* <’ •«
1769
8. I n s t a b il it y a n d sm o o th in g
In nearly all computations, the wave profile, after a sufficiently long time,
developed a saw-toothed appearance, in which the computed positions of the
particles lay alternately above and below a smooth curve (see, for example, figure 4).
The cause of the instability is unknown. Tests showed that, once started, the rate of
growth of the instability, per unit time, was independent of the number of time-
steps. Hence it is not due simply to rounding errors. The growth may be partly
physical, being similar to the growth of short gravity-waves by horizontal compres
sion of the crests of longer waves (see Longuet-Higgins & Stewart i 960 ; Phillips &
Banner 1974). In reality these instabilities are partly damped by viscosity, which
we have neglected.
The instability was effectively removed by the following procedure. A function
f(x) defined at equally spaced points xi (j = 1, 2 ,3,...), and in which alternate points
lie on a smooth curve, can be locally approximated by two polynomials, say
7j = A ( +Щ + - Л+2) <8-3)
and when n = 3 we find in a similar way
Figube 4. Comparison of the profile of a steady progressive wave in deep water (8 = 0.80)
computed by Pad6 approximants from Stokes’s series (smooth curve) and the corre
sponding time-stepped profile (unsmoothed) represented by the oircles. N = 30. The
profiles are compared at times (a) t = 0, (b) t = (c) t = $тг, (d) t — n. Note the growth
of the instability at t = n .
1771
Both formulae were tried for smoothing the functions rp 6j and фр at every 5 , 10 or
20 time-steps. Both worked well and gave a smooth profile with no appearance of
the small-scale oscillations. For certain reasons, as stated in the following section,
the formula (8.3) was generally preferred.
An important and essential test of the method was as follows. For initial conditions
we took a symmetric, progressive wave of finite amplitude in deep water (see
figure 4). A very accurate method of computing the wave profile, based on Stokes’s
expansion (1880 b) has recently been developed by Schwartz ( 1972, 1974). This
nees Pad6 approximants to sum otherwise divergent series. In fact we adopted
a modification of Schwartz’s method due to Cokelet ( 1975) in which the expansion
parameter is taken as
usb/c4, (9.1)
where gcre8t and trough denote the particle velocities at the wave crest and the
wave trough, in a frame of reference moving with the phase-speed c.
We selected a wave corresponding to 8 = 0.80, whose crest-to-trough height is
about 90 % of that of the highest wave in deep water. As initial values for the
numerical integration, we inserted the computer coordinates of the wave profile
(see figure 4) and the corresponding values of the velocity potential ф, in the frame
of reference for which the deep water was at rest, and therefore the motion generally is
time-dependent.
In figure 4 the initial wave has its crest at x = n, and the wave progresses from
left to right. At time intervals of two wave profiles are plotted: the continuous
curve represents the Pad6-approximated profile, and the plotted points represent
the positions of surface particles found from numerical time-stepping by the present
method (with N = 30).
The plotted points in figure 4 are without smoothing, and the growth of the
saw-toothed instability from the third to the fourth profile is apparent. Actually,
the unstable wave is found to conserve total energy to one part in 100 up to t = n,
when the computations break down.
Figure 5 now shows the computed points, compared with the steady-wave profile,
at time t = 2k, when smoothing has been applied after every 6 time-steps. As can
be seen, the profiles are indistinguishable except to the sharpest eye. For this
particular plot the 7-point smoothing formula (8.4) was used. The flux Q of equation
(7.1) varied between б x 10" 6 and 4x 10"3, hovering around lO" 4 for most of the
calculation and showing no tendency to increase. The mean surface displacement y9
varied from 3 x 10-5 to 1 x 10-3. The total energy E lay between 0.06938 and
0.07007 (the value from Pad6 -approximants is 0.06996). The wave energy decreased
by 0.14 % at each smoothing with the 7-point formula, compared with 0.06 % using
the б-point formula. Although asymptotically less accurate, the б-point formula
1772
seemed to perform better in practice. When the number N of integration points was
raised to 60, only 0.003 % of the energy was then lost at each smoothing.
The numerically calculated profile also remained very closely in-phase with the
Pad£-approximated profile, using the calculated phase speed of the steady wave.
These checks suggest strongly that the numerical method remains accurate over
the time-scales in which we are interested, which are of the order of one wave period.
F igure 5. Comparison of the profile of a steady wave calculated from Stokes’s series (smooth
curve) and the corresponding time-stepped profile, with smoothing, represented by
circles, at time t = 2n. (This is twice the maximum duration in figure 4.) N = 30; 7-point
smoothing every 6 time-steps. The two profiles are indistinguishable.
10 . P l u n g i n g breakers
symmetric progressive wave of § 9, for which $ = 0.80. The energy E0 of this wave
is 94 % of the maximum possible energy Emof a progressive wave of the same
length in deep water (which corresponds to £ = 0.92; see Longuet-Higgins 1975 a;
Cokelet 1975). At time t — 0 , a wave crest is at x = n, and the wave progresses from
left to right (see figure 6 ). The pressure applied at the surface is taken to be
[j30sin$sin(a; —ct) (0 < t
(10.1)
У
1
0 x
-1
1
(b)
F igu re 6. A time-sequence of profiles for the pressure amplitude p0 = 0.146 at times (a) t = 0,
(b) t = *7t, (o) t = fit, (d) t - л, (e) * = ( /) « = К (?) f = fit. = 60; 6 -point
smoothing every 6 time-steps. The surface pressurep 0 is applied tilH = я; then the wave
runs free.
increased to 60, since 30 points proved too few. The 5-point smoothing formula was
used every 5 time-steps.
Figure 6 shows a sequence of wave profiles at various times for p0 = 0.146. The
dimensionless time interval between plots is except for the last three plots,
which are separated by a time interval of %n. The wave begins at (a) (t — 0 ) as a
2 Vol. 350 . A.
1775
steady wave and slowly increases in amplitude as it moves to the right. The pressure
forcing is removed at (d) and the wave is thenceforth free. The front face quickly
steepens, and at (g) the crest is overhanging.
Figure 7 is a comparison of the profiles of the waves corresponding to p0 = 0.0729,
0 .100, 0.126, 0.146 at the same time t = 5.066 after applying the pressure. It is clear
F i g u r e 7 . A c o m p a r is o n o f w a v e p r o file s a t t h e s a m e t im e t = 6 .0 6 6 a f t e r f i r s t a p p l y i n g t h e
s u r f a c e p r e s s u r e , (a ) pQ = 0 .0 7 2 9 , (b)pQ = 0 .1 0 0 , (o ) p0 = 0 .1 2 6 , ( d)p9 = 0 .1 4 6 .
1776
that the larger the pressure forcing, the farther towards breaking are the waves.
Notice that the profiles remain smooth and free from instability.
Figure 8 shows a close-up of the overturning crests, in the case p0 = 0.0729. This
is a series of plots of a small region near the crest of the wave, as seen by an observer
moving horizontally with the speed of an infinitesimal wave of the same wavelength
(i.e. (ж—i) is plotted). The positions of the particles have been marked with small
circles, and each profile drawn by connecting adjacent particles with straight lines.
(x - t )
F ig u h e 8. Close-up of the wave-crest at successive times, with a plotting interval of л/160.
p0 = 0.0729; 6.596 < t < 6.890.
On the other hand the imaginary lines joining consecutive positions of each small
particle give an idea of the particle trajectories, in this frame of reference.
Similar profiles for the cases p0 = 0 .100, 0.126 and 0.146 are shown in figures 9-11.
Perhaps the most remarkable feature of figures 8-11 is that the crest does not
tend to develop a sharp angle, as for the highest steady wave, before overturning
2-2
1777
takes place. Instead, a smooth jet of fluid is ejected from the forward face of the
wave.
Our method of computation appears valid till well after the surface is vertical.
However, the curvature at the forward tip of the wave appears to increase in time,
so that the computations cannot be continued beyond a certain point.
Figure 12 shows the complete wave profile corresponding to the last wave crest
plotted in figures 8-11. We notice a tendency for the more highly forced waves
(12 c) and (12 d), which are also more energetic, to start breaking at lower values of
the wave height.
Finally in figures 13-16 we show the time-variation of the kinetic, potential and
total energies, respectively, for each value of the forcing pressure p0. The oscillations
between kinetic and potential energy suggest that there are standing-wave com
ponents in the motion. The total energy, however, increases smoothly to a constant
value. Also indicated in figure 16 are the values of the final energy E divided by the
(x -t)
energy, Emi of the most energetic progressive wave, that is Em—0.07403, for
$ = 0.92.
We mention the results of the checks listed in § 7. The flux Q remained at about
10~5 to 10-4 for most of the calculation, increasing to 10-3 near the end. The mean
level y8 was of order 10-4 for all the calculations, except near the last profile for
p0 = 0.146, when it reached 1 x 10-3. The total energy E remained very constant
after t = n for all runs, its largest fluctuation being a decrease of about 0.2 %.
The above calculations were programmed in Fortran rv double precision, and
carried out on the I.B.M. 370/165 at Cambridge. The computer time needed for each
pressure forcing to achieve overhanging waves was typically 35min for 270 time-
steps at 169 К bytes of storage.
(я-*)
F iq u b e 10. Ав figure 8, but with p0 = 0.126. 4.830 ^ < 5.046.
1779
(x -t)
Figubb 11. As figure 8, but withp0 = 0.146. 4.712 < t < 4.928.
1780
instant when the tangent first becomes vertical. From the results of § 10 we conclude
that waves do not necessarily develop a sharp comer or singularity before the free
surface overturns. Instead they can curl over and plunge towards the forward face
o f the wave, and there is nothing to suggest that the flow does not remain smooth
(neglecting surface tension) up to the instant of impact. The Stokes 120° angle may
у
not be a typical fluid flow but simply a highly singular form, intermediate between
a steady, symmetric wave and an unsteady, unsymmetric, plunging wave. In the
gravity-free flows investigated by Longuet-Higgins ( 1972, 19756 ) a mass of fluid is
deformed into a continually elongated jet with a thin tip. A plunging breaker which
is nearly free-falling may behave similarly. It is also possible that most spilling
F igure 13. The kinetic energy Г as a function of the time t, for each pre
(a) p 0 = 0.0729, (6) p0 = 0.100, (c) p0 = 0.126, (d) p0 = 0.14
F igure 14. The potential energy V as a function of the time t, for each pressure amplitude.
(a) p0 = 0.0729, (6) p0 = 0.100, (c) p0 = 0.126, (d) p0 = 0.146.
1782
t
F igure 15. The total mean energy [E = T + V) as a function of the time t, for each pressure
amplitude, (a) p0 = 0.0729, EJEm = 1.37; (b) p0 = 0.100, E fE m = 1.56; (c) p0 — 0.126,
E jE m = 1.73; (<Z) p0 = 0.146, E {E m = 1.88
Our computations, which are for deep water, show incidentally that for the
breaking o f irrotational waves a sloping bottom is not necessary. Indeed, b y elimin
ating the bottom, one of the parameters of the problem, namely the ratio o f the
depth to the wavelength, has been eliminated. On the other hand, similar computa
tions can easily be done for periodic waves in water of any arbitrary uniform depth.
B y varying the magnitude and duration of the applied surface pressure, it should
in future be possible to gain insight into the energy and momentum lost by sea waves
b y wave breaking, both in deep and shallow water, as a function of the energy input,
and hence to gauge the transfer of momentum from the waves to surface currents.
For simplicity we have assumed that the waves are initially progressive. Here
again this assumption can be generalized and we can apply the method either to
standing waves, or to waves that are a mixture of opposite but unequal progressive
waves. These and other possible applications will be studied in future papers.
two years of his research, E. D. C. was supported b y a Graduate Fellowship from the
U.S. National Science Foundation, and for a third year b y a Research Studentship
from the Cambridge Philosophical Society. During a final year, research fees were
remitted b y Cambridge University. To all these bodies past and present we express
our thanks and appreciation.
R eferences
Acton, F. S. 1970 Numerical methods that work. New York: Harper & Row.
Ahlberg, J. H ., Nilson, E . N. & Walsh, J. L. 1967 The theory of splines and their application.
New York: Academic Press.
Biesel, F. 1952 Study of wave propagation in water of gradually varying depth. In Gravity
waves. Giro. U.S. natn. Bur. Stand, no. 621, pp. 243-253.
Buckingham, R . A . 1962 Numerical methods. London: Pitman.
Chan, R. K . & Street, R . L. 1970 SUMMAC - A numerical model for water waves. Tech. Rep.
Dep. Civil Engrg, Stanford University, Calif., no. 135.
Cokelet, E. D . 1975 Steady and unsteady nonlinear water waves. Ph.D. dissertation,
Cambridge University.
Gerald, C. F. 1970 Applied numerical analysis. Reading, Mass.: Addison-Wesley.
Longuet-Higgins, M. S. 1953 a On the decrease of velocity with depth in an irrotational
water wave. Proc. Camb. Phil. Soc. 49, 552-560.
Longuet-Higgins, M. S. 1953 b Mass transport in water waves. Phil. Trans. R. Soc. Lond.
A 245, 636-681.
Longuet-Higgins, M. S. i 960 Mass-transport in the boundary layer at a free oscillating
surface. J. Fluid Mech. 8, 293-306.
Longuet-Higgins, M. S. 1972 A class of exact, time-dependent, free-surface flows. J. Fluid
Mech. 55, 629-543.
Longuet-Higgins, M. S. 19750 Integral properties of periodic gravity waves of finite ampli
tude. Proc. R. Soc. Lond. A 342, 157-174.
Longuet-Higgins, M. S. 1975 b Self-similar, time-dependent flows with a free surface. J. Fluid
Mech. 73, 603-620.
Longuet-Higgins, M. S. & Stewart, R . W . i 960 Changes in the form of short gravity waves
on long waves and tidal currents. J. Fluid Mech. 8, 565-583.
Phillips, О. M. & Banner, M. L. 1974 Wave breaking in the presence of wind drift and swell.
J. Fluid Mech. 66, 625-640.
Price, R. K . 1971 The breaking of water waves. J. geophys. Res. 76, 1576-1681.
Sohwartz, L. W . 1972 Analytic continuation of Stokes’ expansion for gravity waves. Ph.D.
dissertation, Stanford University.
Schwartz, L. W . 1974 Computer extension and analytic continuation of Stokes’ expansion
for gravity waves. J. Fluid Mech. 62, 653-578.
Stokes, G. G. 1880 a Considerations relative to the greatest height of oscillatory waves which
can be propagated without change of form. Mathematical and physical papers, vol. 1,
pp. 225-228. Cambridge University Press.
Stokes, G. G. 1880b Supplement to a paper on the theory of oscillatory waves. Mathematical
and physical papers, vol. 1, pp. 314-326. Cambridge University Press.
1784
I. Superharmonics
B y M. S. L o n g u e t - H i g g i n s , F.R.S.
Department of Applied Mathematics and Theoretical Physics, University of
Cambridge, Silver Street, Cambridge CBS 9EW, U.K. and
Institute of Oceanographic Sciences, Wormley, Surrey, C/.isT.
i. I n t r o d u c t i o n
472 M. S. Longuet-Higgins
bring to bear analytical methods and approximations. Leaving aside the later
stages of plunging breakers, our aim in the present paper is to study analytically
the initial development of the motion, when the flow may be treated as a small,
time-dependent perturbation of a steep, steady wave (as viewed b y an observer
travelling along with the phase-speed). The form of the steady wave, though
evidently nonlinear, can be well calculated, and it is permissible to treat the time-
dependent part of the motion as a linear perturbation of this steady flow.
There is a second reason for undertaking such an investigation. Though the
mathematical existence of gravity waves on deep water has not been in doubt
since the work of Levi-Civita (1925), nevertheless the question of their stability
to small perturbations has remained unsettled. Benjamin & Feir (1967), noticing
the tendency towards disintegration of waves in a laboratory channel, successfully
analysed one particular type of instability, due to the growth of sidebands. This
causes a slow modulation of the wave envelope, with a space-scale much larger
than one wavelength. The question whether gravity waves are unstable on short
length-scales, comparable to, or less than, one wavelength, is not answered by their
analysis.
A t the same time, recent computations of steep, perfectly periodic waves
(Longuet-Higgins 1975; Longuet-Higgins & Fox 1978) have shown the existence
of stationary values in the phase-speed, considered as a function of the wave height
(see figure 1 of the present paper). It is not unreasonable to expect that a maximum
in the phase-speed will mark the onset of a new type of normal-mode perturbation,
probably unstable. Such an instability would be of the same scale as a wavelength,
or less. The observed speed with which overturning of the free surface can take
place suggests indeed that this type of breaking may begin as a fairly localized
instability.
In this paper our aim is to investigate all types of instability, both those on a
space-scale that is less than one wavelength (superharmonics) and those with scale
greater than a wavelength (subharmonics). For reasons of exposition we take first
the superharmonics.
In our problem the basic motion is a wave of finite amplitude, which cannot be
treated as a small perturbation. Hence we develop in §2 a type of analysis in
which as independent coordinates we take the stream-function and velocity-
potential of the flow at any instant, together with the time t. This is an extension
of Stokes’s (1880) method for steady flows. The boundary conditions are then
expressed in the new variables (equations (2.8) and (2 .9 )). To analyse the perturba
tions, we m ay take as basic coordinates the values of ф and ijf in the unperturbed
flow (that is to say, the symmetric wave of finite amplitude) and linearize with
respect to the small perturbations (§ 3 ). The analysis of the perturbations into
normal modes, and the method of calculating the eigenfrequencies, is described
in § 5 and the results are discussed in §§ 6-10.
The calculated eigenfrequencies are shown in figure 1 as functions of the steepness
ak of the unperturbed wave. I t appears that all the superharmonic perturbations
1786
remain stable up to ah = 0.42 at least, but there is indeed evidence for the occur
rence of an instability at about ak = 0.436, the amplitude corresponding to the
maximum phase velocity. For reasons given in §9, other stationary values of the
phase-velocity probably correspond also to the onset of instabilities.
The relation of these unstable modes to the subharmonic instabilities of Ben
jamin & Feir will be discussed in part II.
2. T r a n s f o r m a t io n of b o u n d a r y c o n d it io n s
474 M. S. Longuet-Higgins
To transform the kinematic condition (2 . 1) we note that since
d , 9 , 0 0
d< Yx dx bt
we have dф
(2.6)
from equations (2.3). Now let the free surface be given in the form:
+ *шЖ(ф%Ь). (2 .6 )
Then the kinematic condition can be expressed as
d(F —г/г)/dt = 0 ,
(i I Ь{-Х'У">\ p , д(*’ У) p I n f“ 7)
1 1 + т г ) ) F*+ m r ) F*+ w j ) - °- (2-7)
Finally in equations (2.4) and (2.7) we may use the Cauchy-Riemann relations
for ХфУур Xp, y^. to obtain two boundary conditions for у{ф, t) and F $ , t), namely
3. P e rtu rb a tio n a n a ly s is
(2.9) we obtain two expressions which must vanish on the free surface \Jr = F. But
on expanding each expression in a Taylor series in \jr we obtain expressions which
are to vanish instead on xjr — 0. Thus to order e we have respectively
—(Хф^t + ^fr£t)
+ p (Y +Y^F + ?7)[(7$ + Y\) + (7^ + Y^^F + 2 ( 1^ 77^+ 1^ 77^)] + J = 0, (3.2)
- ( Y t Vt+Y+£t) + g ( Y } + Y l ) ( Y t F + V)
+g[Y(Y$+ Yl)+F+2Y(YtVt+ Y f ^ )] = 0 .
2P 7 = - 1 (3.7)
where P = + Y\,
Q = 2 YYr
(3 .9 )
R = 2YY+,
S = [ 7 ( r j+ r » ) V , )
476 M. S. Longuet-Higgins
4. P e r t u r b a t io n of a u n if o r m flow
6. C a l c u l a t i o n o f a steady, s y m m e tric w a v e
where с represents the stream velocity at infinite depth (equal to minus the phase
velocity in a stationary frame of reference) and # 0, Hlt... are real constants. The
wavelength is taken to be 2n and this together with the choice <7 = 1 fixes the
scales of length and velocity. The crest of the wave is at (ф, iff) = (0 , 0 ). The co
efficients # 0, Hlt ... and с must be chosen so that the condition (3.4) is satisfied.
This condition leads, as is well known, to a one-parameter system of waves in
1790
к = 2rc/wavelength = 1 (6 .3)
in our chosen units. The quantity ak is also sometimes called ‘ the dimensionless
w ave am plitude’ .
Th e Gn were calculated in quadruple precision to order a120, b y the simple algo
rithm described in Longuet-Higgins (1978). The results for the phase velocity с
are shown in fig. 1 of that paper, where c2 is plotted against ak over a range including
the higher waves. The calculations check very precisely the values of c2 from an
earlier calculation (Longuet-Higgins 1975). In particular the maximum value of
the phase velocity as a function of wave steepness are verified. This occurs at
ak = 0 .436 ..., as compared to the maximum wave steepness, which is when
ak = 0 .4434 . A quite independent calculation of the phase-speed b y an asymptotic
analysis valid in the limit as a k ^{a k )mftx has been given by Longuet-Higgins &
F o x (1977, 1978), with practically identical results.
6. P e r t u r b a t io n s of a s t e a d y w a v e
W e shall assume at first that the perturbations £, 17, F are also periodic in ф with
the same basic wavelength 2 n. This temporarily excludes all subharmonics of the
wavelength and also instabilities of the Benjam in-Feir type, though these will be
considered later. In general we have
478 M. S. Longuet-Higgins
The origins of ф and \Jr, in the perturbed motion, are in general arbitrary. How
ever, we have fixed the origin of ф by assuming P(t) to vanish in equation (2 .2 ).
The origin of \fr can now be fixed by assuming that in the expansion of the free
surface \jr = F we have
c0 = 0. (6.3)
Now the functions 5^, Y^, P, Q, R, S can all be expanded in Fourier series in ф
(see appendix A). Of these, Y^, P, Q, S are even functions of ф, involving terms in
cos(ti0/c), and Y^ R are odd functions, involving %тпф. Now substituting the
series (6.1) and (6.2) into equations (3.8) and equating coefficients of соапф
(n = 0 , 1, 2 ,...) and siппф (n = 1, 2,...) we obtain an infinite system of equations
for the unknown coefficients an, bn, cn, dn in the form
ХАх-Вх (Л = -i(r),
where x denotes the column-vector whose elements are (an, bn, cn, dn) and A, В
are matrices of coefficients independent of <r. This system of equations may be
solved by truncating the vector x at successively higher values of n, neglecting
terms greater than n = N, say. Thus at the iVth stage there are (4N + 2 ) unknowns
a0, ..., fljyj ..., bjsf; Cj, •••> j с?!)..., dpj and (T. From the coefficients of cos тьф
(n = 0,1, ...,N) and s i n (n = 1, in the two equations (3.8) we get (42У+ 2)
homogeneous equations, whose coefficients form a square array. The vanishing of
the corresponding determinant gives a characteristic equation for Л, namely
|(АЛ„-В„)| = 0 ,
where N denotes the order of the truncation. The roots of this equation, if they
converge as i^-^oo, give the instabilities of the system. Greatest interest attaches
to those roots, if any, for which i?(A) is large and positive.
Details of the matrices AN and BN, which are important for actual computation,
are given in appendix A. The matrices were constructed in F o r t r a n rv double
precision, for the IBM 370/165 at Cambridge University. The eigenvalue equation
7. T he computed e ig e n v a l u e s
In figure 1 are shown the radian frequencies <r of the computed normal-mode
perturbations, as a function of the steepness ak of the unperturbed wave. These
have been designated n — 1,2, 3,... and n = —1, —2, —3,... on the following basis.
1792
in the frame of reference moving with the speed of the unperturbed wave. This can
be seen from §4 by writing k = \n\ = 2 . Generally, those perturbations of wave
number n ^ 1, travelling in the same sense as the unperturbed wave, have a radian
frequency cr = » - rai (7.2)
in the limit as ak 0.
1793
480 M. S. Longuet-Higgins
Negative values of n correspond to perturbations travelling in the opposite
sense to the unperturbed wave. These have a relative frequency
(T=\n\ + \n\i (7 .3)
in the limit as ak 0 . Thus the perturbation n = —1, having the same wavelength
as the unperturbed wave but travelling in the opposite direction, has a frequency
cr = 2 ; to an observer moving with the speed of the unperturbed wave the frequency
of encounter of crests of the perturbation is just doubled.
(ak)*
0 a05 _ 0.10 _ . 0.15
ak
Figure 2. Aar plotted against (ak)*, showing the decrease in frequency of the normal modes
due to the finite amplitude of the basic wave. Broken lines represent the asymptotes (7.8).
cr = (7-7)
Writing = |n| in this formula we see that the frequency of the first wave is
shifted by an amount
Acr = —\(ак)2\п\, (7.8)
where ak is the steepness of the unperturbed wave and n the number of the harmonic.
Thus regardless of whether n is positive or negative the relative frequency always
tends to be reduced. (The case n = 1 is excluded since then the two wavelengths
are equal.)
Figure 2 shows the computed values of Acr plotted against (ak)2, and compared
to the asymptotic expression (7.8). It will be seen that the shift in frequency agrees
well with the approximate formula up to about ak = 0.30. At higher values of
|n| the relative change in frequency (Acr/cr) is comparatively small.
8. T h e eigenfunctions
where £, tj, Р are the normalized values of the perturbations, and e is a small
constant. Thus £, i), F are equal to etj, eP. The expression (8 .1) represents the
~~2
1796
first order expansion of (3 .1) about yjr = 0 . A finite value of e is necessary in order
to distinguish between the members of each pair, but since quantities of order €2
are neglected there has to be some distortion in the representation.
Nevertheless two predictions can be verified: (1) that the local wavelength of
the perturbations is shorter in the crests of the unperturbed wave than in the
trough; (2 ) that the amplitude of the perturbation is greater at the crest of the
unperturbed wave than in the trough. The local wavenumber, according to
Longuet-Higgins & Stewart (i 960 ), is approximately
|?t| ( 1 +afccosfoc) (8 .2 )
and the local amplitude is
a(l + akcoskx) (a = const.), (8.3)
so that the ratio of the maximum to the minimum values of each quantity is
(1 +ak)/(i—ak). (8.4)
Both predictions (8.2) and (8.3) are well verified by the waves in figure 3.
9. B e h a v io u r a t la r g e v a lu e s o f ak
Figure 1 indicates that the frequencies cr of all the normal modes except n — 1
diminish monotonically as ak increases, at least until ak = 0.42. Beyond this value,
convergence becomes difficult to obtain by the present method.
The mode n = 2 is especially interesting, as it appears to be approaching the axis
cr = 0 the most rapidly. If the frequencies сгг and <r2 of modes n = 1 and 2 coalesce
at a critical wave amplitude a = a12, say, and if
о| ~ K(a12- a ) (К > 0), ( 9 .1 )
then we would expect that, when a > a12, <r2 would become imaginary, indicating
an unstable mode. Figure 4 shows a plot of cr\ against ak at high wave amplitudes.
Though the computations cannot yet be carried beyond ak = 0.42, an extra
polation of the known values certainly suggests that o\ is tending to zero at some
wave amplitude ал2 less than the maximum. It also appears that 2 is close to, if
not coincident with, the amplitude corresponding to the maximum phase speed.
The occurrence of an instability at this wave amplitude might be understood as
follows. Consider a particular perturbation of a symmetric wave which is such as
to change the unperturbed wave to a symmetric wave of slightly greater amplitude.
In general, since the phase speed is a function of the amplitude, the perturbed wave
will travel at a slightly different speed from the unperturbed wave. Hence the
small perturbation (i.e. the difference between the perturbed and unperturbed
waves) will not be a normal mode, since it began by being symmetric about the
crest, and gradually developed asymmetry. However, at the particular wave
amplitude at which the phase speed is a maximum, the perturbation will remain
1797
484 M. S. Longuet-Higgins
symmetric, and will have zero frequency. Thus, at any stationary value of the
phase-speed a new type of normal mode, with zero frequency, becomes possible.
This leads us to expect that points on the axis a = 0 at which the speed is
stationary will be critical points where the modes will branch or coalesce, and at
which new and unstable modes may first appear. (This does not exclude the possi
bility that some critical points will occur when |cr| > 0 .)
wave amplitude, ak
F ig u r e The squares of the radian frequencies <rx and <rt for the two lowest
4.
modes, at high values of the dimensionless wave amplitude.
A p p e n d ix . C o n s t r u c t i o n o f th e m atrices A and В
where
2Un = 2PnC0 + (Pn+l+-Pn-i) ^l+(Pn+2+Pn-2) C2+ ...Л
2T^ = 2Qn D 0 + (Qn+1 + Q n_ i) Dx + ( § n+2 + Qn-2) ^ 2 + •••> / (A 8 )
2Wn — (Rn+l —i?n_x) Dx+ (Rn+2—Rn_2) + •••J
and we have defined
P-i = Pi> Q-j = Qi> R-i = — (A 9)
486 M. S. Longuet-Higgins
and from (A 8 ) and (A ll)
= %Pn C0+ (i ^+1 + Px_j) Сг + [Pn+2+ Рп-ъ) C2+ ...
+ 2^nZ)0 + Z)x+ 2^n_zZ)2+ ... + Q0Dn. (A 16)
On multiplying the above series for Y^} 7^, P, Q, R and S with the Fourier series
(6 .1) and (6 .2 ) for £, tj and F we can now construct the matrices AN and BN of
§ 6 . The general case will be made clear from figures 5 and 6 . These show the layout of
Л3 and B3. The terms whose coefficients are elements of the matrix are indicated
by the row vector (%, ...,d3) above the matrix and by the column vector on the
left. All elements of A3 are zero except those of the submatrices / (1), I{2\ / (3), J(4)
which are as follows:
•P
ьэ
to
/<u =
G 3 GC
2
3
+
C
i+G 3
ъ+х
0
4 2
/(2) =
C
cl c2— o C<-O
< xO t-G0 03- 31
[c ct O
2 03
3 s Co ok-<
c ,-c j
PP
zt
Г2
+ P P,*+PiР*+Рг
P
*P
+ Р„iP
2 2Д 1
»P
+*+P
0
P3+PiР
LР.+Л b+P P j 1
»
=
[P
P»-P
~PiP
2 P
*
-P-P
o0 ш -щ 3 a
- л .
1
U -a P
^— Pl ^-• p j
a
Similarly in figure 5 all elements of BN are zero except those of Ja), ..., J<e), which
are as follows:
"2P0 2PX 2P2 2P3
Jd ) =
2PX P0+ P2+ 2Q'0/с P±+ P3 P2 + P4
2Pt P1+ P3+ 2Q1/c P0+ Pi + 4Q'0/c P1+ P6
.2P3 Рг + Pl + 2QJc Р ^ Р ь + iQJc P0+ P3+ 6Q'0/cJ
a0 ax a2 a3 b0 bx b2 b3 c, c2 c3 dy d2 d3
x7 Vol. 360. A.
1801
488 M. S. Longuet-Higgins
The computations described in this paper were begun while the author was a
guest lecturer in the Institute of Port and Ocean Engineering at the Norwegian
Institute of Technology, University of Trondheim, in March 1977. The author is
indebted to Professor P. Bruun for the kind hospitality of his Department and to
others of his staff, particularly Dr 0. G. Houmb, for many practical arrangements.
R e fer en ces
Benjamin, Т. B. & Feir 1967 The disintegration of wave trains on deep water. 1. Theory.
J. Fluid Mech. 27, 417-430.
Cokelet, E. D. 1977 Steep gravity waves in water of arbitrary uniform depth. Phil. Trane.
R. Soc. Lond. A 286, 183-230.
Levi-Civita, T. 1925 Determination rigoureuse des ondes permanentes d’ampleur finie.
Math. Ann. 93, 264^314.
Longuet-Higgins, M. S. 1975 Integral properties of periodic gravity waves of finite amplitude.
Proc. R. Soc. Lond. A 342, 167-174.
Longuet-Higgins, M. S. 1978 Some new relations between Stokes’s coefficients in the theory
of gravity waves. J. Inst. Math. Applic. (in the press.)
Longuet-Higgins, M. S. <fe Cokelet, E. D. 1976 The deformation of steep surface waves on
water. I. A numerical method of computation. Proc. R. Soc. Lond. A 350, 1-26.
Longuet-Higgins, M. S. & Fox, M. J. H. 1977 Theory of the aJmost-highest wave. The
inner solution. J. Fluid Mech. 80, 721-741.
Longuet-Higgins, M. S. & Fox, M. J. H. 1978 Theory of the almost-highest wave. П .
Matching and analytic extension. J. Fluid Mech. 85, 769-782.
Longuet-Higgins, M. S. & Phillips, О. M. 1962 Phase velocity effects in tertiary wave
interactions. J. Fluid Mech. 12, 333-336.
Longuet-Higgins, M. S. & Stewart, R. W . i 960 Changes in the form of short gravity waves
on long waves and tidal currents. J. Fluid Mech. 8 , 666-583.
Schwartz, L. W . 1974 Computer extension and analytic continuation of Stokes’s expansion
for gravity waves. J . Fluid Mech. 62, 553-578.
Stokes, Sir G. G. 1880 Supplement to a paper on the theory of oscillatory waves. Mathe
matical and Physical Papers, vol. 1, pp. 314—326. Cambridge University Press.
1802
B y M . S. L o n g u e t - H i g g i n s , F.R .S.
Department of Applied Mathematics and Theoretical Physics, University of
Cambridge, Silver Street, Cambridge CBS 9EW, U.K., and Institute of
Oceanographic Sciences, Wormley, Surrey,
1. I n t r o d u c t i o n
In this paper we continue and extend the stability analysis of irrotational water
waves of finite amplitude which was begun in part I (Longuet-Higgins 1978 ). In
that paper we determined the normal-mode perturbations of gravity waves on deep
water, for values of the wave steepness up to about ak = 0.42. Consideration was,
however, restricted to perturbations having the same space periodicity as the basic
wave, or less. The existence of instabilities having a much longer wavelength has
been known for some time (Benjamin & Feir 1967; Benjamin 1967). Our present
aim is to obtain an overall view of the instabilities, including scales both shorter
and longer than one basic wavelength.
[ 489 ] 17-2
1803
490 M. S. Longuet-Higgins
We show first in § 2 how the calculations of the subharmonics can be carried out
by a relatively simple modification of the analysis of part I. The simplest case,
when the perturbations have a horizontal scale of twice the wavelength of the
unperturbed wave, is studied in §3 (see figures 1 and 2 ). This reveals at once the
presence of modes which are unstable over an intermediate range of wave steepness.
By examining a sequence of modes of increasing length scales (§§4 and 6 ) we are
able to delineate the boundary of the unstable zone (figure 7). It is shown that the
lower left hand part of the diagram corresponds to the Benjamin-Feir instability
(§ 6 ). However the diagram shows also that above a certain amplitude each of these
instabilities is temporarily suppressed. At a still higher value of ak, namely
ak « 0.41
a new type of instability occurs, which we show might be expected at around the
value of ak for which the first Fourier coefficient in the analysis is a maximum.
A comparison of the calculated growth rates with the available observations
is made in § 7.
2. G e n e r a l m eth od
As in part I suppose that we are given the basic unperturbed wave motion of
wavelength 2 tc as described
Y = \jr/c+ +Hxe^/ccos (ф/с) + # 2e2^/ccos (2ф/с) + ...,|
Х = ф/с +HXe^/c sin {ф/с) +H2e^ lc sin (2ф/с) +
where the constants Hn have been calculated so as to satisfy the nonlinear con
ditions at the free surface \{r = 0 . Since the only physical constant is g the accelera
tion of the gravity (which we take as equal to 1) there exists a precisely similar
solution in which the wave height and length are each reduced by a factor m,
namely
r = f/ c' + \Щ + я ; cos (ф/с') + Щ e ^ cos (2ф/с') + ...
and similarly for X', where
я , = ( Я1"ЩПIm. n/m integral| (2 2)
ft \o otherwise J
and the new phase velocity c' is reduced by mb\
c' —m~bc. (2.3)
The coefficients C'n in the expansion of Y'^ are
C'0 = 1/c = mtC0 (2.4)
Tjt t t ( miCnlm, n/m integrall
and in general C„ = nE Jc = otherwise. j <2 6)
We now have a basic wave with precisely m wavelengths in the interval 0 < X' < 2 tc,
and we may proceed, exactly as in part I, to find all the perturbations of this wave
having length scales 2n or less. These perturbations are described by equations
1804
3. R e s u l t s : ra = 2
For the odd modes, the eigenfrequencies were found to converge at all values of
the unperturbed wave amplitude up to about ak = 0.42. The results for both odd
and even modes are shown together in figure 1.
The branches have been labelled according to the wavenumber of the dominant
harmonic at low values of ak. Thus n = 6 /2 corresponds to a mode which, at low
ak, has a wavelength 2/5 of the basic wavelength 2n. The even modes correspond
1805
492 M. S. Longuet-Higgins
to even values of the numerator. Thus the modes n = 2/2 and n = 4 /2 have already
appeared as n = 1 and n = 2 in fig. 1 of part I. (For clarity, the vertical scale is
here doubled.) The new modes are n = 1/ 2 , 3/2 and 5/2.
When ak-+ 0 the radian frequencies of the odd modes, like those of the even
modes, tend to the limiting values
ar=\n±ni\. (3.1)
Thus for the three lowest odd modes (n = 1/ 2 , 3/2, 5/2) the frequencies are res
pectively 0.2071, 0.2753 and 0.9189.
wav© steepness, ak
Figutue 1. Radian frequencies of the normal-mode perturbations having twice the wavelength
of the fundamental wave, shown as functions of the unperturbed amplitude ak.
A8 ak increases from zero, the steepness of the basic wave begins to affect
the frequency of the perturbations. It was shown by Longuet-Higgins & Phillips
( 1962 ) that the speed Cj of a gravity wave of wavenumber kx (relative to the
water at infinite depth) is affected by the presence of a second wave of amplitude
a2 and wavenumber k2. To order (a2k2)2 they found
_ Г± (y/kift—(a2^2)2ca ^1 > ^2*1 /g 2 )
^ i {я/к^ —{аг ^ 2с2{к1/к^) fcj < k2,J
1806
assuming the second wave is travelling in the negative direction. By the same
argument as in § 7 of part I, this leads to a change in the frequency of the first wave,
as seen in a frame of reference travelling with the second. Identifying kx with
|n|and a2k2with ak we obtain
It will be noticed that when |n| = \ or 0 the coefficient of (ak)2 vanishes. This
accounts for the zero curvature in the branches n — 1/2 and n = 0 at their junctions
with the vertical axis.
wave steepness, ak
F igure 2. Rates of growth of the normal-mode perturbations having
twice the wavelength of the fundamental wave.
1807
494 M. S. Longuet-Higgins
As ak increases from zero, it appears from figure 1 that the two branches n —3/2
and 7i = 1/2 coalesce, and the joint eigenfrequency is then of the form
cr = ± a ± i P
where a and p are real and we take p ^ 0. The mode corresponding to the positive
sign is then unstable. This instability, as we shall see later, is of the Benjamin-Feir
type.
As ak increases still further, however, stability is unexpectedly re-established
at about ak = 0.366. Thereafter, the frequency of the upper branch remains almost
constant at about or = 0.20, but the frequency of the lower branch continues the
downward trend begun by the branch n = 3/2 before it coalesced with n = 1/2.
Ultimately, at about ak = 0.406, the lower branch meets the branch n = 2/2,
which as we saw in part I represents a perturbation which simply shifts the phase
of the basic wave by a constant. At the junction a new instability springs up, with
a much higher rate of growth, as seen in figure 2. Though the corresponding eigen
function is less well determined than the eigenfrequency, it does appear that this
second instability is much more localized than the first, and may represent the
initial stage of a plunging breaker.
In part I we showed that the junction of the modes n —4/2 and 2/2 probably
coincided with the calculated maximum in the phase speed с as a function of ak.
The junction of n = 3/2 and n = 2/2 is clearly at a different point. How are we to
understand the occurrence of an instability at this lower wave amplitude?
At the critical point the frequency of the perturbation is zero, so that the per
turbed wave travels along with the same phase velocity as the unperturbed wave.
But the new perturbation is odd, not even. The perturbed wave may therefore
represent a train of waves of alternately higher and lower amplitude. Since we are
not at a velocity maximum, however, the natural increase in velocity of the higher
wave due to its greater steepness must be exactly counterbalanced by a decrease in
velocity owing to a shortening of its wavelength. What type of perturbation is
required in order to produce simultaneously an increase in height and decrease in
length of every alternate wave, and a decrease in height and increase in length of
the intervening waves ? On reflection it becomes clear that the perturbation must
have (at least) three maxima and three minima in every two wavelengths. Hence
we would expect the wavenumber n = 3/2 to be dominant.
Confirmation of this interpretation comes from the remark that the perturbation,
being odd, contains no even harmonics, in particular no harmonic n = 2/2. Now the
harmonic n = 2/2 = 1is simply the first harmonic of the fundamental, unperturbed
wave train. If the height of every alternate wave is to be increased without adding
to it any first harmonic the magnitude of the first harmonic must clearly be stationary
with respect to any increase in the wave height. But Schwartz ( 1974) foimcl by
computation that the amplitude a± of the first Fourier coefficient of a regular wave
train does indeed attain a maximum value, and diminishes again before the height
1808
is a maximum. From figs. 3 and 8 of Schwartz’s paper it appears that the wave
steepness for which the first harmonic is a maximum is given by
2,a/L = 0.131.
This is equivalent to ak = 0.412, which is not far from the amplitude ak = 0.406,
where the onset of the new instability is found.
The above argument is not exact, because any alteration in the perturbed wave
train cannot be precisely the same as in a uniform train of waves; there must be a
slight distortion of the symmetry in order that neighbouring waves shall join
smoothly. Moreover, our argument would apply equally to all the higher har
monics, which attain their maxima at nearly but not quite the same wave amplitude
(see Schwartz 1974, fig- 8).
When m ^ 3 the normal modes may be separated into mutually exclusive cate
gories in a somewhat analogous way. Let I be any integer such that
0 <l<\m. ( 4 .1 )
From equations (2 .2 ) and (2 .6 ), all the coefficients Я* and Cn will vanish except
those for which n is a multiple of m (see figure 3). From the addition rules for sines
and cosines it follows that all the coefficients Pn, Qn, Rn, Sn defined in (I) will also
vanish except when n = jm ,j = 0 , 1 , 2 , . . . . Hence the only non-vanishing coefficients
in the matrices A and В will be those corresponding to the elements
and to the coefficients of cos пв, sin пв when n = jm ± I. The situation is illustrated
in figure 3 for the case m = 7, I = 2 . Each value of n = jm for which a non-zero
coefficients exists is flanked by two values n = j m ± l for which there may be a
non-zero element of the eigenvector (Oj,...,dN) with the exception that when
m —0 , there is only one value n = I.
So, if we include, at the iVth stage, only the coefficients of cos пв and sin пв having
n ^ Nm, we obtain matrices Л * , B* of order 8N precisely, from which to determine
the normal modes. (When m = 2 , I = 1, the rows and columns coincide in pairs,
and the matrix is of order 4iV only.) This doubling of the order, from 4>N to 8N
I I I I 1 1
m m m
F ig u r e Harmonic components of the class of perturbations having wavenumbers
3.
n = \pm + ql\, where m and I are fixed and p and q are running integers.
1809
496 M. S. Longuet-Higgins
means that to obtain a comparable accuracy, at any given value of N, almost
four times the core-storage is required compared to the requirement for calculating
the superharmonics. Nevertheless convergence of the eigenvalues to at least three
significant figures (and generally many more) was obtained for wave amplitudes
up to and including ak = 0.40.
6. R e s u l t s w h e n m > 2
In order to check the programming of the calculation in Fo r t r a n iv it was found
useful first to try the case m = 4, knowing that the modes corresponding to m = 4,
1 = 1 should be among the complete set of modes for m = 4. (For moderate ak, the
complete set can be found by a much simpler modification of the existing program
for the superharmonics.) This being verified precisely for moderate ak it was then
possible to carry out confidently the calculations for larger values of the parameter.
wave steepness, ak
Radian frequencies o f subharmonics corresponding to the case
F ig u r e 4.
m = 4 . Circular plots correspond to calculated values.
1810
decreases from 0.586 to 0 . Then cr_ increases again to a maximum value 0.25 at
8n = 2 , before decreasing again towards zero in the limit n -+ 0 .
For every I less than m it appears that the modes n = l ± l /m coalesce to form an
unstable mode. As ak increases, this mode becomes stable again. Then one of the
stable branches descends to the axis (n — 8 / 8 ) where a new instability arises.
The corresponding rates of growth are shown in figure 6 . It will be seen that the
maximum growth rate in the Benjamin-Feir type of instability is limited to
P = Im (<7) < 0.0234. (5.2)
This value occurs with the mode n = ( 1/ 2 , 3/ 2 ) when the basic wave amplitude
ak « 0.32.
The second type of instability grows much more rapidly. For practical reasons
(see § 4) the rates of growth could be determined accurately only in the degenerate
case m — 2 and when n = ( 1/ 2 , 3/2), that is to say 8n = (4, 12 ). At an amplitude
ak = 0.42, for example we find
p = 0.126 (5.3)
498 M. S. Longuet-Higgins
or an increase of 121 %. (The increase corresponding to (6 .2 ) is only 14%.) The
trend of the curve on the right of figure 6 suggests that at larger values of ak the
rates of growth are still higher.
wave steepness, ak
F i g u r e 6. Radian frequencies of subharmonics when m = 8.
The points in figure 5 where pairs of modes coalesce and become unstable all
lie on a continuous locus as shown in figure 7. This forms the boundary between
stable and unstable parts of the diagram in the sense that in the ‘ unstable’ part
there is always some unstable mode at the given frequency and amplitude. The
boundary on the left appears to extend to the origin, and the tangent there, as we
shall see, corresponds to the Benjamin-Feir theory.
The boundary on the right of the unstable region is the locus of points in figure 5
where the tangent to the right hand branch of each curve is vertical. For the pair
1812
wave steepness, ak
F i g u r e 6. Growth rates of subharmonics when m = 8 .
upper branch n = 2 —l/m the curvature becomes very sharp, indicating only a
slight degree of interaction. At a value of l/m marginally less than 1 /8 the two
branches appear to intersect, so that the left and right hand boundaries of the
region of instability are joined.
At its lower end, the right hand boundary of the unstable region appears to meet
the axis cr = 0 at about the point where ak = 0.347. It is to be noted that by an
application of Whitham’s average-Lagrangian technique, Lighthill ( 1967 ) found
1813
500 M. S. Longuet-Higgins
that for very slowly varying wave trains at finite amplitude, the Benjamin-Feir
instability should appear when 2a /L — 0.054 approximately. This gives ak ж 0 .34 ,
in good agreement with the present value.
wave steepness, ak
F igu re 7. Stability diagram for gravity waves in deep water. The tangent at the origin
corresponds to the asymptotic theory of Benjamin & Feir ( 1967 ). (The resemblance to a
breaking wave is coincidental.)
6. T h e B e n j a m i n - F e i r i n s t a b i l i t y
The analysis proposed by Benjamin & Feir ( 1967 ) to explain the observed dis
integration of wave trains in laboratory channels envisaged a primary, deep-water
wave of finite amplitude a and radian frequency 0) on which was superposed two
‘ side-bands’ of infinitesimal amplitude ea and of frequencies cu(l ± 8) respectively.
Their analysis indicated that the side-bands would tend to grow in amplitude at
the expense of the primary waves provided that
2{akf > (6 . 1)
The ultimate rate of growth of the side-bands was given by
d(lne)/d* = Zo)8\a?/al- 1)*, (6 .2 )
where a0 denotes the critical amplitude 8a>/k<j2.
The effect of the side-band amplification is to produce a modulation of the wave
envelope with a spatial wavenumber Aк related to $ by the group velocity:
8a)/Ak = dcr/dk = (o/2 k. (6.3)
1814
Л — + <6-6)
for the complex wave envelope eA (^r) of a slowly varying wave train on deep
water.
Here £ and r are the dimensionless variables
£ = ek(x—tdcr/dk), т= e2t, (6 .6 )
and quantities of order e3 are neglected. On assuming a slightly modulated wave
train
A = a { l + a ') e t f , (6.7)
where a' and x' are small perturbations proportional to exp (iAк£ + /3т), one finds
after substituting in (6.5) and considering real and imaginary parts, that
^ = i(A ^ {W -i(A ^ } (6 .8 )
or from (6.4)
p = al(a2/a l ~ l)i, where a\ = J(Ak)2 = \82. (6.9)
So, as before, the motion becomes unstable when (ak)2 > \82. The calculated rate
of growth is similar.
The above theory is vaUd only asymptotically for sufficiently small values of the
wave steepness ak and of the modulation frequency 8 and wavenumber Ak. Note
that in a frame of reference travelling with the phase speed the apparent frequency
of the perturbations will be given by
A<r « \Ak±a\(a2/a l — 1)*. (6.10)
Now the two wavenumbers of each pair of modes which coalesce in the bottom
left hand comer of figure 5 are
n = t ± l / m = l ± A k = 1±28 (6.11)
according to equation (6.4). So to compare with the asymptotic theory we should
write
or = 28 = Ak = |w—1|. (6 . 12)
and from (6 . 1) the critical wave amplitude a0 is given by
<r = V2a 0 fc. (6.13)
Equation (6.13) corresponds to the straight asymptote through the origin in
figure 7. The calculated curve appears to touch this line at the origin. The deviations
from the line are of order {ak)i, however, not ak)2, so that in improving the asymp
totic theory some care should be taken with the expansions in powers of ak or 8.
1815
502 M. S. Longuet-Higgins
To account for this, we remark that the instability can be regarded as a resonance
brought about by the effect of the finite-amplitude wave on the frequencies of the
two side-bands (see Benjamin & Feir 1967, § 1). Explicitly, we have only to equate
the two expressions (3.3) for the velocity, the first with kx = 1 + Дк and the second
with kx = 1 —Ak, to obtain a result equivalent to (6.13). This is because for small ak,
the second-order corrections to the phase velocities are valid over the greater part
of the interval 0 < a < a0. However, in the neighbourhood of the critical point the
curves are approximately parabolic, introducing terms like (a0—a)t. It is these
singularities that are responsible for the deviations of the boundary from its
asymptote being of a lower order than suggested by Benjamin ( 1967 ).
7. C o m p a r i s o n w it h o b s e r v a t io n
Benjamin ( 1967 ) has reported experiments on the spatial rate of growth of side
band instabilities in a train of waves generated in a laboratory channel. The
quantity plotted in his fig. 4 is d(lne)/d#, where z denotes kz/2n. To compare
with our results this must be converted to a time rate of change. We must assume
that d/dt ~ cgb/bz, where cg denotes the linear group velocity, although for finite-
amplitude waves the assumption is only approximately correct. Since in our units
cg = \ this implies that
d(lne)/d5 = 4Ttd(lne)/d< = 4я1т(о-). (7.1)
The experiments were carried out at a constant side-band frequency 8 = 0.10
and variable wave steepness ak. In figures 5 and 7 they therefore correspond to
points along the line гг = 0.10 parallel to the axis of ak. Our calculations of growth
rates in figure 6 were, however, carried out at fixed values of the nominal wave
number n. To obtain the comparable rate of growth we must first find the appro
priate (fractional) value of n and then interpolate in an appropriate manner.
First from figure 7 we see that the critical wave steepness a0k corresponding
to a frequency <r = 0.10 is a0k = 0.083 (as compared to 0.071 on the asymptotic
theory). For any given value of the wave steepness ak > a0kvre can determine the
appropriate value of n by interpolating along a line parallel to the stability boundary
in figure 7 . Thus when 0.083 ^ ak ^ 0.20, we find values of n ranging monotonically
from 9.63 to 9.75.
Now in figure 8 we have plotted the calculated values of
7} = Im(cr)/al against £ = а/я 0 (7.2)
at different values of n. The uppermost curve corresponds to the limiting value
8n = 8 (or to = 1). From equation (6.9) this curve is simply 7) = (f 2 —1)*. The remain
ing curves show the extent to which the calculated growth rates depart from the
asymptotic theory. For any given value of ak we know the corresponding values
of £ and n so we may easily interpolate to find the appropriate value of rj. Then
from (7.1) and (7.2) we estimate
d(lne)/d# = 4na%7j. (7.3)
1816
wave steepness, ak
F ig u r e 9. Comparison of the calculated growth rate (solid curve) with the measured rates of
growth (plotted points) as reported by Benjamin (1967 ). The broken line represents the
asymptotic theory of Benjamin & Feir (1967 ).
1817
504 M. S. Longuet-Higgins
The resulting curve for d(lne)/dx is shown in figure 9 (solid curve), compared
with the measured rates of growth according to Benjamin ( 1967 , fig. 4 ), and the
asymptotic theory of § 6 .
It will be seen at once that the present calculations agree rather well with the
observed growth rates, indeed more closely than the asymptotic theory with which
they were originally compared. The part of the axis to the left of the critical point
is strictly part of the theory, representing zero rate of growth. This passes through
the first plotted observation. The second, representing a positive growth rate, also
lies on the theoretical curve. The remaining points all lie somewhat below the curve,
perhaps on account of dissipation, which we have neglected.
The foregoing measurements all correspond to low values of the amplitude:
ak < 0.17. In preliminary experiments at higher wave amplitudes, using a wave-
maker of the conventional wedge or plunger type, the author has found that
steeper waves tend to become irregular, hindering accurate observations. No doubt
this is directly due to the fact that unless the wavemaker exactly matches the
particle-motion in a progressive wave, transient motions are generated which
quickly turn into instabilities as the waves progress.
One method of overcoming these difficulties and of reaching higher amplitudes
which may be more stable would be to design a wavemaker so as to match more
closely the nonlinear motions in steeper waves.
An attractive alternative is to conduct numerical experiments, possibly with the
forward integration technique devised by Longuet-Higgins & Cokelet ( 1976 ),
inserting the appropriate initial conditions. Investigations on these lines will be
reported in a subsequent paper.
8. D is c u s s io n
and so can grow even though the initial amplitude ak is less than the amplitude of
maximum energy for the average wave train.
In this investigation we have so far treated only inviscid, irrotational gravity
waves on deep water because this is the simplest case with the fewest possible
parameters. Physically the model may apply best to surface waves long enough for
surface tension and viscosity to be negligible. Nevertheless the analysis clearly
needs to be extended so as to include capillarity and the important effects of surface
shear.
In water of finite depth h the work of Whitham ( 1967 ) and Benjamin ( 1967 )
suggests that the subharmonic instabilities may be suppressed at sufficiently small
values of kh. On the other hand from observations in shallow water one would
expect the local, superharmonic instabilities to be enhanced. Further investigations
of these effects are clearly desirable.
R eferences
1. I n t r o d u c t io n
The task of understanding and predicting the formation of breaking waves on deep
water has recently been approached in two different ways. On the one hand a
general method has been developed for calculating the deformation of the free
surface in any irrotational motion which is periodic in the horizontal coordinate
(Longuet-Higgins & Cokelet 1976 ; referred to as paper I). This has been applied
successfully to a variety of initial conditions (see I and also Cokelet 1977 ). On the
other hand a more analytical approach (Longuet-Higgins 19786 , c) has yielded some
understanding of the initial rates of growth of any small perturbations to a uniform
train of waves o f finite amplitude.
The conclusions reached in the more recent papers (Longuet-Higgins 19786 , c)
were somewhat surprising. For instance it was found that there were at least two
distinct types of instability: subharmonic instabilities of the Benjamin-Feir type,
t Part I appeared in Proc. R. Soc. Lond. A 350, 1-26.
In the perturbation analysis of Longuet-Higgins (1978 6 ,c) the normal modes were
calculated in the form ж _ Х{ф> f ) + ^ f ) e_lrt
У = Y (ф, ф) + е Щ , ф) e~i<rtJ 1*
where ф and \jr denote the velocity potential and stream function in a frame of
reference moving with the speed с of the unperturbed wave, and x ,y are rect
angular coordinates (x horizontal, у vertically upwards) expressed in terms of ф, rjr
and the time t as independent variables. X and Y represent the coordinates of the
unperturbed wave, and £, fj the normalized perturbation, e being an arbitrary small
parameter. The free surface, which in the unperturbed wave is given by \jr = 0 , is
given by f = е /(ф )е ~ м (2.2)
in the perturbed state. In fact £, ij and/have both real and imaginary parts, corre
sponding to the ‘ in-phase’ and ‘ quadrature’ components of the normal mode, and
each of these components is given, in turn, by Fourier series in cos {1ф/с) and in
sin (1ф/с), where I is a non-negative integer and фruns from 0 to 2mn/c.
Throughout this paper the units of length and time are normalized by taking
g = 1 and the wavelength of the unperturbed wave to be 2 k .
It was shown by Longuet-Higgins ( 19786 , c) that each normal mode may be
designated by one or more rational numbers (mode numbers) which indicate its
principal wavenumber components when ak is small. For example n = (£, J) de
notes a subharmonic normal mode with principal components having wavenumbers
\ and §. An index + or — may be added to designate that the perturbation is
either growing or decaying, respectively.
In general cr is a complex quantity, and the frequencies occur in conjugate pairs.
The unstable modes are those for which <r has a positive imaginary part. The cal
culations revealed two distinct types of instability. The first are disharmonies, with
wavelengths greater than that of the unperturbed wave. At small wave steepness
ak these are clearly of the Benjamin-Feir type (Benjamin 1967). If we imagine the
side-band frequency Д<r as fixed and the steepness ak of the unperturbed wave to
increase gradually, then these subharmonic modes first become unstable when ak
slightly exceeds ДсгД/2 . The maximum rate of growth is found for a normal per
turbation two wavelengths long, at an amplitude ak ж 0.32. But when ak slightly
exceeds 0.37 the perturbation apparently becomes neutrally stable again: the
instability disappears (a result first predicted by Lighthill ( 1967)). Beyond ak = 0.37
the normal-mode analysis of Longuet-Higgins (1978 c) indicates a narrow range of
neutral stability, culminating in the onset of a second type of instability which (it
was suggested) is more local, being concentrated near the crests of individual waves.
For this type, cr is pure imaginary, and the initial rate of growth is much higher.
It was foreseen that these instabilities might develop directly into breaking waves.
1-2
1822
3. T he t i m e -s t e p p i n g method
In the present paper we aim to test and extend these results by the entirely
independent time-stepping method developed in I. This method is based on the
fact that in an irrotational, incompressible fluid the kinematic and dynamic
boundary conditions describe the evolution of the flow in terms of quantities speci
fied only on the boundaries. For deep-water waves, the only boundary is the free
surface, the region of greatest interest. Its location is given by the coordinates,
= 1, 2 of specified fluid particles. In order to follow the motions
of the surface we need know only the tangential and normal velocities of the par
ticles on it. Given the value of the velocity potential ф* along the surface we find the
tangential velocity 00*/0s by numerical differentiation. To find the normal velocity
00*/0n we use Green’s third identity ((4.2) of paper I) to derive an integral equation
for дф*/дп in terms of other surface-evaluated quantities. Having solved this
equation numerically we march forward in time with the evolution equations
((2.7) and (2.10) of paper I).
The motion is assumed to be periodic in the horizontal coordinate, though not
necessarily in the time t. This assumption enables us to transform the semi-infinite
region of fluid ( ж * , у * - р 1 а п е ) into a finite enclosed region (£-plane) such that the
free surface maps to the exterior boundary (see § 3 of I). The basic numerical
computations are performed in this transformed plane.
Evidently the time-stepping method can be adapted without essential change to
the present problem, provided the spatial period L contains m wavelengths of the
unperturbed motion (ra an integer). In the present paper we confine ourselves to
the case ra = 2 , first because the initial perturbations, as calculated by the normal-
mode method (Longuet-Higgins 1978 b, c), are given most accurately when ra = 1
or 2 , at the larger values of ak. Secondly, the case ra = 1 excludes subharmonic
instabilities and yields only the localized instabilities; and these occur at slightly
higher values of ak than when ra = 2 so the accuracy is correspondingly less. By
considering ra = 2 we shall include both types of instability, and with greater
initial accuracy.
A third advantage of the case ra = 2 is that, of all integer values of ra, this yields
the subharmonic instability having the highest growth-rate, namely when
n = (h$)> ak = 0.32.
In the time-stepping method the reference frame is stationary, relative to the
water at infinite depth. So if we denote the velocity-potential and stream-function
by ф*, ifr*, and the rectangular coordinates by x*, y*, we have the relationsf
x* = x + ct, ф* = ф+ сх*Л
У* = У, т
/r* = i/r + cy*.)
4. E xam ple
t= 0
Ь -
t=\T
F igube 1 . Development of the subharmonic instability $)+ over one half-period of the
perturbation, when ak = 0.32. The left hand column shows the time-stepped surface
profiles. The right hand column shows the resulting perturbation, enlarged vertically
by a factor 5. Time increases downwards. In this case N = 90, e = 0.025.
1825
О х 4тг О х 4тг
F ig u b e 2. As in figure 1, but with e = 0 (zero perturbation). This tests the accuracy of the
time-stepping technique. On the right, the vertical scale is enlarged times 20 .
To determine the rate of growth of the perturbation h(x*, t) over a given interval
of time, say (0 , t)ywe aim to determine the ratio J?(i) of the root-mean-square values
of h(x*t t) and h(x*, 0). By definitionf
It is also useful to define the correlation coefficient between h(x, t) and h(x + * 0. °) by
n, « \h(x,t)h(x + x 0>0 )da:
C(xQ,t) = ----------------------- ;-------------------------------- (5. 2)
[{W ^, 0 }2c b j + 0 )}2da:]*
t In this section x and у are written for x* and y*. This should cause no confusion with the
notation of 5 2 .
1827
We might expect the maximum correlation between h(x,t) and k(x + x0, 0 ) when
x0 = ct, where с is the phase velocity. However, since ^e-I<rt has both a real and
an imaginary part, corresponding to components in-phase and in quadrature with
the initial perturbation, the maximum (absolute) correlation will in general fluc
tuate in time. If T denotes the period of the perturbation (T = 2л/R e (<r)) then at
times t = sT , where s is an integer, we should expect C{ct, t) to be near unity; while
at times t = (s + ^)T we should expect C(ct, t) to be nearly —1. Provided the maxi
mum value of |C(a:0,<)| is near unity, then a useful measure of the growth-rate
should be derived from the corresponding value of jR(f). In fact we would expect
that while the perturbation is still small
Я (0 * « Л p = Im(o'). (5.3)
The actual determination of the integrals in (5.1) and (5.2) must be done with
care. The computer program begins at t = 0 , and has its own method of determining
the time-step length, At. In general the flow will not be computed at exactly
t = \sT , but rather at times before and after this. Hence it is necessary to inter
polate the particle positions to this time. This can be done in a variety of ways.
Let us assume we have a simple equation of the form
dy/dt = /( 0 , (5.4)
and we store у and / at every time-step. Our time-stepping technique is fourth-
order (local errors ~ (A<)5) which means it would be consistent to use five pieces
of information about у and / for interpolation. However it seems a good idea to
use information only at times tx and t2 which bracket the time of interest, t. This
implies a cubic interpolation formula (errors ~ (ДО4) based on y v y2, f x and f 2
(here subscripts denote time levels). We tried interpolating the particle positions
using such a formula, and it was clear that the results were ‘ noisy*. A linear inter
polation (errors ~ (At)2) based on just the particle positions themselves was much
smoother. This is to be expected since the displacements are calculated from the
time derivatives by integration, a smoothing process. In order to have a smooth
interpolant y(t), and to make use of both the displacement and derivative infor
mation at tx and t2, we took the average of two quadratics each fitted through tx and
t2 but with one using f x and the other using f 2. Thus we have
d = - m \ + Ч) - № + « ■ ,
This interpolation formula is exact for quadratics and has local errors which vary
as (At)3.
1828
F ig u b e 4. The correlation coefficient C{ct, i) between k(x, () and h(x + ct, 0) as a function of
the time t for the perturbation (-Jr, f)+ when ak = 0.32 (see figure 1).
T a b l e 1. C o m p a r i s o n o f c a l c u l a t e d g r o w t h -r a t e s o f n o r m a l - m o d e
pertur batio ns
time-stepping with
normal-mode calculation integral equation
, ------------------------------- * -------------------------------4
i-------------- A--------------*
* \T fi ft'
ak c(k/g)* n Re (<г) (я/а) Im (cr) C(cl, t) m t~l \nR
.2072 15.17 .0000 -.9 9 9 0.999 - .0001
0.10 1.005013
ii .2676 11.74 .0000 - .9 9 7 0.995 - .0004
.2098 14.97 .0000 -.9 9 6 1.002 .0001
0.20 1.020203
(I .2406 13.06 .0000 -.9 9 0 1.000 .0000
fib i)+ .2150 14.61 .0132 -.9 9 2 1.219 .0135
0.25 1.031746
НЫ Г .2150 14.61 -.0 1 3 2 -.9 8 0 0.833 -.0 1 2 5
/(*. i)+ .1931 16.27 .0234 -.9 6 9 1.512 .0254
0.32 1.052512
Hi. i>- .1931 16.27 -.0 2 3 4 -.7 8 6 0.752 -.0 1 7 5
.1804 17.41 .0000 -.9 9 4 0.985 - .0009
0.38 1.074399
(f .1309 24.00 .0000 - .8 5 3 0.909 - .0040
0.40 1.082225 * .1840 17.07 .0000 -.9 0 8 1.118 .0065
0.41 1.086045 (*)+ .0000 oo .065
6. R esults
Similar calculations were carried out for the modes n = (£, j ) at the unperturbed
wave amplitudes ak = 0.10, 0.20, 0.25, 0.32, 0.38, 0.40 and 0.41. For the normal
modes under consideration the perturbation theory (Longuet-Higgins 1978 c) pre
dicts that at the two lowest values of ak the waves will be stable, at the next two
they will be unstable, at the next two they will be stable again, and at the highest
value of ak they will be violently unstable. This general behaviour is confirmed by
our computations. Figure 5 shows a graph of the growth-rate /3 plotted against the
unperturbed wave amplitude ak. The solid curve represents the growth-rates
according to Longuet-Higgins (1978 b), and the points represent the present results.
The surprising prediction of a return to stability with increasing wave amplitude
is confirmed reasonably well. Departures of the observed growth-rates from the
predicted ones are probably due to computational inaccuracies, although the finite
amplitude of the perturbation may play a role. The results for the highly unstable
mode at ak = 0.41 are not shown on figure 5 but will be discussed later. In all
cases except ak = 0.32 we took e — 0.0125.
Table 1 lists the quantitative results of the calculations. The first six columns
give the unperturbed wave amplitude and phase speed, the mode of the pertur
bation, its frequency in a reference frame moving with the unperturbed wave, the
semi-period of the perturbation and the growth-rate. These quantities are all
derived from the normal-mode theory. The last three columns give the maximum
correlation coefficient, the r.m.s. amplitude ratio and the observed growth-rate,
all evaluated at t = \T. We shall now discuss the results in detail beginning with
the lowest waves.
1830
]
E- - - - - - - - - - - - 3- - - - E— - - - - - - - - H
E- - - - - - - - - - - - 3- - - - E- - - - - - - - - - - - - 3
E- - - - - - - - - - - - 3- - - - E— - - - - - - - - — j
E- - - - - - - - - - - - 3- - - - E- - - - - - - - - - - 4
E- - - - - - - - - - - - 3- - - - E-- - - - - - - - - - - - - 3
------------------------------------------- -------] t= \ T J----------------------------------- ------------- =
O x 4тг 0 x 4i r
F ig u r e 6. The mode n = $ when ak = 0.10 followed over one half-cycle of the perturbation.
The left hand column shows the time-stepped profiles. The right hand column shows the
resulting perturbation enlarged vertically by a factor 20. Time increases downwards.
e = 0.0126.
is to be expected since this mode is exactly a sine wave in the limit as ak -> 0 (which
is the basis for the notation n = J, signifying one sine wave in two fundamental
wavelengths). Likewise the profile of the mode n = J in figure 7 is predominant^
three sine waves.
The profiles for ak = 0 .20 , n = J and f are illustrated in figures 8 and 9. Com
paring these to figures 6 and 7 we see that the increased nonlinearity of the un
perturbed waves enhances the higher harmonic content of the perturbation profiles,
local bumps and dips begin to appear. The correlation coefficients are —0.996 and
— 0.990, and the growth-rates are 0.0001 and 0.0000 for n = \ and f respectively.
This extremely good agreement with theory is heartening when it is realized that
1832
0.05
t—o
-0.05
10 .05.
t= 0
J- Q 0 5
4i r 0 4 it
The modes (|, f)* at ak = 0.32 have already been discussed in § 3. (£, f)+ is of
particular interest since it represents the fastest-growing Benjamin-Feir mode.
One of the unexpected results of the normal-mode theory is that a wave whose
steepness is near ak = 0.38 will be stable to the modes n = £ and n — |. The results
given in table 1 and plotted in figure 11 certainly suggest this. The correlation
coefficient equals —0.994 for n = and there is a very slight decay in amplitude.
We observe a lower correlation coefficient and an increased rate of decay from
w = f . This is probably due to a variety of factors such as: ( 1) insufficiently con
verged eigenfunctions used as initial values, (2 ) the increased size of the pertur
bation amplitude for the mode n = § due to the manner in which the eigenfunctions
were normalized, and (3) inaccuracies in the time-stepping. Whatever the cause,
1834
t=\T
the graph of С against t reveals a slightly noisy signal. The profiles for the two
modes are shown in figures 11 and 12.
The normal-mode theory predicts that ak = 0.40 is just on the neutrally stable
side of a very rapidly growing instability. We have tested the n = £ mode and have
observed growth but at a rate significantly slower than that of the unstable waves
on either side. This discrepancy is probably due to the factors mentioned previously
and to the close proximity of the rapid-growth region. The profiles are shown in
figure 13. One interesting feature of these is the near step-function shape of the
perturbation at t = 0 and t = \T = 17.07. This is also apparent in the n = J profiles
for ak = 0.38 at t = \T = 8.71 as shown in figure 11. The perturbation elevates
one wave trough and lowers the next. This also happens but to a lesser extent for
1835
0.05
f=0
0.05
t=\T
F ig u r e 10 . Development of the unstable mode ($, §)+ when ak = 0.25, over one half-cycle of
the perturbation. The vertical scale on the right is enlarged times 20.
-,0.05
t= 0
F ig u r e 1 1 . The neutrally stable mode n = £ when ak = 0.38 followed over one half-cycle of
the perturbation. The vertical scale on the right is enlarged times 20.
reduce one wave crest and raise the other with not much alteration of the profile
elsewhere. The crest-plots of figure 15 show the overturning. These computations
were done with 120 points along the profile for increased resolution. The comparison
of theoretical and observed growth-rates is given in figure 16, a plot of In R against
time. The spread of observed values at each data point represents the variability
of R for slightly different methods of calculating the integrals. This gives a rough
indication of the minimum size of the errors involved. The agreement with the
linear theory is quite good especially during the initial stages of growth. At later
times the deviations are not too great even when the perturbation has grown large
enough to cause marked asymmetries in the wave profiles.
1837
7. T he n o n l in e a r stages of grow th
all respects similar to that arising at ak = 0.41. Since wave breaking involves
energy loss, this leads to the disintegration of the original wave train.
To test whether the same phenomenon occurs with lower waves, a similar run
was carried out on the mode (J, §)+ at the lower wave amplitude ak = 0.25. Figure
19 shows the surface profile, after nine successive half-periods. In the final profile
the waves do indeed break (see also figure 20 ). A close-up of the breaking crest is
shown in figure 21 . Again, the close similarity to the local instability in figure 15
will be noted.
The calculated values of G(ct, t) and ft are shown in table 2 . The constancy of ft
throughout the whole stage of development is indeed remarkable, as also is the
closeness of ft to the value 0.0132 derived from the independent normal-mode
analysis.
1839
0.2
*=o
4 0 h
- 0.2
4
v-"
h -
h -
h - 4
h -
H -
H -
H -
2.05
I— ^
0 , 4*
F io u h e 14. Development of the highly unstable mode n = (|)+ when ak = 0.41.
The vertical scale on the right is magnified times 6 .
ak > 0 .4 3 6 Д /2 = 0.3 0.
1840
Figure 16. The mode n = (|)+ when ak = 0.41; enlargement of the wave crest, seen in a frame
of reference at rest. Time interval between profiles is 2 я /б 0 ; the range, 1.37 ^ t < 1.87.
Figube 16. Growth of the mode n = (Jj+whenoi = 0.41, shown by In R(t) ae a function of (.
1 he broken lrne represents the growth fit where fi is the rate corresponding to the normal-
mode analysis.
1841
t=UT
1 20
t= T
The example ak = 0.25 shows that this condition is contradicted. The reason
appears to be that every alternate wave is not only increased in amplitude, but also
shortened relative to its neighbours, so that locally a higher value of the wave steep
ness is attained.
This is confirmed by figure 22 . In figure 22 a is shown the variation with time of
the individual wave amplitude a', defined as half the height of each crest above the
trough in front of it. This is normalized by multiplying by the constant wavenumber
k. The amplitude of each wave fluctuates, and the fluctuations grow in time. Figure
226 shows the individual wavenumber k' corresponding to the horizontal distance
between adjacent troughs. This also fluctuates, by as much as 12 % in either direc
tion, though these fluctuations are almost in quadrature with the fluctuations in a'.
1842
F ig u r e1 8 . Enlargement of the wave crest in figure 2 0 , near the instant of overturning, seen
in a reference frame at rest. Time interval between profiles is 2 я / б 0 ; the range,
3 4 .6 6 t 3 5 .1 6 .
The profiles in figures 15, 18 and 21 are so similar as to suggest that the dynamics
of the final stages of overturning are determined mainly by local conditions near
the wave crest. If this is so then we would expect that the local length-scale I and
time-scale r would be related by
l/дт‘ = 0 ( 1),
where g denotes gravity. In fact if we take I to be the *width *of the crest, measured
by the horizontal distance over which the profile is convex, then I « 0.4 while the
time r for the evolution of the crest in figure 15 is above five plot intervals, i.e.
about 0 .6 . Since g = 1 in these units the above relation is indeed satisfied.
However it will be noted that the time-scale for the initial growth-rate of the
instability when ak = 0.41 is much longer than т since /?, though larger than for
the Benjamin-Feir instabilities, is still only equal to 0.065. Thus is about 15.5.
1843
*=0
,
0 .2
- 0.2
2T
f-----------------------------1
3T
AT
F i g u r e 19. Later development of the unstable mode ($, f ) + when ak = 0.25. The profiles
are seen at times t = \вТ, а = 0, 1, 2, 9 where T is the period of the mode.
F ig u r e 20. Profiles of the mode (ф, f)+ near breaking, when ak = 0.25.
The time interval between profiles is 2 л /4 ; t = 140.86 and 142.42.
1844
T a b l e 2 . O v e r a l l g r o w t h -r a t e o f t h e
i n s t a b i l i t y (£, f ) + w h e n ak = 0.25
t C(ct, t) P' = In R
= 1 4 .6 1 -0 .9 9 2 0 .0 1 3 5
T 0 .9 7 9 0 .0 1 3 4
Iт -0 .9 8 2 0 .0 1 3 1
2T 0 .9 6 6 0 .0 1 3 4
fT -0 .9 6 3 0 .0 1 3 3
ВТ 0 .9 4 6 0 .0 1 3 3
IT -0 .9 1 6 0 .0 1 3 6
4T 0 .8 6 3 0 .0 1 3 6
!т -0 .7 6 6 0 .0 1 4 1
8. C o n c l u s i o n s
By applying the independent time-stepping method of paper I we have confirmed
and extended the conclusions of recent studies (Longuet-Higgins 19786 , c) relating
to the stability of gravity waves. The initial rate of growth of each type of instability
has been verified, and it is confirmed that subharmonic instabilities of Benjamin-
Feir type are indeed confined to a certain range of wave steepnesses. Beyond this
range, the perturbations are neutrally stable, until at about ak = 0.41 a second
type of instability is found which leads directly to wave breaking.
1845
Figure 22. Time-dependence of local properties of the surface: (a) the local wave amplitude о ;
(6) the local wavenumber k'\ (c) the local wave steepness a'k'.
1846
R eferences
B y M . S. L o n g u e t - H i g g i n s , F.R.S.
1. In tro d u ctio n
Over 130 years after the formulation by Stokes (1845) of the general equations for
inviscid irrotational flow, it is remarkable that few if any exact solutions to the
problem of time-dependent free-surface flows under gravity are known. Most known
solutions refer to flows that are steady, or, like progressive waves, are easily re
ducible to steady flows by an appropriate choice of the frame of reference. Among
the time-dependent gravitational flows that have been considered are those for
shallow-water waves and standing waves, but these all appear to involve expansions
in powers of some small parameter, such as surface slope, and hence must be con
sidered as approximations at best.
Other time-dependent solutions, such as the Dirichlet parabola (John 1953),
ellipse (Taylor i960) and hyperbola (Longuet-Higgins 1972) are exact but do not
contain gravity in an essential way.
An interesting method for obtaining solutions to time-dependent free-surface
flows was indeed suggested by John (1953), who in this way derived some special
flows (see also Longuet-Higgins 1976). However, the method cannot be general,
for it assumes a Lagrangian coordinate a) which is an analytic function of the velocity
potential. This would not include, for instance, a progressive irrotational wave
with plane bottom, in which the mass-transport velocity induces a strong vertical
gradient of the mean displacement.
The present situation is highlighted by the absence of any satisfactory analytic
solution to the problem of an overturning wave, although the possibility of accurate
numerical calculation has been demonstrated by Longuet-Higgins & Cokelet (1976,
1978 ).
In our quest for a general method, we may be guided by the following considera-
2. Z AND t AS IN DE P E N D E N T V ARIABLES
z — x + iy, z* = x - i y , ( 2 . 1)
which is an analytic function of z and a smooth function of the time t, such that the
particle velocity и + iv = W* is given by
Here a suffix denotes partial differentiation, and an asterisk denotes the complex
conjugate. The pressure p is then given by the Bernoulli equation in the form
|2S>
Now if F(z,z*) is any differentiable function of both z and z* we have
+x,*(xzt+xa x * - g )
+ Х'<.Хг* + Х * * Х .-9 ) (2-10)
and so
D p/D t = R e [jff» -C ¥ K+ 2 ^ *xrt+ W * x J + f t), (2.11)
whprp
W* = Xz* = T>zfDt. ( 2. 12)
The remarkably compact expression (2 . 11) seems not to have been recognized
previously.
For steady flows or for motions such as progressive waves when referred to the
appropriate coordinates, equation (2 .11) reduces to
The first term on the right represents the pressure change due to vertical motion
in the hydrostatic pressure field. The second term, involving Xa* represents the
effect of the curvature of the streamlines.
16-2
1850
444 M. S. Longuet-Higgins
3. X AND * AS I N D E P E N D E N T v a r ia b l e s
For some problems (for example in waves of finite amplitude) it may be more
convenient to take x and t as independent variables, and to express the space co
ordinate z in terms of them. For steady waves (no time dependence) this method
was suggested by Stokes ( 1880 ).
To find the corresponding expressions in the new variables, we have for example,
by the method of differentials,
dz = zx d^ + ztt df,
(3.1)
= X.dz + Xtcfc.J
On eliminating from these equations and then equating coefficients of the inde-
pendent increments dz and dt we obtain
■ W b -M (3.2)
z x Xt + 2t = 0.J
Hence
Xz= l / z x = W (3.3)
and
* = - Z t/ Z x = - Wzt. (3.4)
To obtain the corresponding expression for Dp/TH, we note that if F(z, t) is any
differentiable function which on substitution z — z(x) becomes equal to G(x, 0
then, by the same argument,
F. = Ox/z x (3.6)
and
I ',= Ot+ X tGx, (3.7)
Ft = Gt - G xzt/z x. (3.8)
xa = - W \ x,
x w --T P V * V «r (39)
* „ = - W z tt- 2 W % z xt~W>z\zxx..
H = J>x/Vt, (3.13)
showing the analogy between (2 . 11) and (3.10). However, since W ( = Xz — I /2*) i®
linear in H is clearly not linear in z, showing that (2 . 11) is generally the simpler
expression of the two.
In steady flows equation (3.10) reduces to
R e(W<fzxx), (3.15)
where q denotes the particle speed. In progressive gravity waves this term must
exactly balance the hydrostatic term at the free surface itself.
4. X A N D 2 B O T H F U N C T I O N S OF 0)
The analysis of the two previous sections may be included in the general case
when x and z are each functions of a third complex variable w, and of the time
that is
X = X(a),t), z = Z(w,*). (4.1)
Thus §§2 and 3 correspond to the cases Z = ш and X = ш respectively. The form
(4.1) also includes as a special case the formalism of John (1953) but is more general,
since John assumed that to was a constant following a particle, and moreover that
(o was real at the free surface. This restricted the solution of flows of a special class.
When we adopt the assumption (4.1), the equation of the free surface is given by a
more general (but real) function of <0and (o*.
Expressions forp and D p/D t can be derived by an argument similar to that of
§ 3. Thus we easily find
X. = X J Z „= W (< d ,t) (4.2)
say, and
Xt = Xt - W Z t (4.3)
leading to
- 2 p = iW W * + (Xt - W Z t) - g Z - f + c . c „ (4.4)
where c.c. stands for complex conjugate. The right hand side of equation (4.4) is
a real symmetric function of w, ы' and t.
Now if jP(z,z*, t) is any real symmetric function of z and z* which by the sub
stitution z = Z((o, t), z* = Z*(o)*, t) becomes 0(o), w*, t) it can easily be shown that
446 M. S. Longuet-Higgins
and
Ft = Qt - Q „ Z t/Z ',—Qu*Z t* /Z j* . (4.6)
Therefore from (2.8) we have
DG/Di = D F /O t = F ,+ W *F ,+ W F *
= a t + ( W * - Z t) Q J Z „ + ( W - Z t* ) G * I Z * . (4.7)
Applying this result to the function —2p of (4.4) we obtain finally
W = X J Z M= Dz/D< (4.9)
and
К = ( W * - Z ,) /Z U= Dw/Di. (4.10)
It can be seen that when Z з to we have
Sw= l , Z t = 0 ; Zww= Zwt = ^tt = 0 , (4.11)
so
= (4-12)
and equation (4.8) reduces to (2 . 11). On the other hand when I = w w e find
W = 1/Zmi К = W ( W * - Z t) = Я , (4.13)
But we shall see that the description of many flows is in the end simpler if we do
not make this assumption.
In order that the solution shall represent a permissible flow having a velocity
field that is both finite and single-valued everywhere within the fluid, we must have
ZJX„ = 1/Ж Ф 0 (4.16)
everywhere within the fluid. This implies, for example, that any zero of Zu within
the relevant domain must coincide with a corresponding zero of Хы.
6. S o m e exact s o l u t io n s
We mention now some known exact (but non-trivial) solutions to the equations
of §4.
First is the Dirichlet parabola (John 1953 ; Longuet-Higgins 1976) in which
Z^w-iR/fi, X=№/t, (6.1)
1853
R being an arbitrary real constant. For, from equations (4.9) and (4.10) we have
provided / = \R2/t*. From (6.3) and (5.4) it is clear that both p and Dp/T)t vanish
on the free surface
(yt*)2 = R(xP + -±*R) (5.5)
(= (6.7)
This was discussed by Longuet-Higgins ( 1972, 1976 ). Both (5.5) and (5.6) have
surfaces which are conics, with axes in fixed directions in space. However, some
asymmetric and more general (irrotational) flows in which the free surface rotates
about 0 with a non-zero angular velocity will be described in another paper (Longuet-
Higgins 1980 a).
The above-mentioned solutions are all gravity-free, and so their application is
restricted to situations where the fluid is in free fall, or where the motion develops
so rapidly that gravity can be neglected. Of great interest therefore is the Stokes
comer flow, which in our formulation corresponds to
The free surface is a pair of lines (or planes) inclined at angles + to the vertical.
To see that this is a solution, we have from equations (4.9) and (4.10)
448 M. S. Longuet-Higgins
and so from (4.4)
p = g((i)2 —ww*+ a)*2). (6.11)
I* (5.13)
Dt 4i o)(o*
6. S i n g u l a r it ie s in the pressure f ie l d
The free surface, being a surface of constant pressure, can have a sharp comer
only when the pressure gradient vanishes. Consider then the condition that the
pressure gradient vanish while the pressure p remains a differentiable function of
x and y, or equivalently of z and z*.
From equation (2.9) and the general formulae (2 .6 ) for differentiation of a smooth
function we have
-2z>x = (xa X * + Хгг*Х,) + (Xzt + Xrt*)-2 ?. ! (6
Re(X=X.* + * « t-0 ) = O, I
(6.2)
1855
We shall be concerned mainly with saddle points that is to say points at which two
different contours p = constant intersect one another. In that case the angle у
between the two contours is given by
cos* у = -— -. (6 .6 )
(P z x -P ,v ) + *Р ху2
Now from (2 .6 ) we have in general
P xx Ргг "1" 2P a *
P xy = i ( P * z - P z ' A (6.7)
P vv = ~ P z z + 2Pzz* - Pz*z»>.
so that
P xx+P yv = *Pzz*>
( 6. 8 )
P x x ~ P w ~ ^(Pzz~b~Pz*t*) . }
~ 2Pzz -= д и + д и я Л 1 ( 6 1 1 )
- 2p = XzzXzz*' '
“ ' - B S f e r (6,12)
1856
450 M. S. Longuet-Higgins
The expressions for the mean curvature and the Gaussian curvature of the
surface p = p{xyy) are also very simple. In fact from (6 .8 ) we have that
м = p „ + p yv = (6.13)
showing that the mean curvature of p is in general negative. Exceptionally where the
velocity gradient vanishes, that is when
7. D i s c u s s i o n
We have seen that the equations for time-dependent irrotational flow of a perfect
fluid with a free surface become remarkably compact when formulated in terms
of the complex variables z - x + iy and z* — x —iy, instead of x and у respectively.
This is because the velocity potential ^ is a function of z and t only, independent
of 2*, and because the pressure p is a real function of z, z* and t which is symmetric
in z and 2*. The kinematic boundary condition T>p/Dt = 0 is also much simpler
than expected.
This simplicity is maintained when z and x are ©ach expressed in terms of a third
complex variable w, which need not be a Lagrangian coordinates, as in the formu
lation of the problem by F. John.
In all the exact solutions known at present it appears that p and T>p/T>t are
related by an equation of the form
where Ф is a real function, symmetric in (o and o>*. In the Dirichlet conics, which
are gravity-free, Ф is independent of ш and w*, and is a function of t only. In the
Stokes corner flow, on the other hand, Ф is a function of (o and ш* only, and inde
pendent of t, while in the self-similar solutions described by Longuet-Higgins ( 1976 )
X, 2 and Ф all involve functions of a similarity variable a>tx, where Лis a constant.
1857
However, equation (7.1) is not the most general relation which ensures the vanish
ing of Djj/D* when p = 0 . For example, we may also have
Цр/D* = Ф1р-\-Ф2рг + . . . . (7.2)
Relations of this kind will be explored in the papers to follow.
R eferences
John, F. 1953 Two-dimensional potential flows with a free boundary. Commune pure appl.
Math. 6 , 497-503.
Lamb, H. 1932 Hydrodynamics, 6th edn. Cambridge University Press.
Longuet-Higgins, M. S. 1972 A class of exact, time-dependent, free-surface flows. J. Fluid
Mech. 55, 529-543.
Longuet-Higgins, M. S. 1976 Self-similar, time-dependent flows with a free surface. J. Fluid
Mech. 73, 603-620.
Longuet-Higgins, M. S. 1980 a On the forming of sharp comers at a free surface. Proc. R. Soc.
Lond. A 371, 453-478.
Longuet-Higgins, M. S. 1980 b A branch-point model of a breaking wave (in preparation).
Longuet-Higgins, M. S. & Cokelet, E. D. 1976 The deformation of steep surface waves on
water. I. A numerical method of computation. Proc. R. Soc. Lond. A 350, 1-26.
Longuet-Higgins, M. S. & Cokelet, E. D. 1978 The deformation of steep surface waves on
water. II. Growth of normal-mode instabilities. Proc. R. Soc. Lond. A 364, 1-28.
Stokes, G. G. 1845 On the theories of the internal friction of fluids in motion, and of the equili
brium and motion of elastic solids. Trans. Camb. phil. Soc. 8 , 287-341 (reprinted in
Mathematical and physical papers, vol. 1, pp. 75-129. Cambridge University Press, 1880).
Stokes, Sir G . G. 1880 Supplement to a paper on the theory of oscillatory waves. Mathematical
and physical papers, vol. 1, pp. 314-326. Cambridge University Press.
Taylor, G. T. i 960 Formation of thin flat sheets of water. Proc. R. Soc. Lond. A 259, 1-17.
1858
By M. S. L onguet -H ig g in s , F.R.S.
Department of Applied Mathematics and Theoretical Physics,
Silver Street, Cambridge and Institute of Oceanographic Sciences,
Wormley, Surrey
1. I n t r o d u c t io n
454 M. S. Longuet-Higgins
overturning of the whole mass of fluid such as occurs in a plunging breaker. This
will be the subject of a subsequent paper (Longuet-Higgins 19806).
That these two aspects are in fact distinct is shown, first, by the observation that
upwards-pointing cusps sometimes occur, and secondly by the demonstration both
numerically (Longuet-Higgins & Cokelet 1976) and by laboratory experiment
(Longuet-Higgins 1977) that in plunging breakers it is possible for the upper part
of a breaking wave to overturn entirely before a sharp corner is actually formed.
Among previous authors who have attempted an analysis of the local flow, Biesel
(1952) gave an approximation which is valid only for small surface slopes, and
which cannot in fact be carried to the limit when a cusp is formed. Beyond this
limit, his solution is topographically impossible. The ‘ spiral vortices *of Tulin (1964)
again are multi-valued flows which cannot be incorporated into a rational
solution without the invocation of ‘ turbulence’. Such solutions are perhaps
more appropriate to the state of the fluid after a wave has broken, rather than
before.
N ot long ago some attention was attracted by the suggestion of Zeeman (1969)
that the formation of a sharp cusp corresponds topologically to one of the seven
‘ elementary catastrophies’ enumerated by Thom (1969). However, Thom pre
supposes that the free surface is defined by the condition that a certain ‘ potential
function’ shall be stationary on the surface. Despite unpublished attempts by
Professor Zeeman and by the present author, no appropriate ‘ potential’ for the
gravity-wave problem has yet been discovered. A t the present time, therefore,
Zeeman’s suggestion lacks any hydrodynamical content. (In a later discussion
Zeeman (1975) has expressed reservations about the validity of this model.)
The method of the present paper is relatively simple. As in paper I we seek a
surface on which the pressure p and its rate of change T>p/T>t following a particle
will both vanish together. In this way both the dynamical and the kinematic
conditions on the free surface will be satisfied. In the neighbourhood of some point
where the free surface has a sharp (though finite) curvature, we can expect there
to be a saddle point in the pressure field, not generally on but outside the fluid.
Such a point has downwards acceleration g, so that to an observer attached to this
point the fluid appears to be in free fall, and the equations are greatly simplified.
In fact this procedure effects for free-surface flows a simplification quite similar
to that achieved for progressive waves by taking axes moving with the phase speed
(see § 2).
In § 3 we therefore seek solutions appropriate to the neighbourhood of a saddle
point of p. We obtain a family of solutions which includes the Dirichlet hyperbolae
(Longuet-Higgins 1972) but is more general. For, whereas the solutions found
earlier had surfaces in the form of conics with principal axes fixed in direction, the
new solutions allow the principal axes to rotate in space; this despite the fact that
the flow itself is irrotational (see § 4 ).
The time history of these flows is described in § 5 and it is shown that the vertex
angle (i.e. the angle between the asymptotes) of the hyperbola can tend to zero
1860
as the time t increases. Moreover, although zero angle is not reached in finite time,
nevertheless arbitrarily small angles may be attained.
Some very interesting special cases are described in §§6 and 7. In one of these
the fluid has the form of a simple rotating slab, whose thickness tends to zero as
t -+ oo. In the other, the vertex angle of the hyperbola tends exponentially to the
value 45°, while the axes of the hyperbola rotate at a constant (but non-zero)
angular velocity.
The formation of a sharp cusp requires not only the vanishing of the vertex angle
in the surface as described by the quadratic terms, but also the presence of some
terms cubic in the space coordinates. In § 10 we show that the exact solutions of
the previous sections may be modified by the addition of cubic and higher-order
terms, which are determined ultimately by the conditions for matching to the outer
flow (this involves the gravitational acceleration <7). Figures 11 and 12 show that
some of the resulting forms may be quite realistic. The argument also suggests
that upwards-pointing cusps are more easily formed than those that point down
wards.
2. F rame of references
456 M. S. Longuet-Higgins
the original frame the flow at large distances is steady or horizontal, then in the
new reference frame the velocity potential must contain a term - gtz which must
be matched to the local flow in some intermediate zone.
In this paper we shall be concerned almost exclusively with the local flow. Since
the flow is regular at the origin О it follows that x т а У be expanded in a power
series:
X = Lz + iAz* + iBz* + ... (2 . 1)
in which the term independent of z, which has no effect on the velocity, is taken to
be zero. The remaining coefficients are complex functions of the time t.
Since the origin moves with the fluid, Xt must vanish at 0. Hence
L = 0 (2 .2 )
and we may verify that the condition for a stationary value of p in a free-fall
reference frame, namely
+ = 0 <2-3)
(see paper I, equation (6.4)) is indeed satisfied.
General expressions for the pressure p and its rate of change q following a particle
have been shown in Paper I to be
where F is some real function of the time t and a dot denotes its rate of change
(and c.c. is the complex conjugate). The formal problem is to find solutions x(z>0
such that p and q vanish together on the same (moving) surface.
In the first part of this paper we shall study solutions in which A is the only
non-vanishing coefficient in equation (2 . 1). Later we shall investigate the possibility
that В also does not vanish.
3. Q u a d ra tic s o lu tio n s
Suppose first, then, that
(3-1)
the remaining terms in equation ( 2 . 1) being zero. On substituting for x *n equations
(2.4) we obtain
- 2p = \{Az2 + 2 AA*zz* + A*z* 2) - 2 F, (3.2)
and — 2q = £[(Л + 2A2A *)z2 + ±A A * zz* + c.c.] —2fi. (3.3)
A sufficient condition for equations (3.2) and (3.3) both to vanish on the same
surface is that the coefficients of corresponding terms are proportional, that is
A + 2 A 2A * 2 {AA* + AA* ) A* + 2A A *2 /
(3.4)
AA* A* F’
We shall now obtain solutions to these equations which are a generalization of
those obtained in Longuet-Higgins ( 1972 ), where it was assumed that A was real.
Now let us write .
A = a e l'r, (3.5)
where a and cr are both real. Then
A = (a + iatr) e|/r (3.6)
and A = [(a —i<r2a)-H ( 2dd--|-a0-)]eUr. (3.7)
So on substituting in equations (3.4) we obtain
(a —ol&2+ 2as) + i(2a<7 + acr) 4a a F .
. ' ■ . — a — C.C. — t j , (3. о J
a + ia<r a2 F
Because each ratio equals its conjugate complex expression we find, by adding or
subtracting the corresponding numerators and denominators,
a -a & 2+ 2a3 2(а& + а&) 4a
(3.9)
a eta cl F’
From the second and third expressions we get
1 + 1 - 4 (3-10)
oc, cr a
! - « * (3. 11)
\
1863
458 M. S. Longuet-Higgins
and on integration with respect to t,
& as Лее2, (3.12)
where Л is an arbitrary real constant.
In the first expression in (3.9) we may now substitute A2a 4 for <72, and then from
the first and third terms we get
ad —4a2 + 2a4 —A2ae = 0. (3.13)
This is a generalization of equation (3.6) of Longuet-Higgins ( 1972), to which (3.13)
reduces when A = 0 . Now multiplying both sides of (3.13) by the integrating factor
2d/а 9 as before we obtain, on integration,
- 8, Aa 4A a 2- r .
a <з “ >
where P is a constant. Hence
d2 = a 4 - A 2ae+ Pa8, (3.16)
a generalization of equation (3.8) of Longuet-Higgins ( 1972).
We shall see that the essential feature of this generalization is that it allows the
existence of a class of flows that is asymmetric, whereas all the flows treated pre
viously were symmetric about the axes of x and y.
The general solution of equation (3.15) above is simply
Aa»d<, (3.17)
4. Form o f th e f r e e su rfa ce
The equation of the free surface is p = 0 , or from equation (3.2)
A z* + 2 A A * zz * + A * z *2 = 4F. (4.1)
Introducing polar coordinates
z = re* (4.2)
and substituting the expressions (3.5), (3.6), (3.12) and (3.19) we have
(d + iAa8) r2 e1*2* ^ + c.c. + 2aV 2 = - 2Sa*. (4.3)
Now writing a + iAa3 = —/?eIe, (4.4)
where cose = —a/fi, sine = —Ла3/Д, (4 б )
and ft1 = а 2 + Л2ав = а4(1 + Р а 4), (4.6)
by (3.15), we see that (4.3) becomes
pr2el(2fl‘Hr+e) —aV 2+ c.c. = 25a4. (4.7)
Referred to cartesian coordinates
x' = rcos [в+\{<т + e)],|
(4.8)
y' = rsin [0 + i(cr + e)],J
460 M. S. Longuet-Higgins
We notice also that in the general case
S2a* S2
a‘b* = j ^ = p (4Лб>
a constant. The angle у between the asymptotes is also given by
b2- a 2 a2 1
cosr= - ^ = 7 = (7Т Щ * (4Л7)
It will be seen from (4.16) that in general a2 > b2, so the angle which each asymptote
makes with the axis of x' is less in magnitude than 45°, and у is less than 90°.
F i g u r e 2. Sketch of the hyperbolic form of the free surface as given by equation (4.14).
у denotes the angle between the asymptotes. The axes of x' and y' rotate about the
origin О with (clockwise) rate of rotation 6.
The rate of rotation of the axes can be found from the identity
On substituting from equations (4.6) and using (4.6) with the differential equation
for a (equation. (3.16)) we find
e = Aaa ? ~ ^ * 1 . (4.19)
1 + Pa*
Together with (3.12) we have then
t i / • -4 (4.20)
* = *(tr+e) = I T P ^
1866
From equation (3.15) we see that for a solution to exist we must have
1 —A2a 2 + Pa 4 ^ 0 . (5.1)
For sufficiently small values of a (excluding а = 0 ) it is clear that a solution will
always exist. In fact from equation (3.16) we have, for small values of |a|,
(5.2)
J a2 a
(we shall agree for the present to take the negative square root in (3.16) so that ±
signs can be omitted). So for small a we have
a ~ l/t (a 0) (5.3)
and we see also from equations (4.18) and (4.21) that in this limit the angle between
the asymptotes tends to zero and the rate of rotation of the axes tends to zero also.
At the opposite extreme when |a| -> oo solutions also exist, provided P Ф 0 . In
the limit а -* oo we see from (3.16) that
The parameter A is evidently a time scale for the motion. In (5.7) we have fixed the
lower limit of integration at / = 0 , but any other choice simply changes the origin
of the time t, that is to say it shifts each curve in figure 3 horizontally by a fixed
constant.
\
1867
462 M. S. Longuet-Higgins
If on the other hand 0 < P < £A4, equation (5.6) has four real roots, a = alt ± a2,
g i v e n b y 1 /A * a f = / ! = i + ( i - w ) * . l
(5.9)
1/А*<Х| = / I = * - ( i - T77)*J
Separate solutions exist in the ranges 0 < a < ax and a 2 < a < oo, and not in the
intermediate range ax < a < a 2 (see figure 4). In the first interval, the typical
solution may be taken as ^ - 1/Ae ^
(5.10)
A = Л ( ^ - / ! + / 4)*'
The function a = 1/А / is small at large negative values of t, when the angle у
between the asymptotes is also small. Then a increases to its maximum value ax
at / = 0 and diminishes again towards zero as t -> oo.
1868
In the second interval (a2 < a < oo) the typical solution is
H fd f (6 . 11)
The solutions start at t = —t' with a and/ -1 large and у approximately 90°. Then
a and у diminish to a minimum value when a = a2, afterwards increasing again to
infinity 8L8t-+t'.
The value of t' is finite, and some values are given in table 1 for certain values of
ш, along with the corresponding minimum angles ymin.
The rate of rotation of the axes, which is given by equation (4.20), is always of
1869
464 M. S. Longuet-Higgins
constant sign, positive if A > 0 . However, it will be seen that |<j| takes its maximum
value when Pa 4 passes through the value 1, making
д = A / 2P* = 1 /2 mX. (6.12)
This can take place in one or other of the two ranges of a.
T able 1
6. T he s p e c ia l case P = 0
In this and the following section we shall discuss two particularly interesting
special cases which have been excluded from the previous discussion.
The first of these is when P = 0 . Then a2 goes to infinity, = 1/A, and the
solution must lie in the range 0 < a < exclusively. In fact from (3.16) we have
da (1 —A’ a*)*
■ u ■V -A W j*-------- a----- ‘ ( >
Hence a = l / ( < s+A*)t (6 .2 )
Since p = (a4+ P a 8)* = a*, the free surface, given by (4.10), consists simply of
the two paraUel planes x- = ± S a (6.5)
466 M. S. Longuet-Higgins
7. T h e case P - JAa
The second case of special interest is when the two roots ax and a* of § 5 are
coincident. Then we must have ax = a2 = J 2 / A, and from (3.16)
f da
(7.1)
*“ ”Ja2(l-iA2a2
*)
which can be integrated immediately to give
Aa + 7 2
1 - i - J_i (7.2)
А Ла + 2^2 Aa-^/2
Solutions may exist in either of the ranges 0 < a < 2^/A and 2/A < a < oo. The
limiting value a = ^/2/A may be approached either from above or from below.
Interestingly, since
t ----- (ЛД/2)1п|А(ас —a,)|, (7.3)
weseethat |a “ a i| ~ e~Vitlx. (7.4)
In other words, the free surface approaches its limiting form exponentially in t,
that is much more rapidly than when а -► 0. The limiting angle between the
asymptotes is given (see equation (4.17)) by
a = constant, a = a = 0л
I (7.8)
& = constant, cr = 0, J
Substituting now into the equation of the free surface (4.3), we find this becomes
to n ^ = ( v § ? i ’:)*• (7 12 )
8. The a n g le s у and 8
468 M. S. Longuet-Higgins
y/deg
m
y/deg
m
FiaimE 7. The angle у between the asymptotes as a function of the time t
(a) when то > (Л) when m <
1874
The angle у is shown as a function of t/A in figure 7a, b corresponding to m > 0.25
and ш < 0.25 respectively.
In figure 7a the behaviour of у is very similar to that of a in figure 3, except that
the range of у is finite. The broken line corresponds to one of the limiting curves
ш = 0.25. It can be seen that when w is near to but not equal to 0.25 the angle у
tends to ‘ linger’ in the neighbourhood of 45°.
In figure 7 b (ш < J) the behaviour of у is similar to that of a in figure 4, except
that when m = 0 (the rotating ‘ slab’) the graph of у is the straight line у = 0.
Again the solutions fall into two classes. The first consists of those for which у
starts from 90° (at t = —*'), decreases to a minimum (greater than 45°) at t = 0,
and then increases again to at finite time t = t\ In the second class, the angle у
starts from near zero at t = —oo, increases to a maximum at t = 0 and then falls
towards zero again as t -+ -f oo.
In figure 7a we show one of the limiting curves for w = 0.25. However, the other
limiting curve, in which у starts from near 45° and decreases monotonically to
zero, cannot be shown in figure 7b, since it is not symmetrical about the instant
t = 0.
To evaluate the angle of orientation $ we may begin from equation (4.20), which
can be written in the form
(8.3)
(8.4)
which can be calculated from the expressions for у as a function of t/A which have
just been found. In certain cases, however, it is more convenient to evaluate S
independently in terms of £. For, from (3.15) and (8.1) we have
(8 .6 )
(8.7)
( 8 . 8)
Graphs of 8 are shown in figure 8a, b for the cases ro ^ J respectively. In figure 8a
it will be seen that when m > 0.25 the angle of orientation increases monotonically
1875
470 M. S. Longuet-Higgins
to a finite, non-zero value. The limiting values are indicated on the right of the
diagram, and also in table 1. In the special case ш = J, the angle 8 increases without
limit, as we would expect from equations (7.2) and (8.8).
In figure 86 we see the behaviour of 8 when w < J. In this case also 8 tends mono-
tonically to some positive value (the limits are given also in table 1). When m = 0,
as shown by the lower broken curve, the limiting value of 8 is 90°, that is the axes
have turned through a right angle from their position at t = 0. The upper broken
curve in figure 8b corresponds not to equation (8.8) but to the limiting solution of
§7, in which the rate of rotation is uniform, by equation (7.15). In fact we have
simply
8 — t/\ . (8.9)
1876
m\
F i g u r e 8. T h e angle 8 o f rotation o f the principal axes, as a function o f
the tim e t (a ) when m > J, (fr) when w < J.
9. I l l u s t r a t io n s
In figure 96 we show the case w = 5.0 in which the total angle of rotation 8 is
only 39°. W e have also started at time t — 0 with a different initial orientation.
Figure 9c illustrates the critical case w = J when the axes rotate indefinitely with
a uniform angular velocity.
Figure 10 shows two cases when w < J. In figure 10a the axes rotate through 14.5°,
at which instant the vertex angle is 52°, before returning again to a vertex angle of
90°. In figure 106 the angle between the asymptotes starts from 0, increases to
nearly 32° when t = 0, and then falls again to 0.
1877
472 M. S. Longuet-Higgins
In all the examples of figures 9 and 10 it was assumed that the origin 0 was
stationary, or at least progressed uniformly in a horizontal straight line. In general,
however, О is in a free-fall trajectory. Such a trajectory is described by the para-
metric equations x - x, = - <,)»,■>
У -V i = U V -h )’ ) 3>
where U denotes the horizontal velocity and {xv y j denote the coordinates of the
highest point of the trajectory, attained when t = tv By an appropriate choice of
xv y v tx and U we can construct some suggestive time histories (figure 11).
It will be noted that in the trajectory (9.3) we have
dу dy/dt _ U
(9.4)
da; dx/dt g(t - у ’
so that the inclination 6 of the trajectory to the vertical is given by
U
в = arctan (9.5)
g(t —f|)
1878
$ - 8 00 ~ - A / t . (9.7)
Comparing equations (9.5) and (9.7) (and bearing in mind that the hyperbola is
rotating clockwise) we see that the axes of the hyperbola will follow the parabolic
path of О rather well if we have $тлх = 180° and A = U /g (see figure 11 a). This
leads us to define the matching parameter
fi = A g/U . (9.8)
17 Vol. 37i. A
1879
474 M. S. Longuet-Higgins
s -v .
(a)
F ig u r e 11. Local solutions with the origin О in a free-fall trajectory, (а) ш = 0.3, /* = 1 and
= 180°. (b) w = 0.26, ц = 1 and = 204°; (с) та = 0.261, = 2.0 and = 263°.
ft denotes the ‘ matching parameter’ Ag / U .
It should be borne in mind that the tip of the wave is not generally at О but at
a distance a from О given by equation (9.2). But since from (4.17)
10. H ig h e r - o rder t e r m s : m a tc h in g
In this section we investigate the possibility of solutions which are of higher order,
leading to the formation of a cusp at the free surface.
From figure 1 it will be seen that the simplest way in which this can occur is
for the saddle point at О to coalesce with another stationary point of the pressure
field. This second stationary point O' must clearly be a minimum of p. In general
the pair of stationary points may be represented locally by a cubic expression in z.
Accordingly we now write
X = \Az2+ \ B z * (10.1)
and calculate p and Dp/DJ from equations (2.4) as before. To order r3 we now
obtain
— 2p = — 2F + %(Az2+ ^^4*zz* + c.c.) + J(^z3+ 3^*Bz2z* + c.c.)
and —2q = — 2& + Щ А + 2A2A * )z 2+ 4AA*zz* + c.c.]
+ $[(£ + 6A A * B ) z3+ 6[A 2B * + А * В + А * Ё ) z *z * + c.c.]. (10.3)
I f we assume that the four cubic terms in equation (10.2) are in the same proportion
as before to the corresponding terms in (10.3) then we obtain four further relations
for the two unknown functions \B\ and arg.fi. But our assumption is unnecessarily
strong. It is quite sufficient, in order that p and q shall both vanish on the same
surface, to suppose that , / # _ ( , / * ) ( ! + * ), (10.4)
where R denotes a real power series. Provided that we retain the original solution
for the coefficient A of the quadratic terms in x> we easily see that R has to be at
least of third order :
R = i(i21z» + 3i?2z*z* + c.c.) + ... • (Ю.5)
On substituting the above expressions for p , q and R into equations (10.4) and
equating coefficients of the cubic terms we have
The functions A and F being already known from § 3, there are four equations for
the six functions consisting of the real and imaginary parts of R v R 2 and B. The
coefficient В is therefore undetermined by local conditions; if В is given, equation
(10.6) serves simply to determine Ri and Д2.
The addition of higher-order terms to R in equation (10.5) does not alter the
situation, since at each stage new coefficients i?3, J?4, ... are introduced which are
sufficient to accommodate the higher coefficients C, D , ... in x-
Thus the higher coefficients in x arc available for satisfying other conditions, for
example of convergence, and of matching the local flow to a particular flow at large
z. Details of this procedure will be described in a future study.
Meanwhile a particularly simple case may be noted here, namely when
Б = е »^ + т )-2, (10.7)
17-3
1881
476 M. S. Longuet-Higgins
where v and r are constants (r > 0). On substituting for В in equation (10.2) and
defining x', y\ e and £ as in § 4 we obtain for the free surface, as far as the terms in r3,
y '2 = - S a * + lPa*x'*
since e -> 0 as t -> oo. It will be seen that when xf is 0(1) then y ' is 0(a\) and the
cubic terms in equation (10.11) dominate the right hand side. Hence the form of
the free surface tends towards a cusp (figure 12).
Moreover, we have seen that in the outer flow field the fluid, in the free-fall
reference frame, will have an upwards (i.e. negative) acceleration. This makes it
probable that in flows closely matched to the outer field the real part of В will be
negative, producing a cusp pointing upwards. On the other hand in an overturning
flow, for example the tip of a plunging breaker, the region in the tip is connected
only very remotely to the flow in the rest of the wave. In this case it does not appear
necessary for cubic terms to enter the expansion. The appropriate curvature of the
jet can be produced by quartic terms or terms of even higher order.
11. D iscussio n
which is so extremely simple that it seems surprising that it was not discovered
1882
l
F i g u r e 12. The class 1 solution rn = 6.0, with a cubic term included, showing the near
formation of a cusp. In equation (10.7), v = 220° and r = A = 1 . Also S = 1/64.
478 M. S. Longuet-Higgins
time, nevertheless the protruding jet becomes indefinitely thin and would in practice
be easily broken up into water droplets and spray by the action of surface tension
and air currents. In upwards-pointing waves the appearance of a cusp will be
enhanced by the presence of cubic and higher-order terms which will tend to domi
nate over the quadratic terms as t tends to infinity.
Solutions in class 2, on the other hand, represent waves which at first seem to
advance towards breaking, but then subside into their former state. They may
represent limiting forms of the surface in partly reflected waves.
Class 3 solutions may represent the tip of a hydraulic arch.
In classes 1 and 2 our choice of initial instant as being the time when the angle у
is equal to 90° is to some extent arbitrary, and is based on the fact that for angles
у greater than 90° the pressure at О becomes greater, not less than, the pressure
at the free surface. Before this instant, the solutions would hardly be applicable.
Finally, in this paper we have studied only the local form of the free surface
near a region of sharp curvature. It remains to study the flow in the rest of the fluid
(the outer region). In a sense our solutions play a role in time-dependent flows
similar to that played in steady progressive waves by the Stokes 120° comer flow,
or by the almost-highest waves of Longuet-Higgins & Fox (1977). To obtain a
complete picture of a breaking wave, for example, we must match our local solutions
to some appropriate flow in the outer zone, such as will be described in Longuet-
Higgins (19806).
R eferences
Biesel, F. 1952 Study of wave propagation in water of gradually varying depth. Gravity
Waves, U.S. National Bureau of Standards, Washington, Circular no. 621, pp. 243-263.
Greenhill, A. G. 1879 On the rotation of a liquid ellipsoid. Proc. Cam. phil. Soc. 3, 233-246.
Greenhill, A. G. 1880 On the general motion of a liquid ellipsoid under the gravitation of itR
own parts. Proc. Cam. phil. Soc. 4, 4-14.
Longuet-Higgins, M. S. 1972 A class of exact time-dependent, free-surface flows. J. Fluid
Mech. 55, 629-543.
Longuet-Higgins, M. S. 1977 On breaking waves. Proceedings of the symposium on waves on
water of variable depth, Canberra, July 1976. Lecture Notes in Physics no. 64, pp. 129-
150. Berlin: Springer.
Longuet-Higgins, M. S. 1980a A technique for time-dependent free-surface flows. Proc . R.
Soc. Lond. A 371, 441-451. (Paper I.)
Longuet-Higgins, M. S. 19806 A branch-point model of a breaking waves (in preparation).
(Paper III.)
Longuet-Higgins, M. S. & Cokelet, E. D. 1976 The deformation of steep surface waves on
water. I. A numerical method of computation. Proc. R. Soc. Lond. A 350, 1-26.
Longuet-Higgins, M. S. & Cokelet, E. D. 1978 The deformation of steep surface waves on
water. II. Growth of normal-mode instabilities. Proc. R. Soc. Lond. A 364, 1-28.
Longuet-Higgins, M. S. & Fox, M. J. H. 1977 Theory of the almost-highest wave. J . Fluid
Mech. 80, 721-741.
Thom, R. 1969 Stability structurelle et morphogenbse. New York: Benjamin.
Tulin, M. P. 1964 Supercavitating flows - small perturbation theory. J . Ship Res. January,
pp. 16-37.
Zeeman, E. C. 1969 Breaking of waves. Proceedings of the Symposium on Differential Equations
and Dynamical Systems, University of Warwick, 1968, 3. Berlin: Springer.
Zeeman, E. C. 1975 Catastrophe theory: its present state and future perspectives. Warwick
dynamical systems (Springer Lecture Notes), pp. 1-34. Berlin: Springer.
1884
Lake & Yuen (1978) have suggested that in very steep wind waves the modulation -
frequency of the wave amplitude may correspond to the frequency of the fastest-
growing subharmonic instability of a uniform train of waves whose amplitude equals
the mean wave amplitude d. The approximate theory of Benjamin & Feir (1967)
gives this frequency as (ak)fd, where к is the wavenumber and f d the frequency of the
unperturbed waves. This expression applies strictly only to very small values of the
wave steepness ak.
More recently (Longuet-Higgins 1978) the present author calculated accurately
all the normal-mode instabilities of steep gravity waves on deep water. In this note
these calculations are used to determine the frequency of the fastest-growing sub-
harmonic instabilities precisely. When compared with the experimental data of Lake
& Huen, these frequencies show even closer agreement.
1. Introduction
In the study of random seas, questions concerning the fluctuation in the height of
the waves - for example, how many waves, on the average, are there in a wave group ?-
have important applications to naval architecture and ocean engineering, especially
when we are concerned with nonlinear phenomena (such as ship slamming) stimulated
by a resonant response to the waves.
In the past such questions have been treated mainly by linear theories (Longuet-
Higgins 1956; Goda 1970; Ewing 1973). Thus, there is a demonstratable relation
between the group-length of a gaussian stochastic process and the ‘width’ of the
corresponding frequency spectrum, suitably defined. Such linear theories do not
predict any particular value for the width of the spectrum, but leave this as a matter
either for measurement or to be estimated by wave-forecasting techniques.
In very steep waves, however, the waves become non-gaussian, and the frequency
spectrum, as ordinarily defined, becomes contaminated by the presence of second
and higher harmonics which are phase-bound to the corresponding fundamental
Fourier components. So the appropriate width of the spectrum tends to be over
estimated, when spectral moments are used. Against this, it has appeared that the
width of the main peak in the spectrum, under conditions of active wave generation,
can be extremely narrow (Hasselmann el al. 1973).
Now it is known (Benjamin & Feir 1967) that a uniform train of gravity waves of
finite amplitude a in deep water is inherently unstable to certain subharmonic per
turbations. Benjamin & Feir represented these as side-bands, with radian frequencies
o' ± A<r, where cr is the frequency of the fundamental and До* a positive perturbation;
0022-1120/80/4513-1320 $02.00 <g> 1980 Cambridge University Preaa
23 flm 99
1885
706 M . S. Longuet-Higgins
and they showed that the side-bands would grow at the expense of the fundamental
wave provided
Acr/cr < л/2 ak, (1.1)
к being the fundamental wavenumber. Moreover, the most rapidly growing perturba
tion (of this type) has a frequency such that
До-/сг = ak. (1.2)
This interesting result was taken up by Lake & Yuen (19 7 8 ) who suggested that the
fluctuations of the wave envelope in a steep, irregular wave train might correspond to
the most rapidly growing perturbations of a uniform wave train. In support of this
not unreasonable suggestion they measured the ratio of the modulation frequency
f e to the dominant wave frequencyf dof wind-waves in a short channel, with fetches up
to 30 ft and wind speeds up to 30 ft s_1. Despite some scatter, f j f dappeared to increase
about linearly with mean wave steepness (see Lake & Yuen (1978) figure 13; also see
figure 6 below).
Now the theoretical results (1.1) and (1.2) are valid strictly only for small but finite
values of A cr and ak. Recently the present author has calculated accurately the in
stabilities of deep-water gravity waves over the greater part of the range of wave
steepness ak (Longuet-Higgins 1978). These calculations have been confirmed by an
entirely different method (Longuet-Higgins & Cokelet 1978). The results show that the
instabilities are of two distinct types: local instabilities, which lead directly to
whitecapping, and subharmonic instabilities of the Benjamin-Feir type, which
however are confined to the range 0 < ak < 0-36 approximately. It was shown that
the relations (1.1) and (1.2) are valid only at the lower end of the range of ak.
In this note it will be shown that the results of Longuet-Higgins (1978) allow the
frequency of the most rapidly growing subharmonic perturbation to be determined
accurately. When the resulting curve of f e/fd is substituted for the straight line (1.2)
used by Lake & Yuen, rather better agreement with their data is obtained.
rje = |^4(f) | of the envelope function, drawn so as to touch the dominant waves near
to their crests. The envelope is in fact the smoothest such function (in a certain sense)
which can be so drawn. We wish to find the mean frequency (in some sense) of the
fluctuations of 7fe(t).
Now in the proposed model we regard 7}{t) as the result of modulating a uniform
train of waves of finite amplitude, that is an unperturbed wave
with af real. Here x and t represent the horizontal co-ordinate and the time. To retrieve
rj(t) we set x = 0.
It is useful to consider the perturbations of 7} in a frame of reference moving with the
phase-speed с = <r/к. Then 7} itself becomes independent of the time:
} =* F ty ) = S Л,сов( jty/c).
7 (2.3)
= ( 2 -4 )
ч ~ ч (Ф )+ Ш * )> (2-6)
where т)' is a small perturbation of the basic wave rj. Squares and higher powers of
tj' (but not rj) are neglected. It is relatively straightforward to calculate the TLormal
mode perturbations, which take the form
= (2.6)
where Pn and Qn are real and imaginary parts of the complex amplitude function: n is
a rational number characteristic of the particular normal mode; and
708 M . S. Longuet-Higgins
ak
F i g u r e 2. Real part of the radian frequency <r of normal-mode perturbations of deep-water
gravity waves, as a function of the wave amplitude ak (after Longuet-Higgins 1978).
Im ( a )
M . S. Longuet Higgins
ak
F ig u r e 5. The frequency of modulation of the wave envelope in a wave of initial
amplitude a perturbed by the fastest-growing instability.
1891
M . S. Longuet-Higgins
F i g u r e 6. Ratio of the modulation frequency to the dominant wave frequency, accordng to the
laboratory wind-wave data of Lake and Yuen (1978), compared to the most unstable
modulation frequency. Broken line represents Benjamin & Feir’s (1967) theory. Solid curve
represents present calculations. О» Uw = 15fts-1; □ , Uw = 2 0 ft а-1; Д, Uw = 26fts-1;
0 , Uw = ЗОЛв-1; V , Uw = 36 ft s-1.
field rather than say the root-mean-square wave amplitude, which would be somewhat
greater.
Nevertheless one can state, on the evidence of figure 6, that in the steepest waves
encountered in these experiments the frequency of modulation was about 0*2 times
the wave frequency, implying that in a very steep sea every fifth wave, on average, is
the highest. And, further, that this is supported by our calculations.
It must be emphasized that this would be the wave grouping as seen by an observer
or wave recorder at the fixed point x = 0, measuring the elevation as a function of
the time. When seen in space, the wavenumber of the envelope would appear to be
simply
Д п = | ( п 1- п 2). (4 .1 )
For waves of low amplitude, and fairly long groups we have in deep water
ал = Дo 'Ф \c Дn, (4.2)
1892
since the group velocity equals half the phase velocity. Hence the wave groups appear
twice as long in time as they do in space.
The lower the waves, according to these ideas, the longer the groups of waves become,
until the width of the spectrum ceases to be governed by the local nonlinear dynamics
and is determined instead by other factors depending on the synoptic wind pattern
and the path of wave propagation.
For field measurements of ocean waves, the wave age, as given by the ratio c /U
where U is wind-speed, is generally much greater than in the laboratory experiments
reported by Lake & Yuen (1978). It appears that typically c /U lies between 0-6
and 1*5, and hence that ak, according to Sverdrup & Munk (1947) lies between 0-3
and 0*05, but with the bulk of the observations corresponding to c /U ~ 1*0, ak ~ 0-10.
It will be interesting to see to what extent the observed values of f j f d lie near the
theoretical curve of figure 6.
This paper was prepared during February 1979 while the author was a visitor in the
Department of Engineering Sciences at the University of Florida, Gainesville, U.S.A.
The author is indebted to Dr К . Millsaps and other members of the Department for
their kind hospitality and assistance.
REFERENCES
B e n j a k i n , Т . B. <fe F eer, J. E. 1967 The disintegration of wave trains in deep water. Part I.
Theory. J . Fluid Mech. 27, 417-430.
E w c s g , J. A. 1973 Mean length of runs of high waves. J. Oeophys. Res. 78, 1933-1936.
G o d a , Y . 1970 Numerical experiments on wave statistics with spectral simulation. Rep., Port
Harbour Res. Inst., M in. Transport, Yokosuka 9, 3.
H a s s e lm a n n , K . 1973 Measurements of wind-wave growth and swell decay during the joint
North Sea Wave Project (JONSWAP). Deut. Hydrogr. Z. Suppl. A 8 (12).
L a k e , В. M. & Y u e n , H. C. 1978 A new model for nonlinear wind waves. Part 1. Physical
model and experimental evidence. J. Fluid Mech. 88, 33-62.
L o n g u e t - H ig g in s , M. S. 1962 On the statistical distribution of the heights of sea waves.
J. Mar. Res. 11, 246-266.
L o n g u e t - H ig g in s , M. 8 . 1966 The statistical analysis of a random, moving surface. Phil. Trans.
Roy. Soc. A 249, 321-000.
L o n g u e t - H ig g in s , M. S. 1976 Integral properties of periodic gravity waves of finite amplitude.
Proc. Roy. Soc. A 342, 157-174.
L o n g u e t - H ig g in s , M. S. 1978 The instabilities of gravity waves of finite amplitude in deep
water. Proc. Roy. Soc. A 360, 471-605.
L o n g u e t - H ig g in s , M. S. & С о к е ь е т , E. D. 1978 The deformation of steep surface waves on
water. II. Growth of normal-mode instabilities. Proc. Roy. Soc. A 364, 1-28.
S v e b d ru p , H. U. <b M u n k , W. H. 1947 Wind, sea and swell: Theory of relations for forecasting.
UJS. Hydrographic Office, Publ. no. 601.
1893
B y M. S. L o n g u e t - H i g g i n s , F.R.S.
Department of Applied Mathematics and Theoretical Physics ,
Silver Street, Cambridge CB3 9E W , and Institute of Oceanographic Sciences,
Wormley, Surrey G U 8 5UB, U.K.
1. I n t r o d u c t i o n
The sight of the sea surface overturning on itself, as when waves break in deep water
or on a sloping beach, is familiar to most mathematicians. Y e t surprisingly little
progress has been made in finding an analytical description of the free surface
profile and of the corresponding field of flow. Recently Longuet-Higgins & Cokelet
(1976,1978) demonstrated a purely numerical method of computation in short time-
steps, at each step solving an integral equation. But to comprehend the results of
such a computation or to handle them conveniently, the numerical calculation
needs to be complemented by some exact or approximate analysis.
A purely local solution describing how the free surface can develop a sharp
curvature near the overturning tip of a wave crest was described in a companion
paper (Longuet-Higgins 19806). Here we propose to find simple expressions to
describe the process whereby the whole body of fluid overturns. W e are concerned,
however, only with the upper part of the wave, not too far from the wave crest.
It has sometimes been suggested (for example by Price (1971)) that a wave breaks
only after attaining the limiting form described by Stokes (1880) in which the crest
has a sharp comer of 120 ° (see figure 1). Thus Price proposed an approximate
solution in which the 120° corner-flow was taken as an initial configuration. H ow
ever, observation suggests that, on the contrary, waves generally break without
passing through the Stokes configuration, and that the 120° angle is a very special
case, not generally attained. This conclusion is strengthened by the recent demon
stration that a limiting wave, with a 120° angle at the crest, actually has less energy
than is possessed by lower, symmetric waves with rounded crests (see Longuet-
Higgins & Fox 1978). Hence the limiting wave must be very difficult to attain
[ 377 ] x4-a
1894
378 M. S. Longuet-Higgins
2. D e s c r i p t i o n o f th e flo w
X = f i ? 12*, (2.1)
1895
where g denotes acceleration due to gravity. The free surface consists of the two
planes
argz = ±$rc. (2.2)
The lower streamlines are in the water. The upper streamlines represent an analytic
extension of the flow into the ‘ air’ above. Because of the branch-point at z = 0 we
have to assume a cut in the z-plane. In this instance we have taken the cut to lie
along the line argz = Jrc.
Now in an overturning wave the streamlines are not altogether as in figure 1, but
curl back over to the right, as in figure 2. This diagram represents the streamlines of
the potential
X = 2Azi, (2.3)
where A is a complex constant:
A = -a e -* . (2.4)
The streamlines are parabolae with foci at z = 0, and with axes along the line
arg z = 2e. In figure 2 we have taken e = 30°, so the axes happen to lie along the line
arg z = 60° which is part of the free surface in figure 1.
1896
380 M. S. Longuet-Higgins
Re г
3. B oundary conditions
The flow being time-dependent, the free surface is not in general a streamline, but
instead may be specified by the two conditions
p = 0, D p /D t = 0, (3.1)
where p denotes the pressure and J )/D t denotes the rate of change following a
particle.
382 M. S. Longuet-Higgins
the terms being ordered in descending powers of |ш|. Also from (3.2)
4. C o n d i t i o n s a s oj - > 00
I f A vanished for all t, then from equations (3.6) and (3.7) we should have
precisely, so that the vanishing of P implies also the vanishing of Q. From (3.5) the
free surface P = 0 is then given by
This represents the pair of straight lines &Tg((o/(o*) = ± Jtu, that is argz = ± Jtc,
as in figure 1.
When A is not zero, but (o is large, we can still ensure that Q vanishes on the same
surface as P asymptotically, by making
(4.3)
1900
384 M. S. Longuet-Higgins
that A = 0, (4.4)
and
2i(A *u )2— A w *2) — 2 = —i (A o ) + А * й )* ) (o) — (o*) —A(w2—ш * + o>*2), (4.5)
whence
2iA * = — iA — AA
\ (4.6)
—2Li = У * - Л J
and
- 2 ^ = i ( i - i * ) + A. (4.7)
From (4.6) we have
A + A * = 0, (4.8)
so that A is pure imaginary, and also
A = ii. ( 4 .9 )
Then from (4.7)
A = $i*\ (4.10)
Equation (4.4) also implies that
A = constant, A = A 0+ A t , (4.11)
and from (4.10)
where A 0and denote the values of A and F when t — 0. The constants A 0, F 0 and f 1
are at our disposal, A being related to F by equation (4.10).
I f А Ф 0 we may choose the origin of time t so that t = 0 corresponds to the
instant when A is real. In other words we may choose A 0 to be real. Assuming A 0 to
be negative, we may choose units so that
Л 0 = - 1, ( 9 = 1). (4.13)
as in figure 4. W e see then that the expression (2.5), if it is to satisfy the boundary
condition as o) -+ oo, has essentially two degrees of freedom, controlled by the two
real parameters F0 and t .
It would be possible to obtain further relations for determining FQ and by
considering lower-order terms in equation (4.3) but such relations would not
necessarily improve the fit of the boundary conditions except at large cj. Instead we
may choose and P so as to optimize the fit of the boundary conditions over the
finite part of the boundary, in the following way.
а д . ^ ) = / ( Й ) г^ (6.1)
along the free surface p = 0, and then adjust F0 and / so as to make R a minimum.
The latter course has the advantage that the boundary condition is, so to speak,
distributed over the range —oo < у < oo, rather than being concentrated at one or
two discrete points in the free surface.
Adopting the second method, we find for the values of F0 and / the following:
F0 = - 0 .1 7 6 , f 1= 0.124. (5.2)
386 M. S. Longuet-Higgins
(a)
(M
F i g u r e 6. The form of the free surface that minimizes J(DP/D*)*d« at time t = 0, for
U = V = 0, A 0 = — 1 ; jF0 = — 0.176, F = 0.124. (a ) Successive profiles of the free
surface at previous times: t = — 10, — 8, ..., 0. (6) Successive profiles of the free surface
at subsequent times: t = 0, 1 ,..., 6.
1903
The solution is nevertheless valid only for a limited range of the time t. In the
second half of this paper we shall show that, by adding to (2.6) a further term, linear
in 2, it is possible to obtain solutions in which the surface inclination exceeds 90°,
and the free surface develops a sharp comer or cusp.
6. A MORE G E N E R A L M ODEL
N o w let us adopt the more general expression
+ (UA*(o+ U*A<o*+AA*)/<o<o*. ( 6. 2)
Likewise if we substitute in equation ( 3 .3) and retain only the terms of highest
order in |w| we obtain
+ .... ( 6 .3 )
A = 0 = A* ( 6 .4 )
as before, but now
( U + 2i A * ) + A - i ( U * - 2iA) = 0 , \
388 M. S. Longuet-Higgins
The form of the free surface is obtained by writing P = 0 in (6.2). Since from
(4.11) and (6.11) we have
4 « ( 4 o- * W ) + f i ( * + K 0 < (612)
we obtain
tо*—ом* -f cj** = |i(i^ + 17) (a>—(o*)
+ i ( ^ o- J F 0 (<o*-G>**)/oxi)*
+ ( U 2+ V 2—2F0) — 2flt
z = re w, w= (6.14)
we find, to order r*1,
d = in + 8, (6.16)
where 8 is of order r-i. Thus by expanding each side of (6.16) in powers of r_* and by
successive approximation we easily obtain
rS = - , $ , ( * +
rS = + и )г * + ^ ( # - 1 и - ^ У ) 1
P + T J < 0. (6.19)
1905
The second term on the right of (6.17) or (6.18) represents an upward displacement
which is independent of r but increases linearly with the time. W e require that as
t -> — oo, so the normal displacement shall be negative, at least when r oo. This
implies that the coefficient of / in (6.17) or (6.18) shall be positive, that is
> 0. (6.20)
In the simplest case U = V = 0 discussed in §§ 4 and 6 the two conditions (6.19) and
(6.20) are clearly mutually contradictory. In the next-simplest case when V = 0,
U ^ 0, the conditions (6.19) and (6.20) imply that
Hence U < 0, in other words the imposed uniform velocity must be upwards.
A physical interpretation is as follows (see figure 6). On the forward face of the
wave the upward uniform velocity combines with the upward component of flow
in the Stokes comer-flow (seen in the present frame of reference) to produce a larger
value of the absolute velocity, compared with the velocity at an equal distance on
the left, where the velocity in the Stokes comer-flow has a downward component.
N ow from Bernoulli’s equation
gx = + + (6.22)
it follows that, at two comparable positions 1 and 2 on the right and left respectively,
since at large distances <f>t is of order only and so relatively small. The right-hand
side of (6.23) is seen to be positive, showing that the displacement of the free surface
face is downwards on the forward face but upwards on the rear face. The whole wave
therefore tends to tilt clockwise.
F i g u r e 6. A p h y s ic a l in te rp re ta tio n o f th e ro le o f th e u p w a rd s
v e r t ic a l v e lo c it y (U < 0 ) in e q u a tio n (6 .1 ).
1906
390 M. S. Longuet-Higgins
(a)
For given values of U and V we may choose the three variables A 0, F0 and J?* so as
to minimize JQ2ds taken along some central part of the free surface (say— 10
< Y < 10). In the typical case U = — 1, V = 0.5 we find the optimum values
The cause of the discrepancy can be traced to the fact that as t increases Dp/Df,
instead of remaining small, becomes of order unity near the wave crest. In general
the crest appears less sharp-pointed than in observations.
In this section we shall enquire whether the form of the velocity potential pro
posed in equation (6.1) can lead to the formation of a sharp corner at the interface,
and possibly even to a cusp.
As shown in Longuet-Higgins ( 1980a, paper I), the simplest type of sharp corner
corresponds to a saddle-point in the pressure field. The general condition for a
saddle-point, in terms of the velocity potential #(z, t), was shown in § 6 of paper I
to be
= i (7Л)
1908
392 M. S. Longuet-Higgins
F ig u r e 8. The maximum inclination a mex of the free surface as a function of the time t.
------ , From numerical computation (Longuet-Higgins & Cokelet 1978); upper time-scale.
------ , From analytic model (figure 7); lower time-scale.
(<7 being taken as unity). On substitution for x from (6.1) this gives
A+A*=-V. ( 7 .5 )
В = — (a/r)e~I(0+e>, ( 7 .7 )
(a/r)2— 2/?(a/r) + 1 = 0, ( 7 .8 )
where
p = (- t a n £ 0 + \XJr~\ sec \6) cos (0 + e ) - $ V r~ i sec £0sin(0 + e ). (7.9)
1909
a /r = f i ± (f i 2- l ) i (7.10)
provided that
fi> l. (7.11)
When ft = 1, the two roots are coincident. Given r and в (but not a or e) a necessary
and sufficient condition for (7.11) to be satisfied is that
z = x + \y = (f+iT/)2+ 2 i^ . (7.16)
c o s y = l. (7.18)
For given values of U and V, one further condition may be imposed on the
solution. It is natural to specify that D p /D t = 0 at the sharp corner itself. Now in
general we have
Dp/Vt = p t+ xivа+ XtP* ’ (7■1
p t = 0. (7.20)
From the general expression (3.2) for the pressure it follows that we must have
394 M. S. Longuet-Higgins
■*— *y /U 2
x lU 2
396 M. S. Longuet-Higgins
11. Loci of possible sharp comers at the free surface when U = —1. Solid
F ig u r e
curves show admissible solutions. Broken curve shows inadmissible solutions.
and to seek first solutions at large but finite values of r in the neighbourhoods of the
two limiting solutions in figure 10. This was done, by starting with r = 10* and
solving (7.18) and (7.22) by iteration. The value of r was then gradually reduced,
each solution being used as a starting point for the next.
O f the solutions obtained in this way, those starting from the inner of the two
solution points in figure 10 corresponded to high values of $ and were therefore
unsuitable, by the criterion (6.19). Solutions starting from the outer point in
figure 10 corresponded to smaller values of t . Some loci of possible cusps are shown
in figure 11. The corresponding values of
are generally not positive, as is required by equation (6.20). Nevertheless they can
be made small enough to be acceptable.
An example is shown in figure 12, where we have/ = —0.039, which is sufficiently
small that the slight rise in level as t -► —oo for large r may be accepted. The shape of
the forward face, while concave as r oo, is slightly convex at smaller values of r.
1913
8. D i s c u s s i o n and conclusions
W e have argued that, the flow in a breaking wave being multivalued, we must
expect at least a branch-point in the velocity-potential x • The simplest possible
branch-point is of order J, and we have shown that the three-term expression (6.1)
is capable of describing, at least qualitatively, the most obvious features of plunging
breakers, namely the overturning of the forward face and the apparent forming of
a cusp at the tip.
It should be emphasized that the solutions are valid only approximately, and over
a limited range of the time t. A t the later stages of the flow it will be necessary to
match the solution near the tip of the wave to a locally valid solution such as was
described in Longuet-Higgins ( 19806). The reason for our interest in the cusp-like
solutions of § 7 is that these particular forms, with a small crest-angle, are most likely
to be suitable for matching to the local flow.
Likewise, the earlier stages of the flow, when t is large and negative, must be
matched to the flow in the rest of the wave, in a manner similar to that used for the
almost-highest wave by Longuet-Higgins & Fox (1978), but the time-dependence
also being included.
1914
398 M. S. Longuet-Higgins
Thus we see the present solutions as essentially intermediate, in both space and
time, between expressions describing the wave as a whole, which must take account
of boundary conditions such as finite or decreasing depth of water, and the ultimate
stage of breaking described locally as a jet or sharp comer.
The expression (6.1) may possibly be the first three terms of an asymptotically
convergent series of decreasing powers of 2*. Further terms in the series could
perhaps be determined in the same way in which we have determined A (t ). However,
different forms would be required for non-progressive waves. For example, in a
standing wave, with a vertical plane of symmetry, we should require a solution with
at least two singularities, situated symmetrically on either side of the plane, or
reflecting wall.
Naturally, any of the foregoing profiles may be viewed in a frame of reference
moving with an arbitrary steady velocity U ', so that the branch-point appears to
move in the opposite direction with speed — V . Moreover we can obtain new,
solutions relative to axes with any acceleration a, on replacing the gravitational
acceleration g by g ’ = g — a and referring the new solutions to axes pointing in the
direction of g\
The following analysis provides starting values for obtaining solutions in the
more general case.
When U = 0, equation (7.9) reduces to
др/дв = 0, p= 1 (A 2)
cos0 = i ( l - V 3 ) ; 0 = -111.47°, (A 3 )
and
e — 64.41°. (A 4)
The loci of г/а = (г/а) ew are shown in figure 10. The critical point is designated by
P. W e show the quadrants — 90° > в > — 180° and — 180° > в > — 270°, the others
being obtained by reflection in the origin.
In figure 10 the contours of у are obtained from (7.17) by writing U = 0. However,
the loci for T>p(T>t = 0, obtained from (7.22), simplify very considerably in this
instance. For, since from § 6 A * = —A and equation (7.22) implies that
either
A = 0 (A 5)
or
<u-1( —iw* + A */w*) — —c.c. = 0. (A 6)
1915
Consider the first alternative (A 6). In that case Xt — 0 and the motion must be
steady (clearly an approximation). However, assuming this is so, we have from
(7.1) that
*><
(A 7)
II
hence
\ [}/о > -А /ь Р )(-\ а )* + А */ь)*) = 1» (A 8)
that is
( i - B ) ( - i + 5 * ) = 2ew. (A 9)
The real part gives
1 -B B * = 2 cos в. (A 10)
But from (7.4) we have B B* = a2/r*. Combining this with (A 10) gives
This locus is represented in figure 9 by the two dotted arcs closest to the origin in
each quadrant.
When the free surface corresponding to the above values is plotted as a function of
time, however, it is found that the profiles have unphysical features; either they are
strongly convex on their forward face, or else the limiting profiles with sharp comers
are approached from the outside , as t increases, and not from the inside as desired.
W e have shown in § 7 that this difficulty is overcome by taking non-zero values
of U and V.
This paper was begun in Cambridge and completed during a visit by the author
to the Jet Propulsion Laboratory of the California Institute of Technology,
Pasadena, between November 1979 and February 1980. For the hospitality and
assistance given him there, he is much indebted to Dr О. H. Shemdin and Professor
P. A. Saffman, and members of the Division of Earth and Space Sciences headed
by D r М. T. Chahine.
R e f e r e n c e s
400 M. S. Longuet-Higgins
Longuet-Higgins, M. S. & Cokelet, E. D. 1976 The deformation of steep surface waves on
water. I. A numerical method of computation. Proc. R . Soc. Lond. A 350, 1-26.
Longuet-Higgins, M. S. & Cokelet, E. D. 1978 The deformation of steep surface waves on
water. II. Growth of normal-mode instabilities. Proc. R . Soc. Lond. A 364, 1-28.
Longuet-Higgins, M. S. & Fox, M. J. H . 1978 Theory of the almost-highest wave. Part 2.
Matching and asymptotic expansion. J . Fluid Mech. 85, 769-786.
Price, R. K . 1971 The breaking of water waves. J . geophys. Res. 76, 1576-1581.
Stokes, Sir G. G. 1880 Supplement to a paper on the theory of oscillatory waves. Mathematical
and Physical Papers , vol. 1, pp. 314-326. Cambridge University Press.
1917
Time-dependent flows such as occur in breaking surface waves are often most con
veniently described in parametric form, with the coordinate z and velocity potential
X each expressed in terms of a third complex variable oj and the time t.
In this paper we discuss some interesting flows given in terms of elementary func
tions of a) and t. Included are the Stokes 120° corner flow, the 46° rotor or rotating
wedge, and a decelerated upwelling flow, with an exactly plane surface.
Lastly it is shown that a class of cubic flows, which are related to the plane upwelling
flow just mentioned, has a free surface that corresponds with remarkable accuracy
to the forward face of an overturning, or plunging, breaker.
1. Introduction
The phenomenon of overturning exhibited by surface gravity waves in both deep
and shallow water still lacks a satisfactory mathematical description. Figures 1 and 2
illustrate two salient features of the flow. First, it must be highly time-dependent,
with particle accelerations comparable to, or even much larger than, g. Mathematical
techniques for dealing with free-surface flows of this kind have been suggested in some
recent papers (Longuet-Higgins 1980a, b, 1981),f particularly in paper I.
The second feature, illustrated more especially by figure 2, is that when the jet
impinges on the forward face of the wave the relatively smooth flow becomes sharply
discontinuous. Mathematically, the flow can be regarded as a multivalued function.
It is as though the colliding particles are attempting to pass onto another sheet of the
Riemann surface. This in turn implies at least a simple branch point in the velocity
potential, located somewhere in the ‘tube’ of the breaking wave.
(We are here interested only in describing the flow up to the moment of impact.
Afterwards, a different type of description will become necessary, involving perhaps
turbulence and entrained flow.)
In conventional descriptions of irrotational flow, the velocity potential x i® usually
expressed directly in terms of the complex coordinate z in the plane of motion, with
the time t as an additional variable. In papers I and II a more flexible scheme was
suggested, in which x an<^ 2 were each expressed as analytic functions of a third
complex variable o>, and also t. As pointed out in § 2 below, this has the advantage
that the simplest kind of zero, where dz/dto vanishes at some point a) — a>0, say,
corresponds to the simplest kind of branch point in the velocity potential x and its
404 M . S. Longuet-Higgins
A second feature of John’s analysis was to assume the parameter ы to take only
real values at the free surface. This assumption enables a velocity potential x(w, t)
to be found in the interior, corresponding to a given motion of the free surface. For
the general problem, the assumption is probably too restrictive. Nevertheless, our
purpose here will be to examine whether certain simple flows that do accord with
John’s second assumption (namely that ш is real at the free surface) are of any uee in
describing the overturning of surface waves.
Already some results in this direction have been given in a previous paper (Longuet-
Higgins 1976). But all the flows described there were gravity-free, in the sense that
the acceleration g did not appear in the solution. The flows were appropriate to a
reference frame in free fall, or to the cases where the particle accelerations are large
compared with g. As pointed out in §6 below, gravity-free flows correspond to a
certain linear and homogeneous boundary condition for z(w, t). To include gravity,
we must solve a non-homogeneous equation. But, once a particular integral is
found, further solutions of the homogeneous equation may be added as a kind of
complementary function.
As a first example we present a simple parametric description of the well-known
Stokes corner flow, in which gravity is of course included. An analogous flow is the
‘ 45° rotor ’ discovered in paper II, for which a parametric representation is given in § 7.
We then pass on to discuss an interesting type of gravity flow (§8). This can be
described as a ‘ decelerated upwelling ’, in which the pressure gradient from the surface
velocities is exactly balanced by that due to the deceleration. Associated with this
is a family of flows given as simple polynomials in w. Remarkably, it appears that
one of the cubic polynomials agrees very closely with the observed profiles of over
turning waves.
Further discussion is given in § 14.
and an asterisk denotes the conjugate complex quantity. Suppose that гыhas a simple
zero at a) = (i)Q, where 2 = z0. Then in the neighbourhood of this point we have
z ~ zo ~ i ( w“ wo)22W (2-2)
Hence (o -(J 0cc ( z - z 0)i, (2.3)
406 M . S. Longuet-Higgins
also. This represents the simplest type of branch-point singularity. The velocity x*
is 0(z —z0)* and so less singular than in a source of vortex, where it is 0{z —z0)”1.
If, as in § 4, the velocity W * is finite at o> = w0, as well as the potential x> then by
a similar argument W — W0 is generally 0(z —z0)* and x ~ Xo *s O(z —z0)^.
3. Boundary conditions
General expressions for the pressure p and its rate of change D p /D t following a
particle were given in paper I. Thus if g denotes gravity, and if the ж-axis is taken
pointing downwards, from Bernoulli’s equation we have
in which / is a function of t only, and ‘ c.c.’ denotes the complex conjugate of the
preceding terms. Also
Some solutions making use of the equations (3.1) and (3.2) in this general form were
given in papers I I and III.
4. M ethod o f John
The above equations may be considered as a generalization of the approach of
John (1953), who assumed, in addition to the flow being Lagrangian, that the para
meter to was real at the free surface. So at the free surface we have both
W* = ztH (4.1)
# « - * » , (« ) * ? ( « ). (4-3)
and a flow satisfying at least the kinematic boundary condition may be found by
integrating with respect to (o:
{a))zf{(o)d<D. (4.4)
1921
W = ( X j z j * = zt(u>*)t (4.7)
not zt(cj), as it would be if и were Lagrangian everywhere. Only at the free surface,
where (o = oj*, is the velocity given by zt[aj).
For this reason we shall need the more general expressions forp and D p /D t given
by (3.1) and (3.2). Making use of (4.7) these may be rewritten more explicitly as follows.
First
It will be noted that this is a function of со, w* and t. I f F[u), (o*} t) is any such function,
it is clear that
^ = Ft + K F a+ K *F u., (4.9)
where К = D cj/D t is given by (3.3). In the present case, since W* = 2t(w*), we have
= <4' 13>
Hence (4.11) is equivalent to (3.2).
Lastly we note that, though the particle acceleration a at the surface, in John s
formulation, is simply equal to z^t in the interior it must be calculated from the more
general formula
a = ~ z t(w*) = ztt(w*) + K * z tu.{(o *), (4.14)
408 M . S. Longuet-Higgins
5. Frames o f reference
Given a suitable function r(w,*), real when w is real, the boundary condition
with the same integrating function r. We may call zQa particular integral of the equa
tion (5.1). Then zlt in general form, is the complementary function, and z0+ zx the
complete integral.
Since (5.2) can be derived from (5.1) simply by setting g = 0, (5.2) can be regarded
also as a gravity-free form of (5.1), and we may think of solutions to (5.2) as free-fall
solutions, that is solutions to (5.1) but seen in a reference frame moving with downward
acceleration g.
Thus, if z0(w,J) is any solution of (5.1), then
— Zq — \gt? (5.3)
is a solution to (5.2), and conversely. It is easy to verify from (4.8) that the pressure
p(u)} (o*,t) remains unaltered, apart from an additive function of the time t. The
corresponding velocities, however, will differ by the amount^, which may be important
in matching inner flows to an outer flow as t ^ —oo.
In other respects the distinction between a solution to (5.1) and (5.2) need not be
important, particularly in situations where the particle accelerations are large com
pared with g} or are highly variable. In § 6 we shall examine a flow which in one
reference frame, including gravity, is steady, but in another, free-falling frame is an
expanding, self-similar flow.
z-planc
<*)-planc
(b)
О
j ------------■*----------------- V 60° ~7777777t 7777777777*
\
\
F i g u r e 3. Parametric representation of the Stokes comer flow
(a) in the z-plane, (6) in the o>-plane.
so from (4.3) x« = (6 6)
Integration with respect to a) gives
X = ~ iV?2(w “ = I (6-7)
the well-known expression for the velocity potential in a Stokes corner flow.
Paradoxically, when t > ы the particles in the free surface, according to (6.1),
reverse their direction and slide again down the forward face of the wave, instead of
down the rear face as might be expected. The paradox is resolved by considering the
situation in the w-plane (figure 36). There it is seen that the forward face of the wave
(O A in figure 3a) corresponds to the positive real axis of со—t, but the rear face OB
corresponds to the line arg(o> —t) = The point to —t = 0 is a branch point of z,
and when и —t < 0 the particles go onto another sheet of the Riemann surface.
1924
410 M . S. Longuet-Higgins
All the particles in the interior of the fluid correspond to S{(o) < 0), and since o) is
not real the representation there is not Lagrangian. Thus the particles do not have to
follow a path on which o) is constant.
It remains to verify that the two boundary conditions (3.4) are in fact satisfied on
OB. This would follow from the symmetry of the expression (6.9) for but from (6.1)
it is less obvious. However, direct substitution into (4.8) gives us
= irzw (6.14)
with r given by (6.4). Moreover
z<2>= ±geinizS, (6.16)
for which the contours p = 0 are shown in figure 4. One branch, other than the real
axis of oj , passes through the point oj/t = 1. The branch points, given by the vanishing
1925
of zu, are at w / t = 1 and | 3i. Hence a possible domain of flow would appear to be the
shaded area shown in figure 4. However, on further investigation it appears that the
second boundary condition D p / D t == 0 is not satisfied on any of the contours p = 0
other than the real axis. Hence (6.17), and similar expressions of higher order, do not
represent complete solutions.
hence ^
* - V2 '• ( ’
Choosing e «= f 7Г, we may write
ut
X- ф * <7-6>
which identifies the flow with the 46° ‘rotating hyperbola’ discovered in paper II.
The flow is sketched in figure 5 (a).
I4 ILM 121
1926
412 M . S . Longuet-Higgins
(a) (ь)
F i g u r e 6. The 45° rotating corner flow: (a) equation (7.1); (6) equation (7.10).
In each case the flow is extended by reflection in the origin O.
So, apart from the real axis w = to* there is a second free surface p = 0 given by
arg (i) — — \ t i . Setting
О s [ ( l + i ) w - ( l - t ) w * ] e - w, (7.7)
we find also from (4.9) that
(7.8)
For points on the trailing face of the wedge ( oj = <o*) we have zt — —(1 + i)z . Hence
the particles spiral inwards towards the centre along 46° spirals.
For points on the other face, arg w = - { n , we have o)*/o) = t, hence ze - (1 —i )z
and the particles spiral outwards.
The single-term expression (7.1) represents only the special case of the ‘ 45° rotor’
when the boundary has a sharp corner. By adding to (7.1) a second term, so that
z = s ( i ) e -{l+ i)t+ M n l9 + w - i eu-ot+<*/e (7 .1 0 )
we obtain a more general flow in which the free surface is a 46° hyperbola, with rounded
vertex, rotating about its centre 2 = 0 (see figure 6b). The potential x (v, 0 satisfies
,at
(7.11)
a special case of paper II, equation (4.1). In the present formulation z((o,t) has a
branch point (гы- 0) when
o) = ee-iwe, 2 = 2e-«+<*/4f (7j3)
which lies at a focus of the hyperbola, that is in the interior of the flow. But this is
also a branch point of the velocity zt{a)*), so that altogether the flow is non-singular.
The flow just described is itself a highly special case of the class of hyperbolic flows,
with variable angle between the asymptotes, suggested in II as representing the tip
of a plunging breaker. However, their parametric representation would in general
seem to involve transcendental functions of t.
8. U pw elling flows
We now discuss a different, and in some ways simpler, class of flows than the two
types just described. These may be called ‘upwelling flows’ for reasons that will
become apparent.
We first choose a very simple form of (5.1). In fact let us take r = —2/1, independent
of a), and to save writing set . /Q 44
ь ш = w (8.1)
giving x = ~ = - 2(•
This flow is shown in figure 6. The free surface @ (w ) = 0 is the jr-axis, x = 0. The
velocity potential is
( 8. 8)
r 21
and the streamlines are yjr = —^ = constant, (8.9)
which are rectangular hyperbolae. For t > 0 the flow represents a decelerated up
welling, in which the vertical and horizontal components of flow are given by
f. f <8-10>
respectively, the axis of x being vertically downwards.
To verify that horizontal surface x = 0 is indeed a surface of constant pressure,
note that the horizontal pressure gradient is in general given by
—Py = vt + {uvx + w y), (8.11)
x+-a
1928
X
Fioube в. Streamlinee of the upwelling flow (8.7).
where (u,v) = {фх,фу), From (8.10) the terms on the right of (8.11) are equal to
— y/t*, 0 and y jt 2 respectively. Hence p v = 0 everywhere. Alternatively we may
remark that in the Bernoulli equation
the acceleration term фь is exactly balanced by the usual pressure defect - J(V0)2.
For t < 0, фх becomes positive (when x > 0), and instead of a decelerated upwelling
we have an aocelerated downwelling.
itztt (9.1)
to be satisfied when w + w* = 0.
Let us try the polynomial expression
V K = 0,
i tbn- i - nbn,
( 9 '3 )
)
Clearly bn must be of the form At + B, with A and В constant. Since the equations
(9.3) are linear, the simplest solutions of (9.2) can be separated into two fundamental
1929
P0 = t,
Pi - tw + t2,
(9.4)
P2 — trn2+ 2t*w 4- §J3,
and Q0 ш 1,
Qx - w + 2t\n |*|,
(9.6)
Q2 = w2+ 4twIn |t[ -f 4J2(ln |J|- f),
In fact these flows are limiting cases of the self-similar flows described in §9 of Longuet-
Higgins (1976), in the limit as Л -> 1.
On the other hand, the flows represented by Qn are not self-similar, nor is any com
bination of the Pn and Qn involving two or more non-zero coefficients A n or B n.
To find the velocity potential % we note that on the free surface w* = —w, so the
general boundary condition
Хш = * ? (- « 0 z e(w) (10.3)
gives хда = - -f 2t2} (Ю.4)
416 M . S . Longuet-Higgins
td + v j* — 0, 2 + 2* = 2f2, (1 0 .8 )
This vanishes on the free surface (10.8) but not on the parallel surface (10.9). In other
words, (10.9) represents a surface that is no* moving with the fluid particles.
Clearly the non-homogeneous flow (8.3) is a special case of the flow Px when seen
in an accelerated frame of reference.
and since Ww = 2тя- 41, which does not vanish at the branch point, this is a genuine
singularity. The flow must therefore be taken in some domain excluding the focus,
i.e. outside the parabola.
The corresponding velocity potential is found to be
The loci of p = 0 in the planes of w and z are shown in figure 7. On the branches of
the locus other than w + w* = 0 it is found that D p /D t does not vanish, nor ie it a
function of t only, so these do not correspond to free surfaces.
Similar remarks apply to the flow z =• Qz.
1931
417
418 M . S . Longuet-Higgins
i=P>
(b)
F i g u r e 8. The cubic upwelling z = P ,. (a) Free surface in the г-plane;
(6) contours o f p = 0 and branch points in the to-plane.
A t the free surface, w being imaginary, we may write w/t = ifi} ц real, so on separating
real and imaginary parts in (12.1),
The profile is shown in figure 8(a). I t intersects the ж-axis when ft = 0 and ±<J%,
that is in the points z/t* = $ and —-у- respectively, the latter being a double point on
the curve. Stationary values o f у occur when ft1 = § and z/i4 = - f i V t ? * *
1933
aspect latio o f the loop, that is to say the width of the loop at its widest section
divided by the distance between the vertex and the node, is
= 0-3629. (12.3)
The maximum width is one-third of the way from the vertex to the node.
The branch points o f z are given by the vanishing of
420 M . S. Longuet-Higgins
0-6
£
02
about the real axis o f z. Nevertheless, since the pressure p has a stationary point
(saddle point) on the real axis between I and n, it appears that only a slight asymmetric
perturbation o f the flow would be sufficient to alter the free surface to the form hi,
corresponding very nearly to the surface o f a plunging breaker. Near the tip o f the jet,
the perturbed flow may locally take the fqrm described in paper I I.
Figure 13 (b) shows the same flow z = P 3but at a different time t *= 0*5. Apart from
the perturbation, which is seen to be a small part of the total flow, the surfaces in
figures 13 (a) and 13 (6) are precisely similar.
1935
----- 26=C(r-/o)4
20
«5ч_
10
10 IS 20
In the expression for the pressure p the terms of highest order in w are given by
z — P 3+ iijQs (13.2)
1936
422 M . S. Longuet-Higgins
v/3-ellipse
Figure 14. Successive profiles of the free surface for the flow z = P, + 0‘ltQ,.
where tj is a real, positive constant, < 1, represents a flow in which the asymptotes
rotate clock wise, as does the nodal loop (see figure 14). This agrees qualitatively with
the numerical profiles shown in figure 10.
14. Conclusion
W e have shown how a parametric description o f the fluid flow, involving only
elementary functions o f the parameter o>, can describe some outstanding features of an
overturning gravity wave, particularly the asymptotic form of the front face of the
wave. W e have confirmed the suggestion in I I I that the analytic description probably
involves a branch point. We have also described the flow near the tip o f the ‘ je t ’, in
a special case.
A complete solution should ideally give an account o f the whole development of
the flow, including the rear face o f the wave. There is reason to believe that this may
be possible by combining or perturbing the solutions given in the present paper.
REFERENCES
J o h n , F. 1963 Two-dimensional potential flows with a free boundary. Commune Pure Appl.
Math. 6, 497-503.
L o n o u e t -H ig o in 8, M. S. 1976 Self-similar, time-dependent flows with a free surface. J. Fluid
Mech. 73, 603-620.
L o n g u e t - H i g g i n s , M. S. 1980a (Paper I) A technique for time-dependent, free-surface flows.
Proo. R. Soc. Lond. A 371, 441-461.
L o n g u e t - H i o o i n s , M. S. 19806 (Paper II) On the forming of sharp comers at a free surface.
Proc. R. Soc. Lond. A 371, 463-478.
L o n g u e t - H i g g i n s , M. S. 1981 (Paper II I ) On the overturning of gravity waves. Proc. R. Soc.
Lond. A 376, 377-400.
M i l l e r , R. L. 1967 Role of vortices in surf zone prediction, sedimentation and wave forces.
In Beach and Nearshore Sedimentation (ed. R. A. Davis & R. L. Ethington), pp. 92-114.
Tulsa, Oklahoma: Soc. of Economic Palaeontologists and Mineralogists.
424 M . S. Longuet-Higgins
N e w , A. L. 1981 Breaking waves in water of finite depth, British Thtor. Mech. Colloq., Bradford,
6-9 April 1981 (abstract only).
Stokes, G. G. 1880 On the theory of oscillatory waves. Appendix B : Considerations relative to
the greatest height of oscillatory waves which can be propagated without change of form.
Mathematical Physical Papers, vol. I , p p . 225-228. Cambridge University Press.
V in je, T. Sc B re v io , P. 1981 Breaking waves on finite water depths: a numerical study. Ship
Res. Inst, of Norway, Rep. R - 111.81.
1939
Experiments have shown that bubbles approaching an air-water interface give rise
to axisymmetric jets projected upwards into the air. Similar jets occur during the
collapse o f cavitation bubbles near a solid surface. In this paper we show that such
jets are well modelled by a Dirichlet hyperboloid, a hyperbolic form o f the
better-known ellipsoid. The vertex angle o f the hyperboloid is calculated as a function
o f time and found to agree with the observations o f Blake & Gibson (1981) and others.
The jet is initiated, according to this model, when the vertex angle passes through
2 arctan\/2, or 109-47°, at which instant the fluid accelerations become large. This
compares with a vertical angle of 90° in the corresponding two-dimensional flow.
Further experiments demonstrate that an axisymmetric standing wave, when
driven beyond its maximum amplitude, can break by throwing up a jet of the same
hyperbolic form. Hyperbolic jets may occur commonly in free-surface flows.
1. Introduction
In a recent paper Blake & Gibson (1981) studied the growth and collapse of vapour
bubbles close to a free surface, both numerically and experimentally. One o f their
experiments is reproduced, in modified form, in figure 1. As the bubble rises and
expands it can be seen to raise and accelerate a thin layer o f fluid in the form of a
spherical dome. Then (starting at about frame no. 30) there emerges a remarkable
jet rising from the dome along the axis of symmetry. The jet ultimately approximates
a circular cone with diminishing vertex angle. Such a phenomenon is clearly of
environmental interest for transfer o f water droplets, salt nuclei and organic particles
from the-ocean to the atmosphere by bursting bubbles (see Blanchard & Woodcock
1980). As indicated by Blake & Gibson, similar jets issuing from the interior surface
o f a cavitation bubble near a solid surface, and directed towards the surface, are
probably responsible for much of the damage to propellers and other fast-driven
hydraulic machinery.
The numerical calculations described by Blake & Gibson agree with their observa
tions as far as the initial formation o f the jet. However, the mechanism still appears
somewhat mysterious. In the present paper we propose an analytic model for the
development of the jet. The model is in fact a hyperbolic form o f the flow known in
astrophysics as a Dirichlet ellipsoid, and described for example in Lamb (1932). The
hyperbolic form, however, does not appear to have been seriously considered, perhaps
on account o f the infinite masses involved, and certain singularities in the time
dependence, described below. A two-dimensional hyperbolic form has been studied by
Longuet-Higgins (1972, 1976) in connection with the breaking o f surface waves, and
generalized in more recent papers (1980, 1982).
1940
104 M . S. Longuet-Higgins
Frame no. 10
F i g u r e 1. J e t produced b y a gas bubble approaching a free surface (a fte r B la k e & G ibson 1981).
T h e dashed line represents the tra je cto ry o f the assumed centre o f the h yp erb oloid.
The theory for the three-dimensional hyperboloid is given below in §2, and the case
o f axial symmetry is discussed in detail in §3. One special feature o f the flow is the
occurrence o f a ‘ jo lt’ , or local infinity in the acceleration when the vertex angle o f
the hyperboloid equals 2 arctan\/2, or 109*5°. This in fact corresponds to the initiation
o f the jet. A comparison with Blake & Gibson’s observations follows in §4, with
reasonable agreement (see figure 4).
The degree of correspondence between the observations and this simple theoretical
model prompted the author to enquire whether a similar flow might be observed in
other situations. One example that came to mind was the jet thrown up by a standing
water wave when excited by a vertical or horizontal oscillation of the container
(Faraday 1831). In §6 we describe a simple experiment which in effect verifies that
these jets are, in suitable circumstances, well described by the same theory.
In the appendix we enquire further into the initial ‘ jo lt’ and give a physical
explanation for it. Finally, in §7 we summarize the conclusions and note possible
implications for the modelling of a breaking wave.
1941
where a, 6, с are functions of the time t only, and a dot (") denotes time-differentiation.
Equation (2.1) represents, in fact, the most general expression for a pure straining
flow that is symmetric in regard to the three coordinate planes. The velocity vector
!* !* )• (2-2)
and by continuity we have a b с __
J J - + т + - = 0, (2.3)
a b с
so that abc = 2L 3, (2.4)
is a constant o f the motion.
The fluid being assumed homogeneous and o f unit density and gravity negligible,!
the pressure p at any point is given by
Surface tension being also neglected, we may take as boundary conditions that both
p and D p /D t vanish on the same surface (cf. Longuet-Higgins 1972, 1976). Thus the
coefficients o f corresponding terms in (2.5) and (2.6) must be in proportion. Hence
* + 1 * + * £ + « ' (2 .7 )
a a b b с с г
a2 b2 c2 (2Л° )
a hyperboloid with semi-axes a, b and c. Moreover, the continuity condition (2.3) can
be written aa — bb—cc = 0, (2.11)
106 M . S. Longuet-Higgins
so t 12 , 14
b — —, 0 = ----- r-, (2.15)
a a2
(2.16)
' 0 - Э - *
hence p®/
(2.17)
± Г ( 1“ 5 = )а *-
On the other hand we may have
Axisymmetric flow. This corresponds to
с = 6, (2.18)
so ab2 = 2L 3, аг - 2 Ъ г = и г \ (2.19)
and
(2.21)
hence
(2.22)
- ± f !'1
In both (2.17) and (2.22) the right-hand side is expressible as an elliptic integral.
I t is noteworthy that, in the general case, a second surface p = ps(t) may be found
from (2.5) and (2.6) provided that p = ps and D p /D t = ps represent the same surface.
The equation o f the surface must clearly be
x2 у2 z2 F+2ps
(2.23)
a2 fe2 c2 F
from (2.5), and from (2.6) a similar equation with (/ + 2ps) /Z1on the right. I f we then
take ps = AF (2.24)
with Л constant, the new surface is a hyperboloid similar to (2.10), and with semi-axes
a', b ', c' given by ,
- = i = - = ( l + 2A)l. (2.25)
a b с
This represents a hyperboloid o f two sheets, having the x-axis as axis of symmetry
(see figure 2). A section by the plane z — 0 gives the hyperbola
^ ! _ 2 ' ! == i (3-2)
a2 b2
To fix the ideas we shall discuss the lower branch, for which x > 0.
the lower limit being suitably chosen. The angle between the asymptotes is 2y, where
tan у = —
6
= —
f i
= —
21
j w-o)
a a oc*
108 M . S. Longuet-Higgins
V (3 -8 )
Lastly „ .. §t/!a - 3
F = - a a = - lP a a „ = (1 _ a _ 3)2- (3.9)
(1—a -3)2'
Consider the limit o f (3.5) as a - » oo. When a > 1 we have
)da, (3.10)
r = 1,
and л 2L
Й ~ { Г ^ 5 * [1 'Н (Т _Т 0 ) (3 ,5)
I t can be shown that the limiting form of the surface is a paraboloid whose dimensions
contract like t~2 about a point of similarity 0 lying on the axis o f symmetry at
one-quarter the distance from the vertex of the parabola to the focus. This limiting
form is one of the self-similar flows discussed in Longuet-Higgins (1976, §3).
Consider on the other hand the limit as а - » 1. Writing a = 1 + y , rj > 0, and
substituting in (3.5), we have
= £ [ i - ( i + 7 ) - 3)]irf7 ~ - ^ i , i . (3.16)
Hence
а ~ 1 + ( ^ г т ) 1, ( 3 1 7 )
(3.18)
Also, 2* ,
(3.19)
Thus as t -+0 the velocities (represented by a) tend to infinity like l~b and the
accelerations (also the normal pressure gradient) tend to infinity like Hence the
motion starts with a weak ‘ jo lt’ or shock, just as in the well-known two-dimensional
case (Longuet-Higgins 1972). Further discussion o f this phenomenon is given in the
appendix.
Equation (3.19) shows that the limiting angle between the asymptotes is
lim 2y = 2 arctan 2* = 109-47° (3.20)
compared with 90°, or 2 arctan 1*, in the two-dimensional case.
1945
110 M . S. Longuet-Higgins
The function r was calculated numerically by quadrature over the complete range
1 ^ a < 10. Representative values are given in table 1.
Figure 3 shows the development o f the hyperbola as a function o f the dimensionless
time т. For illustration we have shown not only the free surface p — 0, but also a
second surface p = ps = AF(t), with Л = 1*5.
W e note that values o f a lying in the range (0 < a < 1) correspond to angles 2y
greater than (3.20). However, we then have d2 < 2b2 and U2 would be negative. The
normal pressure gradient at the free surface also changes sign, and there is reason
to doubt that the flow is stable (cf. Longuet-Higgins 1972).
a r 2y R /l F /U 2
100 0 0000 109-47° 2-0000 1-5000
1-04 00090 106-26° 1-8491 1-3335
116 00676 97-08° 1-4863 0-9610
1*36 0-2078 83-45° 10813 0-5963
1-64 0-4417 67-91° 0-7436 0-3401
2-00 0-7798 53-13° 0-5000 01875
2*44 1-1887 40-71° 0-3359 01033
2-96 1-6951 3104° 0-2283 00578
3*56 2-2862 23-78° 0-1578 00332
4-24 2-9604 18-40° 0-1112 0-0197
500 3-7165 14-42° 0-0800 0-0120
5*84 4-5538 11-44° 00586 0-0075
6-76 5-4719 9-20° 00438 00049
7-76 6-4706 7-49° 00332 0-0032
8-84 7*5496 6-16° 0-0256 0-0022
1000 8-7089 512° 00200 0-0015
Frame number 2y r
30 104° 0
40 47° 1-125
50 24° 2-25
60 15° 3*375
70 10° 4*5
80 8*5° 5-625
90 7° 6-75
100 6-5° 12-6
T a b l e 2. Measured values of the angle 2y between the asymptotes in figure I
1947
F i g u r e 4. Time history of the angle 2y between the asymptotes according to (3.6). The plotted
points correspond to the observations in figure 1 .
In theory, since the reference frame in §§2 and 3 above is inertial, i.e. in free fall,
the trajectory o f the centre С should actually be in a free-fall parabola. However,
the elapsed time between frames 30 and 100 - 5*5 ms - is so short that the curvature
o f the trajectory is quite negligible: the maximum deflection would be about one part
in 103.
To obtain a dimensionless time r, it was noted that the angles 2у corresponding
to the two frame numbers 50 and 80 were 24° and 7°, corresponding to r = 2*25 and
6-75 respectively. The values o f r for the other frames were then inferred by linear
interpolation (and extrapolation) giving the numbers shown in column 3 of table 2.
These have been plotted in figure 4. The agreement with the theoretical curve will
be seen to be quite good. Especially notable is the observed angle 104° in frame no.
30, which is only a little less than the critical angle o f 109-5°. I t appears a coincidence
that this was one o f the frames selected for display.
In figure 10 of Blake & Gibson (1981) there was one further figure (10.11)
corresponding to a frame number 142. Extrapolation gives for this frame r = 12-6
and hence, from the model o f §3, 2у = 3-6°. The angle measured from the photograph
is about 3°, so that the agreement is maintained even as far as this point.
Near the base o f the jet in figure 1, where fluid is apparently being drawn into the
hyperboloid, close agreement is not to be expected. In fact the photographs from
frame 50 onwards show the formation o f a smaller, reverse jet directed into the cavity,
which by frame 100 has actually penetrated the lower cavity wall. Nevertheless, given
the unplanned character of the initial conditions, it is remarkable that the form of
the upwards jet should be so stable.
1948
112 M . S. Longuet-Higgins
The downwards jet appears similar to that produced by a shaped charge (see
Birkhoff et al. 1948). These authors model the phenomenon by a steady flow. Their
observations, however, indicate a uniform rate o f strain in the jets, more in accord
with the present theory.
The asymmetrical collapse o f cavitation bubbles in the interior o f a fluid or near
a solid surface has been discussed by Benjamin & Ellis (1966), Plesset & Chapman
(1971) and many others. From the theory o f §3 one would expect the maximum
acceleration o f the inwards jet to occur when the ‘ angle o f indentation ’ of the bubble
reached a critical value o f around 109-5°. Such a result appears not inconsistent with
the numerical calculations shown in figures 1 and 2 o f Plesset & Chapman (1971),
for example. However the development o f an inwards jet is necessarily limited in time
and space by the presence o f the cavity walls, so that comparison with the ideal
hyperbolic form is less feasible than for a bubble breaking a free surface.
Jets formed by a hollow liquid surface (without lining) have been produced
artificially by Bowden (1966), and are probably quite a common phenomenon. They
appear to be similar to the jets arising from the bowl o f any small bubble (diameter
1-2 mm) bursting at a free surface (see Blanchard & Woodcock 1980; MacIntyre 1968,
figure 1). Similar jets may model the ‘ splash’ produced after a solid body has fallen
rapidly through a free surface; see Worthington (1908).
114 M . S. Longuet-Higgins
F i g u r e 5. Glass beaker placed on a vertical Derritron vibrator, in a state of rest. The vertical flanges
on the axis are to suppress asymmetric modes of oscillation.
Figure 6. Crest profiles of successive standing waves (frequency approximately 3*43 Hz) excited subharmonically
by a vertical oscillation (frequency 6*64 Hz). The interval between frames is about 0-29 s.
1951
115
М . S. Longuet-Higgins
1953
7. Discussion
W e have shown that the axisymmetric jet produced by a gas bubble near a free
surface is well modelled by the inertial, inviscid, hyperbolic flow described in §3.
Moreover, a very similar jet has been shown to occur in axisymmetric standing waves
driven beyond their limiting amplitude. By inference we may expect that such jets
are a possible characteristic o f other types of unsteady free-surface flow. In particular,
we may take note o f a previous suggestion (Longuet-Higgins 1980) that the tip of
an overturning gravity wave may be suitably modelled by a hyperbolic jet of
generalized form, in which the principal axes are in rotation.
I t should be emphasized that the jets discussed in this paper are to a high degree
inertial, that is independent o f both gravity and surface tension. In the initial stages
o f formation, the time scale is too short for these other forces to exert an effect. A t
later stages both gravity and surface tension may become effective. Gravity causes
the jet to fall back, and the origin to describe a free-fall trajectory. Surface tension
can cause the jet to break up into droplets.
However, the evident stability o f the jet under surface tension is remarkable. The
reason may be that the flow is not simply a uniform inextensible stream, as is assumed
in the well-known perturbation analysis o f a jet of circular cross-section (see Lamb
1932, §274). In the hyperbolic jet, on the contrary, the free surface is constantly being
extended, with the consequence that any wavelike perturbation of the flow is being
drained o f energy by the radiation stresses. Conversely, if time were reversed so that
the free surface were being contracted, the flow would be highly unstable, quite
regardless o f surface tension.
In two-dimensional breaking waves o f plunging type, the jet is continually being
stretched in the direction o f motion, but not in the transverse direction. One therefore
expects instabilities to appear at first across the flow, with crests aligned parallel to
the jet. This is what is commonly observed.
The author is indebted to John Blake for arousing his interest in cavitation bubbles
and for subsequent stimulating discussions, and to Norman Smith for assistance with
the experiments described in §6.
118 M . S. Longuet-Higgins
F i g u r e 8. Illu stration o f the stream lines and the free surface in a tw o-dim en sion al wedge. (T h e
v e rte x angle 2у is shown increasing.)
which the position (ж, у) o f a particle is given as a function o f t and the initial position
1 = J M L = a6, (A 2)
d{x0,y0)
A direct but very brief analytical proof is as follows. From (7.3) we have
- + y
x - = o, (A 6)
x у
hence
(A 7)
x у
say. Thus
^ -(х г - у г) = 2 (x x -y y )
at
= 2 G (xx+yy)
where ( r,6 ) are polar coordinates in the plane. The expression in brackets on the right
represents the radial acceleration, which vanishes identically. Hence x2—y2 is a
positive constant. But, because of the symmetry of the trajectory about в = 45°, we
see that when x - * y then x2tends to y2. These simultaneous requirements are possible
only if x and у each tend to infinity. In other words, the velocity becomes infinite.
Turning again to the axisymmetric case we can give a very similar physical
explanation. The essential difference is that in the axisymmetric case the trajectory
o f a given particle is no longer symmetric about a line through 0 inclined at 45° to
the x-axis. Instead, the tangent to the trajectory becomes normal to the radius vector
at a different angle у = arctan y/2 , as shown in figure 9.
A formal proof is as follows. I f p denotes (y2+ z2)b, the equation o f continuity now
requires that each trajectory satisfy
hence
* = - ^ = G, (A 10)
x P
1956
120 M . S. Longuet-Higgins
say; therefore
j i (x2- 2 p i ) = 2 (x x -2 p p )
= 2 G (xx+p p )
REFERENCES
B e n ja m in , Т. B . & E l l i s , A. T. 1966 The collapse of cavitation bubbles and the pressures thereby
produced against solid boundaries. Phil. Trans. R. Soc. Lond. A 260, 221-240.
B e n ja m in , Т. B. & U r s e l l , F.
1954 The stability of the plane free surface of a liquid in vertical
periodic motion. Proc. R. Soc. Lond. A 225, 505-515.
B i r k h o f f , G., M a c D o u g a l l , D . P ., P u g h , E. M. & T a y l o r , G. I. 1948 Explosives with lined
cavities. J. Appl. Phys. 19, 563-582.
B l a k e , J. R. & G ib s o n , D. C.
1981 Growth and collapse of a vapour cavity near a free surface.
J. Fluid Mech. I l l , 123-140.
D. C. & W o o d c o c k , A. H. 1980 The production, concentration and vertical
B la n c h a r d ,
distribution of the sea-salt aerosol. Ann. New York Acad. Sci. 338, 330-347.
B o w d e n , F. P. 1966 The formation of microjets in liquids under the influence of impact or shock.
Phil. Trans. R. Soc. Lond. A 260, 94-95.
D i r i c h le t , P. L. 1860 Untersuchungen uberein Problem der Hydrodynamik. Abh. Kon. Ges. №4'ss.
Gottingen 8 , 3-42.
E d g e , R. D. & W a l t e r s , G. 1964 Period of standing gravity waves of largest amplitude on water.
J. Geophys. Res. 69, 1674-1675.
F a r a d a y , M. 1831 On a peculiar class of acoustical figures, and on certain forms assumed by groups
of particles on vibrating elastic surfaces. Phil. Trans. R. Soc. Lond., pp. 299-340.
F u l t z , D. & M u r t y , T. S.
1963 Experiments on the frequency of finite-amplitude axisymmetric
gravity waves in a circular cylinder. J. Geophys. Res. 68, 1457-1462.
H a u b r i c h , R. A., M u n k , W . H . & S n o d g r a s s , F. E. 1963 Comparative spectra of microseisms
and swell. Bull. Seism. Soc. Am. 53, 27-38.
Lam b, H . 1932 Introduction to Hydrodynamics. 6th edn. Cambridge University Press.
L o n g u e t - H i g g i n s , M. S. 1950 A theory of the origin of microseisms. Phil. Trans. R. Soc. Lond.
A 243, 1-35.
L o n g u e t - H i g g i n s , M. S. 1972 A class of exact, time-dependent, free-surface flows. J. Fluid Mech.
55, 529-543.
L o n g u e t - H i g g i n s , M. S. 1976 Self-similar, time-dependent flows with a free surface. J. Fluid
Mech. 73, 603-620.
L o n g u e t - H i g g i n s , M. S. 1980 On the forming of sharp corners at a free surface. Proc. R. Soc. Lond.
A 371, 453-478.
L o n g u e t - H i g g i n s , M. S.
1983 Rotating hyperbolic flow: particle trajectories and parametric
representation. Q. J. Mech. Appl. Math. 36 (May 1983).
L o n g u e t - H i g g i n s , M . S. & U r s e l l , F. 1948 Sea waves and microseisms. Nature 162, 700.
M a c I n t y r e , F. 1968 Bubbles: a boundary-layer ‘ microtome ’ for micron-thick samples of a liquid
surface. J. Phys. Chem. 72, 589-592.
M c I v e r , P . & P e r e g r i n e , D. H. 1981 Comparison of numerical and analytical results for waves
that are starting to break. In Proc. Symp. on Hydrodynamics in Ocean Engineering, Trondheim,
Norway, August 1981, pp. 203-215. University of Trondheim.
M a c k , L. R. 1962 Periodic, finite-amplitude, axisymmetric gravity waves. J. Geophys. Res. 67,
829-843.
1957
R O T A T I N G H Y P E R B O L I C FLOW:
PARTICLE TRAJECTORIES
AND PARAMETRIC REPRESENTATION
By M. S. L O N G U E T -H IG G IN S
SUM M ARY
A class of time-dependent, irrotational flows in which the velocity potential has the
form \ A z 2 (where z is a complex coordinate and A is a complex function of the
time) are of special interest in relation to the flow in the jet of a plunging breaker. In
this paper we calculate the trajectories of marked particles, and make use of the
result to construct a semi-Lagrangian representation of the flow. Special attention is
paid to the limit when the free-surface profile tends to an infinitely thin hyperbola,
and to another special case (the “ 45°-rotor”) in which the axes rotate with uniform
angular velocity, and the angle between them is constant at 45°.
1. Introduction
The problem of describing mathematically the breaking of gravity waves in
deep or shallow water involves, at the very least, solving the equations of
motion for time-dependent inviscid flow with a free surface. Such solutions
are rare, but an interesting class of flows, which may describe the flow in the
tip of a plunging breaker, was recently pointed out by the present author ( 1 ).
In these the surface takes the form of a hyperbola, in which the origin falls
freely in a parabolic path while the principal axes rotate non-uniformly in
time. The angle between the asymptotes also varies, and in an important
class of cases it tends to zero, giving rise to a slender or sharp-pointed jet.
The solutions in (1) were given essentially in Eulerian coordinates, using a
velocity potential \ as a function of the complex coordinate z and of the
time t. If the flows were steady, the particle trajectories would be given
simply by the streamlines Im \ = constant. But in a time-dependent flow the
streamlines continually vary, and only touch the trajectories instantaneously.
The first aim of the present paper is to determine the particle trajectories
analytically.
A second aim is more far-reaching. As shown in a companion paper (2)
the complete description of an overturning wave probably will involve a
solution given in parametric form, with \ and z each expressed in terms of a
third complex variable o>, and t. Now in the simplest type of parametric
representation, due originally to John (3), it is assumed that at the free
surface w takes only real values, and moreover that at the surface (but not
with a and cr real functions of the time t. It was shown in (1) that (2.1)
represents a flow with a time-dependent free surface, attached to the
particles, provided that
cr, = A a 2 (2.3)
and
aa„ - 4a f + 2a4- \2a 6= 0, (2.4)
A ,z 2 4 -2 A A * z z * + A * z * 2= ~2S<*4, (2.7)
(3.2)
Hence
(3.4)
(3.5)
250 M . S. L O N G U E T - H IG G IN S
(312)
where
ш = Р/А4.
J
2 = R exp ^A-1 а " 1е,фdcrj (3.13)
in which the constant R depends on the lower limit in the integral. For
points on the free surface, R may be related to S by substitution in equation
(2.7).
tw ~^ Ur + Jt) = 0. (4.4)
This generally has two independent solutions t](1) and V 2) say, so that
where с = bla. So the solution to (4.2) involves only one arbitrary constant c,
as we would expect. T o find z we have now from (4.1) that
zf = A * z * , z* = A z (5.1)
zn = A * z * + A * z * t (5.2)
and substitute for z * and z* from (5.1). This gives
252 M. S. L O N G U E T - H IG G IN S
Dividing by A * and noting that z (1>* and z (2)* are independent we have
L = L *, M = M* (5.8)
i.e. L and M are both rea/. Conversely, if L and M are real, then (5.5) is a
solution of (5.1).
Now let z = z (1) denote any solution of (5.1) and hence (5.4). A second
solution Z to (5.4) may be obtained by writing
Z = uz(1), (5.9)
Hence
(5.12)
ц z A*
which on integration gives
u, = iK A *fz 2, (5.13)
that is
* { a * z * J ( A * / z 2) d t+ A * / z J = - e - ‘M * z * J ( A / z * 2) dr, (5.17)
1964
R O T A T IN G H Y P E R B O L IC F L O W 253
e i7| (A * / z 2) d t = Ф, (5.18)
we obtain
г г * (Ф + Ф * ) + е * = 0 . (5.19)
Since the first terms are real, it follows that e'y = ± l . Without loss of
generality, take eiy = l. Then (5.19) becomes
z z * / (z z \ = 0; (5.22)
in other words the lower limit of integration in (5.14) must correspond to
|z|=o°.
T o summarise, we have shown that if z ~ z (l) is any solution of (5.1) and
(5.4), then a second solution z = z (2) is given by
z ^ = i z ^ \ ^ - 2dt, (5.23)
z„ = A A * z + A * z * , (5.25)
so that the equation (2.7) of the free surface can also be written
254 M. S. L O N G U E T - H IG G IN S
6. The case Л = 0
In this very special case, equations (2.3) and (2.8) show that crt and 8, both
vanish, and the principal axes of the hyperbola are stationary. Accordingly
we may take
<r = 0, A = a, real, (6.1)
z„ = z, (6.3)
so that
and substitution in (3.3) or (5.1) shows that L and M are both real.
Moreover the functions
that is to say they are rectangular hyperbolae, the origin assumed stationary.
By (2.4) the function a satisfies
aa„ —4a? + 2 a 4 = 0 ( 6.8)
and hence
Н д т т к ч !- ( 6 '9 )
7. The 45°-rotor
An interesting case mentioned in (1) is when m = J and
Aa = constant = V2, (7.1)
1966
R O T A T IN G H Y P E R B O L IC F L O W 255
so that by (2.3)
Aat = 2 , (7.2)
and
The angle between the asymptotes is constant at 45°, and the whole flow
rotates with a constant angular velocity - 8 t = - A -1 (see Fig. 1).
On choosing the time scale so that A = 1, equation (5.4) becomes simply
zn + 2izt - 2 z = 0. (7.4)
Thus
25 6 M . S. L O N G U E T - H IG G IN S
Therefore
(7.7)
(7.8)
(7.9)
L M —2S. (7.11)
The same result can be obtained, though at greater length, by solving
equation (4.4) for 17 and making the substitutions indicated in sections 3 and
4.
Consider (7.9); when t is large and negative the first term on the right of
(7.9) is dominant and so
zt e —( l + i)z. (7.12)
This is the equation of a 45° equiangular spiral. The particles are far out
along the trailing face of the rotating hyperbola and are spiralling inwards
towards the vertex (see Fig. 2). On the other hand when t is large and
positive the second term is dominant and so
(7.13)
The particles are far out on the leading face of the hyperbola and are
spiralling outwards.
A t intermediate values of t the two terms in (7.9) may be comparable. In
fact the vertex of the hyperbola corresponds to
8. Slender hyperbolae
In one particularly interesting class of flows described in (1), the angle 7
between the asymptotes tends to zero as * —►«>, while the orientation 6 of
the axes tends towards a constant limiting value. Such flows may be of
particular interest for comparison with the tip of a plunging breaker.
1968
cr = - a - i a 3+ ( j b w - & ) a 5 + . . . , (8.1)
1969
258 M. S. L O N G U E T - H IG G IN S
and hence
= ~cr + б€Г3 —(io^r + Т2б)сг 5 + . . . , (8.2)
OJ- 1 = _ 0г-1_1 0.+ (Х ш __7_)0.з_|„ ? ( 8.3)
p ( p - l ) + 2p = 0. (8.5)
Hence p = - 1 or 0, and we find for the two independent solutions
and hence
leading to
z « o - - 1 - i- (J + | ic )o - - £ c o -2 + ....... (8.9)
To express this result in terms of r, we note that when m > J the integrand
in equation (2.6) is always negative. Hence t is monotonic in a, and we may
choose the lower limit of integration as a = oo. Then
(8 .10)
Ja a (1 - a +WGL p
making t —>0 as a —><» and vice versa. When |a2 —w a 4|<l,. the integrand
may be expanded in powers of a. Hence
, = tl+Ji-ia+(iOT-J)a3-...}, (8.11)
where is a constant, given by
ti = lim ( f - —Y ( 8. 12 )
a-+0 \ aJ
For numerical values of ti see Table 1. The series on the right of (8.11)
involves only odd powers of a. Thus writing
(8.13)
1970
R O T A T IN G H Y P E R B O L IC F LO W 259
20 -0-5568 1*7175
1*0 -0*2656 2-1565
0-5 0-0997 2-8542
0-4 0*2698 3*2146
0-3 0*6413 40196
0-25 oo 00
1 1 /w 3\ 1 /ОЛА.
а ~ Т ~ 2 Р + \ 6 + 8 ) ^ “ - - - ( 8 Л 4 )
1 1 / т а г 1 \ 1 / o i c 4
° = — +^ ~ \ T s + (815)
From (8.9) and (8.15) we find that
K = R *. (8.17)
Hence R is real.
The solution corresponding to JR = 1 and с = 0 is clearly
z (1) = ( ' + * + . . . , (8.18)
(8.19)
This shows further that in (8.16) the parameter с occurs at most linearly in
the coefficients of t,n.
T o carry the expansions further it is now easiest to use equation (5.4) in
the form
(«t /а “ ia 2)z f ~ = 0, (8.21)
1971
260 M. S. L O N G U E T - H I G G I N S
. ш InУ
zw= f'+, ....
24t 120t'
(8.22)
,( 2) _ i 1 i
+ ---------,
2*' 6 f'2 6 f'3 10f'4
v j L 2- M 2 = 2S. (8.24)
9. A semi-Lagrangian representation
For the reasons given in section 1, we are interested in finding for the flow
(2.1) a Lagrangian or semi-Lagrangian representation of the type proposed
by John (3). In this, the position coordinate z is expressed as a function of a
complex variable <o and of the time, in such a way that on the free surface <o
is real and also Lagrangian, that is to say the velocity and acceleration are
given by zt(<o, t) and z„(o>, f). In general, however, the velocity equals
z,(oj*, f), not z,(<t>, f). Only on the free surface are these expressions
equivalent.
The representation has the advantage, first, that the free-surface con
dition can be expressed in the simple form
(9.1)
where r(co, r) is to be real when <o is real. Second, the velocity potential, if
required, can be found from
X «,=X 2z„ = z?(<»)z„(<o) (9.2)
which vanishes only at a focus o f the hyperbola, that is in the interior of the
fluid. But
and hence
F„ = irG ,}
(9.9)
G tt = irF,l
where r must be real for real со.
The non-rotating case of section 6 corresponds to F real and G pure
imaginary, say G = »G'. Then (9.7) becomes
tn + 2aat ] ^ q n- 2 o
^ +2а^ + 2 а = ^ ф - 2 а , (9.14)
a ,+<*2 a, —a
r2 = - F ttG\JFG' = a? - a 4. (9.16)
1973
262 M. S. L O N G U E T - H I G G I N S
a? = a 4( l + P a 4), (9.17)
so that
Then
z = F (r)cosh a> + G(0sinh<«) ( 10 .2)
is a semi-Lagrangian representation of the flow (2.1) if we take
(a ,00,, t),'
F = (2 S / w № Г(а, f),l
(10.3)
G = (2 S )iH c(<
(a,0, f). J
Proof. W e have to show that these expressions satisfy equation (9.7) and
also equations (9.9) with a suitable choice of r.
First, since z = H (a , c, r) is a solution of equation (3.1) we must have
H* = AH (10.4)
as required.
T o prove (9.9) we establish the following lemma.
R O T A T IN G H Y P E R B O L IC F L O W 263
Again differentiating each side and substituting for z„ from (5.27) we find
But by (2.6)
a2 = a 4( l —a 2 + tzra4). (10.15)
So (10.14) gives
£tl = iN~1/aja4Z = irZy (10.16)
provided that
N 2 = ш. (10.17)
264 M. S. L O N G U E T - H IG G IN S
But for points on the free surface, equation (2.7) is satisfied by z = fz(1).
Hence equation (10.20) is indeed satisfied provided that
W e have then only to verify equation (10.18) for one particular value of f, or
in the limit as t —> °°.
Consider now the solution of section 8. From the expansion (8.22) we
easily find that the two sides of (10.18) are equal asymptotically provided
Nm/l = - w , (10.23)
that is
m/l = m*, (10.24)
and comparing (10.22) and (10.24) we have
11. Discussion
For points on the free surface, the trajectories given by the two represen
tations of sections 8 and 10 should be equivalent.
T o see this, we may set in (8.24)
L - (2S/or)* cosh o>, M = (2S)* sinh со, (11.1)
where z (1) and z (2) are the two functions given in (8.22). It will be seen that
this agrees precisely with the representation of section 10 .
Consider the interpretation of (11.3). Equations (8.22) show that, in the
limit as t —> °o,
z (1)~ r ' + i , z (2)----- i/2t\ (11.4)
so
z ~ {2 S !m )K t'+ i) cosh a>. (11.5)
1976
R O T A T IN G H Y P E R B O L IC F L O W 265
The point a) = 0 is at z ~(2S/w )H f' + i), which travels with constant limiting
velocity
C = (2S/m)K ( И . 6)
and if we take the origin at this point, the limiting position of each particle
as t —> oo is given by
x' = |(2SM*ta>2,
( 11 .8)
y' = -(2 S )W 2 r.
Since the transverse angular velocity of a particle is, by (3.7) and (6.9),
0t = + о-), ~ a 2( l + со-), (11.10)
the particle gains on the tip of the hyperbola by a relative angular velocity
W e note that for the 45°-rotor, in which the angle between the asymptotes
does not tend to zeo at any time r, a semi-Lagrangian representation
corresponding to ( 10 .2 ) is
where
F = (2S)*H (a,0, t),
(П .1 3 )
G = (2S)*HC(<
(a ,0 , (),J
and
and o) is real. These values satisfy (7.11). Moreover, the necessary condi
tions (9.1) and (9.3) are also satisfied provided
r = 2. (11.16)
1977
266 M. S. L O N G U E T - H IG G IN S
It can also be shown that z = H (a, c, t)y с real, represents the general
particle trajectory as derived by the method of section 4.
( 12. 1)
and
(12.2)
(12.3)
1 1
t~f+ h, o '------ (12.4)
I
provided that
(12.5)
and
( 12.6)
Some values of crQ and tj calculated in this way are given in Table 1.
T o calculate z (1) and z (2) we may compute two trial functions za(t) and
*ь (0 beginning with any convenient initial conditions, say
za( 0) = zaO> zb( 0) = zb0, (12.7)
( 12.8)
hence
(12.9)
af
1978
R O T A T IN G H Y P E R B O L IC FLOW 267
( 12 . 10)
z- = z "- 2 + 2 ( | L a/’
the two initial values z 0 and Zj being calculated from the expansion
2 = г ° + й Н й ' 2 + й ' 3 - - ( 1 2 1 1 )
za = aa)zm + a^’z
( 12.12)
zb = bm z w + b m zm :)
t Неге Д/ is double the increment used in the calculation of (12.1) and (12.2).
1979
Fio. 3. (Contd.)
1980
R O T A T IN G H Y P E R B O L IC F L O W 269
Fig. 3. ( Contd.)
where a (1), a (2), b(1) and bi2) are constants. So as t-+ ° ° and /-►
z ^ a V + a V , (12.13)
where s(1\ s(2) are the truncated series for z (1), z(2) on the right of (8.22).
Hence
and
In practice a(1) and a(2) can each be determined from a sequence of values at
uniformly spaced values of f, using Shanks transforms; and similarly for b(1)
and b(2). A check on the results is that
270 M . S. L O N G U E T - H I G G I N S
Finally we solve equations (12.12) for z (1) and z (2) in terms of za and zb.
The trajectories calculated in this way are shown in Fig. 2, in the typical
case m = 0*5. The instantaneous surface profiles at successive times are also
shown, corresponding to equal increments of / (J~l is shown as a function of
t in Fig. 3 of (1)). As can be seen, the profile at time t = 0 is a rectangular
hyperbola, and as time increases the angle between the asymptotes di
minishes monotonically towards zero. A t the same time the principal axes
rotate clockwise through a finite angle.
The trajectories start with a marked component of velocity inwards into
the fluid, as the angle between the asymptotes diminishes. Later, as the
vertex angle tends to zero, the trajectory is determined more and more by
the ejection of the fluid to infinity, in the x -direction.
In Fig. 2 it has been assumed that the origin О is in a free-fall reference
frame. In a stationary reference frame, however, О moves in a parabolic
trajectory, with downwards acceleration g. T o transform the trajectories to
such a reference frame it is necessary to add to z the terms Uf +£gr2, where
U and g are (complex) constants representing a uniform velocity and the
acceleration of gravity respectively. Figure 3 shows examples, with ш = 0*3
to 2-0, in which U is taken as -1 *5 i and g as -1 , with A = 1 as before.
It will be seen that as w diminishes from 2-0 to 0-3 the total angle 8
through which the hyperbola turns gradually increases. As ш approaches
0-25, the angle у between the asymptotes “ hesitates” near 45°, and the total
angle 8 increases without limit, as in Fig. 1.
REFERENCES
1. M. S. L o n g u e t - H i g g i n s , Proc. R . Soc. A371 (1980) 453.
2. , J. Fluid Mech. 121 (1982) 403.
3. F. John, Comm uns pure appl Math. 6 (1953) 497.
4. M. S. L o n g u e t - H i g g i n s , J. Fluid Mech. 55 (1972) 529.
5. , ibid. 73 (1976) 603.
6. G. N. W a t s o n , Theory of Bessel Functionsy 2nd Ed. (Cambridge University Press,
1966).
1982
JOURNAL O F GEOPHYSICAL RESEARCH. VOL. 88. NO. С14. PAGES 9823-9831, NOVEMBER 20, 1983
Quantitative information on the strength and size distribution of whitecaps in a given wave field is
very scarce. During the MARSEN field experiments, observations of surface elevation were made with a
capacitance wire wave recorder attached to a free floating spar buoy. Automatic analysis of the records
with a differentiating circuit and counter allowed a histogram of jump-heights to be constructed corre
sponding to any preset critical rise rate R of the surface elevation. Over a certain range of R the
histogram was nearly independent of the precise value of R. This occurred usually when 0.6 < R/c0 < 1.0
where c0 was the speed of the dominant waves. Records were obtained in wind speeds ranging from I to
14 m/s. At 14 m/s the number of “jumps” indicating either steep or breaking waves was of the order of 1
every 100 wave periods. It is shown that this number is consistent with previous theoretical estimates,
and with visual observations of whitecap coverage. Because of the dispersive properties of waves, white-
capping in deep water is intermittent. Both theoretical calculation and laboratory experiments lead us to
expect steep or breaking waves to induce separated flow, with a high local input of momentum to the
wave field. It is concluded that such local events could contribute significantly to the total horizontal
stress exerted by the wind.
1. In tr o d u c tio n disc w ould leave un tu rn ed a sim ilar disc of sm aller diam eter.
B reaking waves in deep w ater are intim ately involved in a T hus the m ethod ap p aren tly was unreliable.
n um ber o f processes at the air-sea boundary, including the W e therefore resorted to the m ore direct m ethod described
horizo n tal stress exerted by wind [Banner and Melville, 1976], in section 2 below, in which the aim was to detect and m ea
th e generation and dissipation of surface waves, the vertical sure the small ju m p s o r discontinuities in surface elevation
mixing in the upper ocean, the exchange of heat and gases associated with spilling o r plunging breakers. T h e instrum ent
(enhanced by vertical mixing), and the generation o f aerosols designed for this purpose is described in section 3. After suc
by b u rstin g bubbles [ Blanchard and Woodcock, 1957]. cessful testing of a m odel in the 1980-ft (604 m) wave channel
O n the oth er hand, the frequency of occurrence of breaking at W orm ley, the p ro to ty p e was first deployed from the N ord-
waves and also their intensity have been little studied. C ertain wijk observation tow er off the D utch coast, as a p a rt o f the
descriptive properties are indeed im plicit in the Beaufort wind M A R SEN field pro g ram (see section 4). An analysis of the
scale. But the best know n quantitative inform ation seems to results will be given in sections 5 an d 6. W e also discuss the
be in term s o f w hitecap coverage [see, e.g., Monahan, 1971]. im plication of the results for w ind stress and wave generation.
D isregarding any possible subjectivity in the visual observa
tions. this still gives only an indirect m easure of the dynam ical 2. P r in c ip l e of M easurem ent
characteristics of the breakers. M oreover, the scatter in the
d a ta when plo tted against local wind speed suggests strongly Suppose th at the elevation r\ o f the sea surface is m easured
th a t the local wind speed is not the only relevant physical as a function of tim e i a t an alm ost fixed h o rizontal position
param eter. (x, y), as in Figure 1. A progressive wave passing the recorder
T h ere is clearly a need for direct, instrum ental m easure will generally show a sm o o th rate o f rise dtj/dt. But if a break
ments of b reaking waves to give quantitative inform ation of ing wave passes the recorder, we expect a sudden ju m p in
their physical properties such as the height of the “ju m p ” at surface elevation. F o r a plunging breaker (Figure la) this is
the surface, the distribution of w hitecap dim ensions, and the relatively large; for a spilling b reaker (Figure lb) it is smaller.
intensity of the accom panying turbulence. M oreover, these We rem ark th a t spilling breakers m ay start either as instabil
q u antities need to be related to the local wave field and to the ities on the forw ard face of the w ave o r as small-scale plunging
tim e history of any relevant winds. breakers near the w ave crest, which then spread dow n the
As a first step in this direction, we have (since 1974) tried forw ard face [Longuet-Higgins and Turner, 1974].
various m ethods of m easuring breakers in deep w ater. O ne Suppose then th at the o u tp u t voltage V{t) from a capaci
m ethod was to record their acoustical signature. However, the tance wire recorder is passed through a differentiating circuit,
m ain difficulty with (passive) acoustical detection is th at the as in Figure 2. W hen dV/dt cxceeds a certain threshold value
signal depends rather strongly on the distance from the source ц, say, a gate G triggers an integrating circuit I. This integrates
to receiver; thus under breaking wave conditions only the the signal dV/dt between the time t, when dV/dt rises through
m ean acoustical intensity can be easily interpreted. /j until the next instant i 2 when dV/dt falls below ц again. T he
W e also attem pted to calibrate breaking waves by their resulting “ju m p ”
ability to o verturn floating discs of different diam eters. O n the f'l av
whole, larger waves overturned larger discs. However, it was
found th a t occasionally the sam e wave that overturned a large
J = \ T idt~ K('2) “ K(t,) (1)
is then autom atically sorted into a histogram by the co u n ter
Copyright 1983 by the American Geophysical Union. C, which divides the range o f ju m p s into intervals with preset
divisions:
Paper number ЗС0481.
0 148-0227/83/003C-0481 S0S.00 0= -,J m (2)
9823
1983
4. A n a ly s is o f R e s u lts
Figure 5 shows a typical length o f record taken from run 2
on September 27, 1979. O n the trace below can be seen the
jum ps J as detected by the circuits o f Figure 2. T he scaJc of
the jum p trace is the same as th at of the record above.
T he way in which the histogram of ju m p heights J depends
Fig. 1. Illustration of (a) a plunging breaker and (b) a spilling
breaker passing a vertical wire recorder. on the critical rate of rise R can be seen, in a typical case, in
Figure 6. The vertical coordinate N, represents the total
num ber of jum ps, in a 10-min record, which exceed (i — 1)5.
T hus the upperm ost curve f = 1 in Figure 6a represents the
It is convenient to m ake the J , correspond to integer multiples total num ber of jum ps in the record; the next curve i = 2
o f a certain ju m p height Дft = 6, say. represents the total num ber of jum ps exceeding <5, and so on.
T h e resulting histogram depends clearly on the selected Clearly, as R increases, so N , decreases m onotonically. How-
values of \i and S, to an extent which we shall aim to m ini
mize.
3. T h e M e a s u r in g A p p a ra tu s
L o n g u e t - H ig g in s a n d S m ith : M e a s u re m e n t o f B r e a k in g W a v e s 9825
ever, for i Ss 2 there is a certain range of R over which each of analyzed in this way. T he given value o f R denotes the m id
the curves for N , is nearly horizontal, in this case 6.5 point, approxim ately, of the corresponding range.
m /s < R < 9.0 ш/s. O ver this range the distribution is nearly Also show n in T able 2 is the m ean wind speed U , as record
independent of R. ed by an anem om eter on the lower, a t a height of 20 m above
An exception is the curve for i = 1, which generally de m ean sea level; and the phase speed c0 of the do m in an t waves
scends m ore steeply than the others. This curve, which in in each record, th at is, the speed of a regular train of waves, of
cludes ju m ps of very small am plitude, will clearly be affected low am plitude, having the sam e num ber of zero crossings as
by any high-frequency noise, and we shall therefore disregard the actual record.
it.
T h e corresponding histogram o f ju m p heights is found by 5. I n t e r p r e t a t io n
tak in g the differences
H ow should we interpret the results of Table 2? C onsider
AN , = N l - N , . l (3) first the relation of R to c0. Figure 10 shows schem atically a
a t some particular value of R w ithin the chosen range. Figure regular wave progressing to the left. T he ap p aren t rate o f rise
6 b shows the histogram when R = 8.2 m/s, for example. of the surface at a given point is clearly
T he anaJysis of other typical records is show n in Figure 7-9.
In Figure 9 there is clearly no range of R over which the
curves N , can be considered as horizontal. In such a case we
conclude that there are no significant jum ps in the record. where с is the phase speed and Ax = cAf. But Aq/Ax = tan a,
Table 2 shows the results from all the records that were where a is the angle of inclination o f the surface to th e hori-
1985
L o n g u e t - H ig g in s a n d S m ith : M e a s u re m e n t o f B r e a k in g W a v e s 9827
TABLE 1. Times of Observation With the IOS Surface Jump Meter 18. T his show s a single occurrence o f w hat m ay have been a
From the Nordwijk Platform plunging breaker actually striking the sp a r buoy. In the re
Wind Wind m aining cases we m ay sup p o se th a t the waves were on the
Run Date Time Speed, m/s Direction point of breaking.
T h e question o f the phase o f th e ju m p s relative to th at of
2 Sept. 27 1554-1614 8.0 240° the corresponding w ave m ust also be exam ined. It seems that
3 Sept. 27 1654-1714 7.5-7.0 240°
4 Sept. 28 in every case the ju m p s occurred on the forw ard face of the
1430-1450 4.0 25°
5 Sept. 28 1525-1543 1-2 25° wave, that is ahead o f the crest, as seen on the record. Some
6 Oct. 4 0830-0850 8.4 165° allow ance m ust be m ade for th e phase shift in the response of
7 Oct. 4 0915-0935 8.0-6.4 160° the buoy. T he appendix gives a simplified theoretical analysis.
8 Oct. 18 0920-0940 13.7 300° F ro m F igure 12b it appears th a t generally the total phase-lag
9 Oct. 18 1005-1035 13.7 300°
10 Oct. 18 1038-1115 14.2 300° {Фл + Фв) *s negative. T his im plies th a t th e phase of the re
11 Oct. 19 1015-1100 11.4 220° sponse leads the phase of the surface elevation. At high fre
14 Nov. 20 1100-1140 6.6 036° quencies the total phase lag tends to zero, so high-frequency
15 Nov. 20 1140-1200 5.2-4.3 039° “jum ps” are registered w ithout appreciable time lag, b u t their
16 Nov. 23 1230-1355 10.8 252° occurrence is delayed relative to a longer wave. In oth er words
17 Nov. 24 1330-1440 6.7-7.2 220°
a “ju m p ” is further dow n the forw ard face o f the w ave than
appears on the record. H owever, the to tal phase shift \фА
+ фв\ is not large, being less th an 30° for waves o f period 7 s
zontal. H ence
o r less (corresponding to phase speeds c0 of less th an 11 m/s).
R = с ta n a (5)
6. D is c u s s io n
or
T he last colum n o f T able 2 indicates th a t un d er the given
a = arctan (R/c) (6)
range of conditions the n um ber N o f ap p aren tly breaking
N ow in a regular progressive gravity wave, the m axim um waves in a 10-min record was typically between 0 an d 8.
inclination of the surface is 30.37° [see Longuet-Higgins and Is this result reasonable in term s of present know ledge?
Fox, 1977]. H ence the m axim um value o f R/c is tan Let ro denote the p ro p o rtio n of d o m in an t waves th a t have
30.37° = 0.586. Any progressive wave in which R/c exceeds whitecaps. Table 2 indicates th a t o n O cto b er 18, for example,
this critical value m ust be breaking. T hus it is interesting th at the wave period was aro u n d 6.0 s, so th at the n um ber of waves
in T able 2 all th e ratio s R/c0 lie som ew hat above this value. in 10 min was a b o u t 100. W ith N = 3 this gives w = 3
O u r conclusion m ust be that the jum ps registered occur x 1 0 " 2, which is within the range calculated theoretically by
mainly on the forw ard face of steep, unstable waves. They are Longuet-Higgins [1969].
n o t necessarily either w hite-caps or plunging breakers. N ever Again, if the foam from a typical “ju m p ” has a persistence
theless, in the absence of a com plete know ledge o f the local time T„, then the p ro p o rtio n of the sea surface covered by
dynam ics o f steep, progressive but uniform waves, it seems foam (the visual w hitecap coverage (VW C) will be
reasonable to suppose th at these are m ainly waves th a t are
VW C = ш Tp/T (?)
ju st a b o u t to break, it n o t actually breaking.
E xam ination of all the plots of N t versus R showed that in where T is the wave period. F o r salt water, Monahan and
only one case did N 2, N 3, ••*, not become zero when R > c0 Zietlow [1969] found Tp = 3.85 s; w hence V W C = 2 x 10- 2 .
(that is, when the angle o f inclination exceeds 45°). T he excep This value lies well w ithin the range of o bservations reported
tion is in F ig u re 6, for the run from 1005 to 1015 on O ctober by Monahan [1971, Figure 2] for a w ind o f 14 m/s.
Fig. 5. (e) A typical length of record taken from nin 2 on September 27,1979. (b) The corresponding trace of jumps.
1987
18/Ю/79. 1005-1015
Fig. 6. (e) The number Ы, of jumps of magnitude h belween (i - l)£ and IS, as a function of the critical rate of rise Л,
during the 10-min interval 1005—1015 (local lime) on October 18. (fc) The corresponding histogram of jump heights;
AN,
хтггх
20/11/79 . П 0 0 -Ш 0
Дм,
ja
L o n g u e t - H ig g in s a n d S m ith : M e a s u re m e n t o f B r e a k in g W a v e s
23/11/79 1 2 5 5 -1 3 0 5
An,
5 -
In spite o f the prelim inary n ature o f the observations it m ay when the w ave is near its p o int o f m axim um steepness. Since
be interesting to discuss the possible im plications o f this result the speed o f the w ave relative to its envelope is (c — ct) or
for tw o aspects o f air-sea interaction: the horizontal stress ab o u t i c , w hitecaps are seen interm ittently, with a frequency
exerted by the w ind and the input of energy to the waves. ab o u t half th a t of the waves themselves [see Donelan el. a i,
H orizontal stress. It was show n very clearly by Banner and 1972].
M elville [1976] th a t the occurrence of w hitecaps and short T he separation of the airflow m ust therefore be in term ittent
breaking waves n ear the crest o f a steep gravity wave tended also. N o precisely ap p ro p riate calculations of the airflow are
inevitably to induce a separation of the airflow over the wave. available, and it is reasonable to com pare the onset of sep ara
M easurem ents indicated th a t the flux of horizontal m om en tion w ith som e recent calculations o f the flow started from rest
tum from the air to the w ater was thereby increased locally by over a steep but regular train o f w aves [ Longuet-Higgins ,
a factor of o rd er 50. T hough occasional separation of the 1980]. T he variation o f the d rag coefficient c D with tim e is
airflow m ay also occur when the waves are not breaking, show n in Figure 11; cD begins at t = 0 w ith a relatively high
nevertheless, it seems probable th at breaking waves induce value of 0.15 due to the form ation o f a circulating eddy behind
sep aratio n far m ore strongly and consistently. each crest. O nce the eddy is form ed, however, th e m om entum
T w o questions then arise: H ow frequently are breaking of the airflow changes less rapidly, an d the d rag coefficient
waves observed in the ocean, and w hat p ro p o rtio n of the total falls drastically to a level w hose m ean is far less, by perhaps
w ind stress can the local patches of breaking be expected to tw o orders o f m agnitude. T he high d rag is thus a m om entary
acco u n t for? phenom enon, lasting for a tim e TD o f o rd er L/(U — c), where L
W ith regard to the first question we note first th a t wave is the w avelength and U the w ind speed. Since the wave
b reaking is an essentially interm ittent phenom enon, especially period T is equal to L/c, it follows th at
in deep w ater. T his is due to the fact th a t surface waves are
T J T Ф c/(U - с ) (8)
b o th dispersive and irregular; high waves occur in groups,
w ith individual waves traveling through the group at a speed с which, in the case of the d ata o f Table 2, is of o rd er 1 a t least.
w hich is greater than the group velocity cr T hus a wave will T he contribution of steep waves to the to tal m ean d rag coef
start a t the rear of a group, grow to its m axim um steepness as ficient is then of order
it passes thro u g h the group, and die aw ay as it reaches the
front of the group. A w hitecap will be seen only m om entarily сD' = wCmiU ~ c)2/c2 (9)
where c D0 denotes the initial d rag coefficient for th e high
waves. Conservatively, we m ay take, from F igure 11, 4=
TABLE 1 Analysis of Data 0.075. In table 2, for the d ata of O cto b er 18, (U - c)/c is of
u. order 0.3, giving cD' = 0.6 x 1 0 " 3. T his is sm aller th an the
Co R.
Date Time m/s m/s m/s N custom ary coefficient cD 4» 1.5 x 10“ \ b u t nevertheless o f the
sam e order of m agnitude.
Oct. 18. 1979 0922-0932 13.7 7.6 6.5 3
Oct. 18, 1979 0935-0945 13.7 9.2 8.0 1
Oct. 18. 1979 1005-1015 13.7 9.9 8.2 6
Oct. 18, 1979 1035-1045 14.2 10.4 8.5 7
Oct. 18. 1979 1055-1105 14.2 10.6 8.5 5
Oct. 19. 1979 1020-1030 14.4 9.5 _ 0
Oct. 19, 1979 1051-1101 14.4 7.9 6.5 5
Nov. 20, 1979 1100-1110 6.6 6.7 5.0 2
Nov. 20, 1979 1130-1140 5.2 5.2 _ 0
Nov. 23. 1979 1230-1240 10.8* 8.4 5.5 1
Nov. 23, 1979 1255-1305 10.8* 8.4 5.5 1
Nov. 23. 1979 1320-1330 10.8* 8.4 _ 0
Nov. 24. 1979 1335-1345 6.7 6.6 5.0 2
Nov. 24, 1979 1412-1422 6.9 6.7 5.0 2
TTTmT77rfrf1'
•Heavy rain. Fig, 10. Interpretation of the rate of rise R for a progressive wave.
1989
Hence
a = 2 n/TD = 0.598 rad/s
K/a = - In ~ - 0.63
n D
Theoretically, the vertical displacem ent у o f the buoy from r j + K t j + w024 = (A12)
its equilibrium position m ay be assum ed to be governed by
the equation
y+C\y\y + wo 2y = 0 (A l)
J K y 2 dt m I C|yiJ dt (A2)
у+ K y + o»0гу - 0 (A3)
W hen К < 2w0, this has solutions of the form
w here
a = (a>0J-i/Cl),/J (A5)
an d A an d i are constant. T he “period" TD of the dam ped
oscillation is
TD = 2n/a (A6)
L o n g u e t - H ig g in s a n d S m i t h : M e a s u r e m e n t o f B r e a k in g W aves 9831
H ence th e response to the sinusoidal oscillation shifts larger th an 30° occur only for w ave periods T greater
th an ab o u t 7 s.
Y = ае‘ш (A 13)
w here a an d со are real constants, is given by Acknowledgments. The authors are much indebted «о B. McCart
ney for assistance in the design of the elecironic circuits used in the
П— (A 14) apparatus, to R. Spanhoff and his colleagues at the Rijkswaterstaat
for collaboration and logistic support on the Nordwijk tower, and to
provided О. H. Shemdin and his group at the Jet Propulsion Laboratory,
California Institute of Technology, for allowing us to use channels on
( - c o l + iKa> + (o02)be~‘* = [ - w 2 + iK (l - p)w ]a (A15) their magnetic tape recorder. Their work was carried out under con
tract with the National Aeronautics and Space Administration. This
manuscript was begun in England and completed during a visit by
{oj02 — со2) + iK io
one of the authors (M.S.L-H.) to the Department of Engineering
{а/Ь)е‘* (A 16) Sciences at the University of Florida, Gainesville, Florida, during
- t o 2 + iK (t February 1981. Comments and discussions were provided by mem
bers of his class.
and th e p hase lag ф will be given by
B y M. S. L o n g u e t -H iggins , F.R.S.
Department of Applied Mathematics and Theoretical Physics, University of
Cambridge, Silver Street, Cambridge CBS 9EW , U.K.
and Institute of Oceanographic Sciences, Wormley, Godalming,
Surrey GUS 5UB, U.K.
1. I n tr od uc tio n
Since the pioneering work o f Whitham ( 1967) and Benjamin & Feir ( 1967) it has
been known that gravity waves on water o f infinite depth are unstable, at least
to certain subharmonic perturbations. Precise calculations o f the normal modes
by the present author ( 1978 a) and by McLean ( 1982) have shown also the existence
o f some unstable perturbations with length scales shorter than the basic
wavelength.
While many o f the instabilities are three-dimensional, some important and
interesting examples occur even in two dimensions (one horizontal and one
vertical). Thus it was shown in Longuet-Higgins ( 1978a; hereinafter referred to
as I ) that some superharmonic disturbances, with a dominant wavelength one half
that o f the original wave (i.e. having wavenumber n = 2 ) tend to become unstable
when the steepness ak o f the unperturbed wave exceeds about 0.43. Although the
calculations could not be carried accurately beyond about ak = 0.42, nevertheless
they were consistent with the suggestion (made on physical grounds) that a
transition to instability occurs at the wave steepness ak for which the phase speed
с is a maximum. From independent calculations (Longuet-Higgins & Fox 1978)
this was known to be at about ak = 0.436 (less than the maximum steepness
ak = 0.4434).
[ 269 ]
1992
f = f (? U ). <2 J >
the two boundary conditions expressing that the pressure p is constant and that
a particle at the surface remains at the surface become respectively
-(УфУ1+У*Х1)+ЯУ(Уф + У \ )+ \ = ® <2'2*
3. T he basic wave
in which the coefficients an are all real, and units have been chosen so that the
1994
Y = £ * am cos mQ (3.2)
TO-0
where в = ф/с (3.3)
and an asterisk means th at whenever a 0 occurs in the sum it is to be replaced by
Ja0. F urther oo
—cYj, = £ bn созпв,
n-o
00 (3.4)
- с У ф = £ bn sin пв,
a \ ^ 0 + ( а 0 + а 2 ) ^ 1 + (а 1 + а з ) ^ 2 + (а 2 + а 4 ) ^ 3 + - - - =
(3.11)
а 2 ^ 0 + ( а 1+ а з ) ^ 1 + ( а 0 + а 4 )/ 2 + ( а 1+ а 5)^ 3 + -* - =
th at is to say B x F = G, (3.14)
bo bx b2 6 3
0 b0 bx b2
where B = (3.15)
0 0 60 b,
4. T h e n o r m a l m o d e s ; l i m i t a s <r->0
To solve the perturbation equations (2.8) and (2.9) it was noted in I that (f, rj)
being conjugate functions, tending to 0 as i/r/c-^ао, may be expanded in the form
0?-i£> = E + (4.1)
n- 0
and th a t in general ^
/ = S ( r n + ii„ )e ‘" ^ , (4.2)
71—0
where a n, fin>y n and $n are real constants. Substitution of these expressions into
(2 .8 ) and (2 .9 ) and equating the coefficients of cosn# (n = 0 , 1 , 2 ,...) and sinn#
(n = 1 , 2 , 3 ,...) yields a system of equations for the constants л п,/Зп, y n, Sn and the
radian frequency or, which can be solved numerically by successive truncations.
However, we are here interested chiefly in the limiting case when <r-*0. From
(2 .9 ) we see immediately th at in the limit
Л - 0, / —Уо- (4.3)
Equation (2 .8 ) then reduces to
Pv + QVf + ^ + s V 0 = 0, (4.4)
to be satisfied when \jr = 0.
1996
274 M. S. Longuet-Higgins
We rem ark th a t this equation can be derived directly from the steady boundary
condition 2y{yl + y \ ) = - 1 on xjr = F (4.5)
(compare (2.5)) by writing
y = Y + e T j } F — ey0 (4.6)
and then considering the coefficient of e.
5. N ormal modes (c o n t in u e d )
To proceed further with (2.8) we must evaluate Q, R and S. From (2.10) and
(3.4) we have oo oo
—cQ = £ * E a n 6n[cos(m*f n )e + cos(m —n)6], (5.1)
n —0 TO-0
Ql + R l = 2 E ttf-m^m+ 2 E a m fy+ж —0
m-0 m-o
by (3.12). Hence for I = 0 ,1 ,2 ,3 ,... we have
On substituting these expressions into the boundary condition (4.4) and equating
to zero the coefficients of J, cos#, cos 26, ... and sin#, sin 26, ... we obtain a linear
system of equations for
a = (aQ, a l ta 2, ...),]
(5.15)
= [fiv fit ’ ■••) J
and y0, which may be written in the form
' M 0 S
i
x ( oe ; p ; To )
т _ (5.16)
0 N i 01
where
(Po + ^o) (Л + ^i) (^2 + П) (P3 + P3 )
(Pi + Pi) (P0+ P2+ Q0) (Pi + P>) (P* + P<1)
M = (P2+ P2) ( P ^ P t + Q j (Pq+ Pa+ 2Qq) (/> + />) (5.17)
(P3 + P3) (P2 + ^ + 0 2) (P1 + P5 + 2«1) (P0 + Pe + 3Q0)
and О denotes the zero square matrix, 0 a zero column vector. Finally
S = ( S 0, S l ,S 2,...)T (5.19)
6 . C o n d i t i o n s w h e n <t = 0
We shall make use of the following lem m as:
M = BxC, (6.1)
N = В x £>, (6.2)
in which В denotes the m atrix (3.15) and C ,D are the symmetric matrices:
2 ((1,4*0.,)
(aj + aJ (2a24-2a2) (Зл34-Зй3)
(1 4- a04- 2a2) (2^14" 3a3) (3a24-4a4)
C = (2a24- 2a2) (2a14-3a3) (1 4- 2a04-4a4) (З ^ -Ь б ^ ) (6.3)
(3a34-3a3) (3a24-4a4) (3®! 4* 5as) (14- 3a04-6ae)
and
(l4 -a 0- 2 a 2) (2«! —3a3) (3 a 2 —4 a 4)
(2aj —3a3) (1 + 2 a 0—4a4) ( 3 a j - 5 a 5)
£> = (6.4)
(3a2—4a4) (3a2—5a6) (1 4 - 3 a 0 — 6 a e )
Moreover, M x b T = S, (6.5)
N x b'T = 0, (6.6)
where S is given by (5.19) and
* “ (боА Л . « 0 , ( 6 -7 >
and N x p = 0 ( 6 . 11)
This represents a simple phase-shift of the original wave through the horizontal
distance ec.
Now (6.11) implies also th at the determinant of N vanishes. Hence the first row
of N is linearly dependent on the others. The corresponding equation is thus
redundant and may be replaced by a condition on the phase, for example that
A = 0. (6.16)
The modified system then becomes
N' x ft'1 = 0, (6.17)
where fi' = [/32,J33, ...) and N' is the matrix derived from N by omitting the first
row and column, th at is
(Po-P* + 2Qo) « -/» -
( /> - /> + 2 0 0 (/>0- ^ e + 3<20) (6.18)
Now, as in the proof of (6.1) given in the Appendix, it may be shown that
N' = B x D \ (6.20)
where D ' is the matrix derived from D by omitting the first row and column, that
is
(1 + 2a0—4<z4) (3oj 5a5)
D' = (3ах- Ь а ъ) (1 + 3a0 —6 ae) ( 6 . 21 )
B ut it has already been shown numerically th at | D' \ does not vanish anywhere
in the range of uniform waves (see Longuet-Higgins 1 9 8 4 6 ), and particularly not
2000
278 M. S. Longuet-Higgins
when E = E mgLX. Hence }ЛГ | does not vanish anywhere in the range, and there are
therefore no antisymm etric normal modes a t zero frequency apart from a pure
phase-shift.
Since any perturbation to the original wave form may be expressed as the sum
of a symmetric part and antisymmetric part, this shows th a t there are no other
antisymmetric normal modes of any form a t zero frequency, except a t points where
the phase speed has a stationary value, i.e. when dc = 0 .
The form of the normal mode, a t points close to a point of stationary phase speed,
is presumably given by
a = [a0, a lfa2, ...] = a(c) + e(Aa + /a/f) + 0(e2), (6 .2 2 )
where a <c) denotes the values of the Fourier coefficients a t the critical wave
amplitude, a and p correspond to solutions of (6.10) and (6.11), and A, fi are
constants. To determine the ratio A\ц and the corresponding (non-zero) value of
the radian frequency cr the full equations (2.8) and (2.9) may be carried to higher
order in e.
7. D iscu ssio n
On multiplying the ith row of В (see equation (3.15)) by the jth column of (A 1)
and using the definition of Pt in (3.8), we see th at the terms in 6n alone yield the
2001
inner product (/^_д + Рщ). On the other hand, from equations (3.12) and (5.3)
the terms in b{ and yield 0 if i < j , and (j —i)Q i 4 if i > j . This proves (6.1).
To prove (6.2) let us define
-(бо —60) (0—b1+ a1) ( 0 - 6 2+ 2a2) (0 —63+ 3a3)
(b r-b j (b0~ b 2+ a0) (0 —63+ 2a,) (0 —64+ 3a2)
(b2- b 2) ( ^ - Ь з + aJ (60- 6 4 + 2a0) (0 —65 + 3 0 ^
—
(6 3 )
6 3 (62—64+ a2) (6i —66+ 2ai) (b0- b 6+ 3a0)
(A 2)
Then it is clear th at
D[ = D (A 3)
(where a prime denotes the matrix derived from a given matrix by omitting the
first row and first column). As before, we now have
B x D l —N t, (A 4)
where
(Po-Po) (Рг-Рг + 0) (P2- P 2 + 0) (Рз-^з + 0)
(Pi~ Pi) (Po-P* + Qo) ( P i-P * + 0) (Рг-Р* + Ъ)
(Рг ~Рг) (Pi - Р з + Qr) (Po-P* + 2Qo) (Pi-Pb + 0)
(P* ~~P3 ) (Pi~~Pi + Ф2) (P i-P b + W i) ( / > - / > + з д 0)
(A 5)
B ut in D 1 the elements of the first row all vanish. So from (A 4) we have
B' x D\ = N[ (A 6)
and since В' = В (see equation (3.15)) this is (6.2).
To prove (6.6) we note that since the elements in the first row of N t all vanish,
it follows from (A 4) that
b' x D = 0T (A 7)
and on taking the transpose matrix of each side of this equation we obtain
D x b ' T = 0, (A 8)
D being symmetric. We now multiply on the left by B, using the commutative
law for matrices, to obtain
( B x D ) x b'T = 0 (A 9)
and (6.6) follows, by equation (6.2).
Lastly, to prove (6.5) note first that from the definition of St in (5.13) we have
^ = (^ + ^o)^o + № + ^ + 0)fti + № + P2 + 0)62+ . . . ,
Si = (JP1+ / >i)60+ (/i + ^ + ^o)6i + (A + ^3 + 0)62+ . . . 1 (A 10)
^3= ( P 2+ P2) 604-(P1-fP3+ <?1)61+ (P0+ P4 + 2 g 0)62 + . . . >
280 M. S. Longuet-Higgins
From (5.17) and (6.6) it is clear th a t St is the inner product of b with the (I —l)th
row of M. This proves the lemma.
This paper was begun while the author was visiting Cal. Tech. J e t Propulsion
Laboratory, Pasadena. These and other results were presented at the Workshop
on Surface Gravity Waves a t U.C. Santa Barbara on 31 May-1 June 1984; also
a t a meeting a t the University of Bristol on 3 July 1984. To Dr M. Chahine, Head
of the Earth Sciences Division of J.P .L ., and to Professor M. Tulin, U.C.S.B., the
author expresses thanks for their hospitality and assistance.
I am indebted to Dr P. A. Saffman and Dr M. Tanaka for interesting corre
spondence and discussion.
R eferences
Benjamin, Т. B. & Feir 1967 The disintegration of wave trains on deep water. J. Fluid Mech.
27, 417-430.
Longuet-Higgins, M. S. 1975 Integral properties o f periodic gravity waves o f finite amplitude.
Proc. R. Soc. Lond. A 342, 157-174.
Longuet-Higgins, M. S. 19 7 8 a The instabilities of gravity waves of finite amplitude in deep
water. I. Superharmonics. Proc. R. Soc. Lond. A 360, 471-488.
Longuet-Higgins, M. S. 19786 Some new relations between Stokes’s coefficients in the theory
of gravity waves. J. Inst. Maths. Applies. 22, 261-273.
Longuet-Higgins, M. S. 198 4 a New integral relations for gravity waves of finite amplitude.
J. Fluid Mech. (In the press.)
Longuet-Higgins, M. S. 19846 Bifurcation in gravity waves. J. Fluid Mech. (In the press.)
Longuet-Higgins, M.S. & Fox, M. J. H. 1978 Theory of the alm ost-highest wave. Part 2.
Matching and analytic extension. J. Fluid Mech. 85, 769-786.
McLean, J. W. 1982 Instabilities of finite-amplitude water waves. J. Fluid Mech. 114,315-330.
Tanaka, M. 1983 The stability of steep gravity waves. J. phys. Soc. Japan 52, 3047-3055.
W hitham, G. B. 1967 Variational methods and applications to water waves. Proc. R. Soc. Lond.
A 299, 6-25.
2003
1. I n tro du ctio n
Steep, irrotational gravity waves on deep water display in simple form some
striking nonlinear wave phenomena. These include the presence of local maxima
in the phase speed с or the energy density E (Longuet-Higgins 197s); the
bifurcation of regular, uniform waves into steady wavetrains of non-uniform
amplitude (Chen & Saffman 1980; Saffman 1980; Longuet-Higgins 1985 a ) ; and the
existence of subharmonic instabilities (Benjamin 1967; Longuet-Higgins 1978 a, 6;
McLean et al. 1981; McLean 1982). All these phenomena are already present in the
analytically simpler two-dimensional waves, as well as in more general three-
dimensional wave motions.
The first accurate calculations of the normal modes of perturbation for gravity
waves of finite amplitude were made by Longuet-Higgins (19780,6) and were
confirmed by Longuet-Higgins & Cokelet (1978). These revealed the existence of
two types of subharmonic instability. The type I, or Benjamin-Feir, instabilities
were present, generally in the range 0 < ak < 0 .41 , though some modes returned
and where
dfl = {$da0,d a v da2,...)
(3.3)
dc 2 = (dc2, 0 , 0 ,...)T.
7*»
2006
170 M. S. Longuet-Higgins
In class m /l waves, the regular series, in which the wavelength is 2n /m , is
characterized by
= 1 (34)
“ mn+i = 0. J =
where an is the corresponding Fourier coefficient for class 1 waves. In this series
we have damn = ( \ / m ) d a n and damn47 = 0. On the bifurcated branch, however,
we seek solutions such th at
< 4 ,,+ ,# 0, 1 = 1 ,2 ,... (m—1). (3.5)
Now when m is odd (and greater than 1) it will be seen th a t the corresponding set
of equations (3.1) breaks up into \(m —1) subsets, of the form
C(m, I) da(m, I) = 0 (3.6)
where C(m, I) is a submatrix of C, containing only the rows and columns numbered
I, (m ± l ), (2m ± l ) y (3m ± l), etc., and similarly for da. For solutions of (3.6) to exist,
the determinant of C(m, I) must vanish. And on substituting for the am n from (3.4)
and suppressing the primes we obtain
d- - ° - ,s ', )
1
2«, За,
[ 1 + s a '] ( , + s ) “ > i K b
° i 2°. H b i 3° > н ь
( l + s b
4а4
[ 1 + ( 1 + s ) “ -] i K b
3«, 4a4
K b ; [ , + K b ]
(3.8)
Equation (7.4) of I is the special case when m = 3 and I = 1.
When m is even we obtain (Jra—1) equations similar to (3.7) and one equation,
corresponding to I = £m, which is the same as th at for m = 2 and I = 1, and is in
fact equation (6.5) of 1.
In the general case, Dm t is clearly a function of l/m only. It will also be seen
th at Dm t remains invariant if I is replaced by [m —I) so l/m is replaced by (1 —l/m).
So the curve of ak against l/m is symmetric about l/m = J.
To find the first roots of (3.7) we calculated the ratios D m JD"m t (where D'
denotes the principal minor of D) as a function of the wave amplitude
a = а х+ a3-f a5+ — (3-9)
The first zeros of D m t/D"m j are shown in table 1 for some representative low values
of m/l. These are plotted against l/m in figure 1. I t can be seen th at these lie on
2007
a smooth curve, almost a parabola with vertex at l/m = The very small difference
between m /l = 2 and m /l = 3, which was found by Chen & Saffman ( 1 9 8 0 ), is also
accounted for by the symmetry about l/m = £.
In the limit when m-> 00 and I remains fixed, Dm t takes the limiting form
ьэ
3az 4a4
—
2a, j (l+«o)
w
а
«1
2аг 2a x i 3a3 (1+2 a0) 4a4 3a,
2a, 3«, j 2a j 4a, (1 + 2 a 0) 5a6
3as 3at , 4a 4 3a! 5a5 (1 + 3a0)
1
1
The first zero of D ^ /D ^ was also calculated and is shown in table 1 and
figure 1. This clearly lies on the same smooth curve.
We note that the limiting value itself cannot be expected to correspond to a
bifurcation, because any train of waves having infinitely slow modulation would
be indistinguishable from a uniform wavetrain.
2008
172 M. S. Longuet-Higgins
4. N orm al m o d e s : g en er a l v alues of th e fr e q u e n c y
y - % = E (an + i/?n)e*"<*+W c,
n-0
(4.3)
/= S (y„ + ian)eIn(*+ W c ,
Thus
(l, Ш ( во I ) ( 3
XAn V2 = Bu Vt A
(4.7)
XA v x = B22v 2,]
and on eliminating v 2 and v x in turn, we have
(AM21—B22 A 12l B n ) v x = 0 ,|
(4.8)
(A2A 12—B l l A 2l1 B 22) v 2 = 0 /
The forms of the submatrices A 12, A 2l, Bu and B22 are given in Appendix В
5. S u p e r h a r m o n i c s : t h e m o d e n — 2
To check the method, and to verify the calculations of Tanaka ( 1 9 8 3 ) on the
stability of the lowest superharmonic mode, the above method was first applied
to superharmonicsj in the range of wave steepness 0 < ak < 0.43.
The Fourier coefficients an were computed directly by the method of paper I,
up to order n = 600. These agreed well with an asymptotic expression for a n, valid
at large wave steepnesses, which has been derived in Longuet-Higgins (1 9 8 5 b). The
eigenfrequencies, or rather their squared values, were then computed by the
method of §4 , a t selected values of the wave steepness ak , and with values of N
as great as 72 when ak ^ 0.425, and 144 when ak = 0.43. The results for the two
lowest eigenfrequencies and <r2 are shown in table 2a. In the same table are
shown the first differences AN between successive convergents, and the ratios r of
successive values of A^. As this ratio is nearly constant in each case, we may
estimate the final value by Richardson’s extrapolation formula
/00 = / t f + A W ( 1 ” r )> (6 -1)
174 M. S. Longuet-Higgins
1 9 7 8 a) and should be precisely zero. The estimated value of erf therefore gives
some indication of the accuracy of the computation. This suggests th a t the last
decimal place of the calculated values of v \ is uncertain. Nevertheless when
plotted in figure 2 it can be seen th at there is close agreement with T anaka’s ( 1 9 8 3 )
curve. In particular there is no doubt th at <r\ changes sign between ak = 0.425
and 0.430, very close to the energy maximum a t ak —0.4292.
F ig u re 2. Square of the radian frequency, <r£, for the lowest non-trivial superharmonic mode;
+ , Longuet-Higgins ( 1978 ) ; x , Tanaka ( 19 73 ); ©, present paper.
6. S ubh a rm o n ic modes
F ig u r e 3 . The radian frequency Re (cr) aa a function of the wave steepness ak , for the simplest
subharmonic mode, n = (£,|).
The simplest mode, n = (£, §), is shown in figure 3. I t will be seen th at the curve
has (by symmetry) a vertical tangent at D where cr = 0,ak = 0.40496 (see table 1).
Figure 4 shows the two next lowest modes, n = (},J) and (J, J), which are in fact
all roots of one analytic equation (cf. §§4,5). These pass through zero at a steepness
ak = 0.40433 (see table 1). At the point of zero frequency neither curve has a
vertical tangent, but each passes continuously into the reflection of the other in
the axis cr = 0. At higher values of ak the combined curve does have a vertical
tangent, at which a new instability appears. This can be appropriately labelled by
the combined symbol 4л. = (1,3,5,7) or 8n = (2,4,10,14).
Figure 5 shows the intermediate modes 8n = (/,16 —/), where / is odd. In each
2012
F ig u r e 4 . The radian frequency Re(<r) as a function of the wave steepness ak , for the
subharmonic modes 4n = (1,7) and (3,5).
2013
F ig u r e 5. The radian frequency Re(or) as a function of the wave steepness ak, for the modes
8n = (1,15) (3,13), (5,1») and (7,9).
2014
(2.14) 178
<=>
Re(<r)
M. S. Longuet-Higgins
Fiqueb 6. The family of frequency curvee 8n ~ (1,16—1), including the euperharmonic 8n = (0,16).
(2,14)
Re (a)
Bifurcation and instability in gravity waves
ak
F ig u r e 8. Regions o f stab ility for modes of types I and II.
201 7
F ig u r e 9. Rates of growth Im(o*) for modes of types I and II, over the range
0.34 < ak < 0.43.
To mark the boundary between unstable and neutrally stable modes of the
present type, we may draw a smooth curve through all the critical points where
the frequency curves have a vertical tangent. The result is figure 8 .
On the left of the diagram the stability boundary meets the axis cr = 0 in the
point A where ak = 0.347. This has been shown to be the limiting wave steepness
for slow modulations, as given by Whitham’s averaged Lagrangian theory
(Lighthill 1 9 6 7 ; Longuet-Higgins 1 9 6 8 b). Here we note th at the point A is quite
distinct from the limiting point В discussed earlier in §3. This may be seen simply
2018
7. D isc u ssio n
The subharmonic instabilities th a t we have studied are of two kinds. Those with
dominant wavenumbers n = (v,2 —v) may be called type 1 , and have relatively
low growth rates. They have been identified as Benjamin-Feir instabilities (see
Longuet-Higgins 1 9 7 8 6 ). They extend throughout the range of ak betweenO and E.
The point E is not associated with a bifurcation.
Secondly, we have type 2 instabilities, with dominant wavenumbers
n = (v, \ —v, 1 + y, 2 —v), and relatively high growth rates, see figure 9. In two
dimensions, these instabilities extend throughout the range from В to E. The case
v = \ is associated with a bifurcation at B.
There are presumably higher types of instability in ranges of ak beyond the
point E.
Tanaka ( 1 9 8 5 ) has shown numerically that the next superharmonic n = 3
becomes unstable at ak = 0.442, close to the next energy extremum d E = 0.
Presumably the frequencies of the subharmonic modes could be calculated
similarly. However, for an understanding of these and other results at still higher
values of ak it may be better to use an asymptotic analysis based on the theory
of the almost-highest wave (Longuet-Higgins & Fox 1 9 7 8 ).
Some authors have suggested th at the type II instabilities, and other instabilities
of higher order, are essentially three-dimensional, being finite-amplitude examples
of the higher order resonances noted by Zacharov ( 1 9 6 8 ). There is, however,
nothing three-dimensional about the phenomena discussed here. This contrasts
with the situation in shear flows, for example, where stretching of vortex lines
produces some essentially three-dimensional effects which would be absent in two
dimensions. Yet irrotational waves in only two dimensions still exhibit a wealth
of interesting nonlinear properties. In this paper we have taken the alternative view
th at such properties will be most easily understood, both numerically and
analytically, by first concentrating attention on the simplest possible system,
namely the two-dimensional case. From recent papers (Saffman 1 9 8 0 , 1985*
Tanaka 1 9 8 3 , 1 9 8 5 ; Longuet-Higgins 1 9 8 5 a) there is evidence that this approach
has already proved fruitful.
Since this investigation was begun, the author has learned of a report by Branger
et al. ( 1 9 8 4 ) in which some of the above results were obtained, but apparently with
less accuracy. To accelerate convergence, these authors used the ‘e-algorithm’ of
2019
and where
Рп — 2 bn+jb} ,
J- о
oo (A 2 )
Qn = 2 Яп+jbj.
i- 0
An exception to the rule could occur if the limiting form of the instability is a pure
phase-shift.
We shall now show that for a subharmonic normal mode of class m /l , the
condition that | M | = 0 is equivalent to the condition that D m t = 0.
I t was proved in II that
M = J3 x C , (A3)
where
Л Ьг Ь,
0 К 6. Ь2
B= 0 0 К Ьг (A 4)
0 0 0 ь,
and С is the matrix (3.2). In a subharmonic normal mode of class m the only non-zero
components of the unperturbed wave are those of order n = j m , where
j = 0 ,1 ,2 ...... Hence also the only non-zero Pn and Qn are those of order jm.
Now in a subharmonic of class m /l (that is, one having a wavelength m times
2020
the unperturbed wavelength), the only harmonic components arc those of order
i j m ± l ) (see Longuet-Higgins 1 9 7 8 6 , §4). Hence the condition for the frequency
of such a perturbation to pass through zero is th at the matrix M {m , /) be singular,
where M(m, I) denotes the matrix constructed out of the rows and columns of M
numbered (j m ± l ). But all the elements of В vanish except those numbered sm,
where 5 is an integer. Hence we find, corresponding to (4.3), th at
M(m, I) = B{m) C(m, I), (A 5)
where
0 bm 0 ^2m
0 b0 0 bm 0
0 0 fro 0 bm (A 6 )
1=
0 0 0 bo 0
0 0 0 0 bo
A p p e n d i x B. F o r m s o f t h e m a t r i c e s A i} a n d B fj
We give here, for reference, the forms of the submatrices A l2, A 2l, B n and B22
introduced in §4. These will be clear from the case N = 2. Thus
(60+ 60) ( K +6,) (62+ 62) "
bi (60+ 62) №,+6a) 0
<5
<2
L -ь2
1
1
1—
-b 9 (b0- b i )
( + 0)
6 0 6 ( . + ,)
6 6 ( + 2)
6 2 6 (Л+Р.) [Pt + P t )~
Ьг ( + 2)
6 0 6 (6 1 + ) 6 3 (/u n ) (Л + /У
ОI —
b2 b> ( + 4)
6 0 6
(Л + /у (П+Л) (B 2)
*. i - b 0+ b 2) (- 6 1 + ) 6 3
\ ()
. 6 2 b3 (- 6 0 + 4)6 -
2021
Bn =
(/>,+/>) ( П + ^ + ЗУо) I № + «,) (tf0+ « 4>
I Co 0
0
. i о *0. -
! о 0 ■
0 1 -в о 0 1
B z2 = ! о -2 ^ o
w h e re Pn a n d Qn a r e g i v e n b y e q u a t i o n s (4 .2 ), a n d
T able 1. P o in t s o f b if u r c a t io n
m /l l/m ak
2 .5 .404961
в .375 .404813
3
3 .3333 .404687
4 .25 .404335
6 .1666 .403823
8 .125 .403501
oo .402233
T a b l e 2. (a ) C o n v e r g e n t s t o <
t\ and (t\
N г? Л г <т\ A
ak = 0.41
40 .000657 .078297
339 897
48 .000318 .49 .077400 .36
165 325
56 .000153 .48 .077075 .34
80 112
64 .000073 .48 .076963 .32
38 36
72 .000035 .076927
ale = 0.42
40 .002949 .049594
1043 2412 .45
48 .001906 .67 .047 182
694 1096
56 .001212 .65 .046086 .43
448 475
64 .000764 .63 .045611 .39
284 186
72 .000480 .045425
2022
T a b l e 2 . ( a ) ( cont.)
N Л r
ak = 0.425
40 .006803 .032706
48 .005327 1476 4190
.88 .028516 .55
56 .004032 1295 2324
1054 .81 .026192 .55
64 .002978 .77 1279
816 .024913 .53
72 .002162 681
.024232
ak = 0.43
80 .004862 -.0 1 6 2 7 4
96 1349 2074
.003513 .68 -.0 1 4 2 0 0 .94
112 911 1957
.002601 .74 -.0 1 2 2 4 3 .84
128 675 1639
.001926 .77 -.0 1 0 6 0 4 .79
144 528 1288
.001408 -.0 0 9 3 1 6
(6 ) E s t i m a t e s or o-f, cr\ a n d a 2
ak Nr
.41 300 .00000 .07691 .2773
.42 300 .0000 .04531 .2129
.425 300 .0006 .0235 .153
.43 600 .0003 -.0 0 4 5 .067 i
T able 3. F r e q u e n c i e s o f s u b h a r m o n i c m o d e s o f t y p e 8 n — (1, 1 6 — /) i n t h e
RANGE OF WAVE STEEPNESSES ak FROM 0 .4 TO 0 .4 2 5
ak °Y <r< 07 *7 <*i <*r
I= 1
.400 .05028 .0 .00508 .0 .22420 .0 .24092 .0
.4025 .05040 .0 .00156 .0 .22772 .00506 .227 72 .00506
.405 .05051 .0 .00253 .0 .22236 .00890
.00890 .22236
.40625 .05057 .0 .00485 .0 .21944 .00923 .21944 .00923
.4075 .05063 .0 .00741 .0 .31632 .21632 .00856
.00856
.410 .05077 .0 .01344 .0 .2139 .0 .20479 .0
.415 .05107 .0 .03236 .0 .2179 .0
.0 .16350
.420 .05147 .0 .086 .037 .2185 .0 .037
.086
.425 .0518 .0 .069 .095 .2187 .095
.0 .069
1= 2
.400 .09892 .0 .01459 .0 .23044 .0 .17967 .0
.4025 .09911 .0 .00681 .0 .23126 .0
.0 .16718
.405 .09931 .0 .00278 .0 .231 85 .0
.0 .15281
.40625 .09941 .0 .00860 .0 .23209 .0
.0 .14454
.4075 .09951 .0 .01545 .0 .23230 .0 .13516 .0
.410 .09973 .0 .03524 .0 .23265 .0 .11008 .0
.415 .10022 .0 .0667 .0595 .23315 .0 .0667 .0595
.420 .1008 .0 .0592 .0983 .2335 .0983
.0 .0592
.425 .1017 .0 .0493 .134 .2339 .134
.0 .0492
Z= 3
.400 .14426 .0 .03130 .0 .21523 .0
.0 .11388
.4025 .14450 .0 .01787 .0 .21560 .0
.0 .09831
.405 .14475 .0 .00197 .0 .21592 .0
.0 .07622
.40625 .14487 .0 .021 10 .0 .21607 .0 .0
.05683
.4075 .14499 .0 .03791 .02723 .21621 .0 .02723
.03791
.410 .14525 .0 .03664 .05488 .21647 .0 .03664 .05488
.415 .14580 .0 .03374 .09033 .21695 .0 .03374 .09033
.420 .14648 .0 .03016 .1206 .21743 .0 .03016 .1206
.425 .14744 .0 .0254 .152 .2180 .0 .0254 .152
2023
R e fe r e n c e s
1. I n t r o d u c t i o n
In almost all sea states under the action of wind there is a rather broad spectrum
of wavelengths, ranging from long gravity waves of period several seconds down
to short gravity waves and capillary waves, which are highly responsive to the
local wind speed. The latter are known to be modulated by the longer gravity
waves on which they ride (Evans & Shemdin 1980) and it has been shown by
numerous laboratory and field studies (Keller & Wright 1975; Wright et al. 1980;
Plant & Keller 1983; Hoogeboom 1985) th at these modulations are responsible for
backscattering of centime trie radar waves from the ocean surface, and hence are
the principal factor in imaging the surface by X-band and L-band radars.
Nevertheless, the problem of how to account for the observed distribution of
short-wave energy with regard to the phase of the longer waves has remained
unsolved.
I t has long been known theoretically th at straining of the short waves by the
orbital motion of the long waves, together with work done against the short-wave
radiation stresses, tends to produce shortening and steepening of the shorter waves
near the long-wave crests (Longuet-Higgins & Stewart i 9 6 0 ). Recently, some much
more accurate calculations of this effect according to the principle of action
conservation (Longuet-Higgins 1987) have shown the importance of including full
nonlinearity of the longer waves, in some circumstances. A linear model for
ehort-8 urface waves, which includes both growth and dissipation, has been
proposed by Smith (1986). Nevertheless, all models so far proposed are unrealistic,
in th at it has been assumed, first, th at the short-wave steepness is somehow rigidly
[ 19 1
2025
20 M. S. Longuet-Higgins
determined, rather than having a distributed probability, and second, that the
height of the longer waves is uniform and given. The purpose of the present paper
is to replace these by more realistic assumptions, and indeed to show that the
randomness of the longer waves plays an essential role in distributing the short
wave steepness.
The basic idea, introduced in §3, is to suppose th at the short-wave steepness 8
a t the crest of a long wave of amplitude A has some probability density depending
on A , and then to relate the density on one wave crest to the density on the wave
crest immediately before. This procedure is made possible by the fact that we
already know, at least approximately, the joint probability density of the
amplitudes A 1 and A2 of two successive waves. (The approximate joint distribu
tion, sometimes called the bivariate Rayleigh distribution, has been used with
success in predicting the properties of wave groups.) With suitable assumptions
regarding the history of the short waves over the intervening time, including
growth by wind action and possible dissipation by breaking, we are able to show
th a t the density p(a\A) of the short-wave slopes on a long wave of given amplitude
A satisfies a certain integral equation, which can be solved in a straightforward
manner by iteration (§4). In fact, the presence of both growth and dissipation are
essential to the convergence of the solution.
In the subsequent sections, we present some results for typical values of the
parameters. These show how the solutions depend upon the rm s steepness of
the longer waves, and on the wind-induced growth rate. I t is notable th at the
short-wave density p(e\A) can be biomodal, with a peak both a t the limiting
steepness s0 and at some lower value of в. In § 6 we investigate the effect of group
length of the longer waves. Somewhat different distributions are obtained for ocean
swell (long-wave groups) and wind waves (short groups), respectively. In §7, we
investigate the result of adding a certain amount of noise to the process of wave
growth.
Sections 8 and 9 contain a physical discussion and a statement of the conclusions.
where L and С are the wavelength and phase speed of the long waves, and cg is
the group velocity of the short waves. Assuming cg -4 С .we have
r « L / C = T, (2.2)
approximately.
2026
Sea-surface roughness 21
?«)
During this time the wind will generally act on the short waves so as to increase
their slope, and if there were no dissipation by wave breaking the short waves
would grow by a factor e^T, where /? is a time rate of growth, related to the wind
speed (see, for example, Plant 1 9 8 2 ). T hat is to say
8 j 8 x = ев , В - fir. (2.3)
On the other hand, if there were no generation or dissipation of short-wave
action, the straining of the long waves would induce a change in the short-wave
steepness depending only on the vertical elevation rj of the waves, for a given long
wavelength (Longuet-Higgins 1 9 8 7 , figure 9). Thus we should have
(2.4)
where К = 2 n /L is the wavenumber of the long waves. In fact, it is a good
approximation to take
f{Ky) = (2.5)
I t will be sufficient initially to retain only the first term in this expression. Then
from (2.4) and (2.5) we have
«*/«1 = e2-08*M«-^>. (2.6)
Supposing now th at in the absence of wave breaking the rate of growth due to
the wind is independent of the long waves, we shall have altogether
в ^ ^ е Я + ^ - 'Ч у = 2.08#, (2.7)
provided th at neither s, or 8г is limited by breaking. In such a case it is convenient
to write (2.7) reciprocally in the form
8l = 82F(A1, A i )t (2 .8 )
where F{Av A 2) = e~B^ {A'-A*). (2.9)
However, the e m s steepness 8 of the short waves is necessarily limited by
breaking. Here we shall assume a sharp limit
*1 ^ e0’ ^ so> (2.Ю)
2027
22 M. S. Longuet-H iggins
an appropriate value of s0 being given by (5.3). Hence 8l is given by the smaller of
(2.8) and (2.10), and e2 by the smaller of (2.7) and (2.10). The situation is shown
schematically in figure 2 .
3. P ro b a b ility d en sities
We propose now to treat both the slope s and the amplitude A as random
variables. The distribution of the long-wave heights will be regarded as given, and
independent of the short waves over a sufficiently long period of time. The problem
then is to determine the probability density of s at given wave amplitude A.
Now the density p(A v A 2) of successive wave amplitudes A x and A 2 for narrow
spectra has been derived and used in connection with the theory of group lengths
(Rice 19 4 4 - 1 9 4 5 , 1 9 5 8 ; Kimura 1 9 8 0 ; Longuet-Higgins 1 9 8 4 )- In fact it is the
‘two-dimensional Rayleigh’ density
f( 11 * ' M l * 4 ( 1 - * 2 M 7 K >
in which A and к are constants and / 0(z) is the modified Bessel function of order
zero. A is equal to the r m s value of the surface elevation rj and к is a ‘groupiness
param eter’ explicitly related to the spectral density of rj. I t can be shown that for
narrow spectra к* « 1 — 4 n2v*, where v is a dimensionless bandwidth (Longuet-
Higgins 1 9 8 4 ). The density of A x or A 2 alone is the Rayleigh density
p(A) = (A / A 2)e~At/i7[i. (3.2)
2028
Sea-surface roughness 23
Hence the probability density of A x given A 2, which is obtained by dividing
p {A x, A 2) by p {A 2), is
Р (А М г) = (1_ ‘t ) i l gi e“» [e-J-Ч /„(*£, S ,)]|
By using the model of § 2 we shall now derive an equation for the probability
density p(s\A) of the short-wave slope on long waves of a given amplitude A.
p(«U)
F i g u r e 3. Sketch of the density p(a\A ), showing the continuous part ф(з, A ) and the singular
part P(A)S(8—80).
First, the fact th a t the slope 8 is limited by breaking means that, for a typical
value of A , there must be an exceptionally large probability density in the
neighbourhood of 8 = s0, as shown in figure 3. In fact, the density p(s\A) has two
components, a continuous component, which we denote by ф{з, A) in 0 < 8 < s0,
and a singular component P(A)S(s —80) in the neighbourhood of s = s0, S(x) being
the Dirac delta function. Normalization of clearly requires th at
|* V (« ,^ )d e + P ( ^ ) = 1. (3.4)
but this takes somewhat different forms according as the continuous or the singular
parts of p(s^A^) отр(8^Аг) are considered. First, the contribution to ф(82, А г) from
Ф(81»A ) presents no difficulties. 8Xis related to s2 by (2 .8 ), so th at we have
24 M. S. Longuet-H iggins
Moreover, the integral in (3.4) must be taken over the range of A x for which 8X< 80,
hence
0 < A l < Лг+у-Ч .В + 1п( 40/«г)] = Н{аг, А г), (3.7)
say. The function H is shown schematically in figure 4.
from (2.10). Therefore, on reversing the order of integration in (3.8) the double
integral becomes
(l/y « ,)/4 A )p (A M i). А = Ж «..Л )- (311)
Altogether, then, when 0 < 8X< s0 we have
ГН(в0.Аш) . i
Ф(8г> A%) — I ф{8ХУА х) F(AX, A 2)p (A 1\A 2)d A x+ ——P(Ax)p(A^A 2)>
Jo 2
A x = H(s2, A 2). i
On the other hand, when s2 = s0 we may integrate each side of (3.5) to obtain
P(A,) = + (3.13)
Physically, this expresses the probability P(A2) as the sum of two parts. The first
comes from the lower long waves, as a result of steepening by straining and by
Sea-surface roughness 25
the effect of the w ind; the second part comes from the slightly higher long waves,
as a, result of regeneration by the wind only, and in spite of the straining.
Equations (3.12) and (3.13) are clearly a coupled pair of integral equations to
determine ф(з, A) and P{A) simultaneously.
4. S o l u t i o n o f e q u a t i o n s (3.12) a n d (3.13)
To find ф{8,А) and P{A) we may proceed by successive approximation.
Arbitrary starting values 0 (1>(s, A) and Z*1^ ) satisfying the condition (3.4) may
be substituted on the right of (3.12) and (3.13) and used to calculate 0 (2)(s, A) and
F*Z)(A). Substituting these again on the right-hand sides of the equations, we may
calculate 0 (3)(s, A) and /* 3)(Л), and so on, the process imitating the development
of the short wave probabilities from one wave to the next. If the functions ф{п){8, A)
and P^n)(A) converge, we have a solution.
This was tried with two different sets of initial conditions, namely
фЫ(8,А) = в ; \ P<1>M) = 0 (4.1)
and Ф{1)(з,А) = 0, = 1. (4.2)
W ith either set, the procedure was found to converge to the same solution,
although with (4.2) the convergence was somewhat faster; typically four decimal
places were obtained after only seven or eight iterations.
At each step, the accuracy of the integrations was checked by evaluating the
left-hand side of (3.4) fdr ф — ф{п+1) and P = P<n+1>; for it may be shown th a t if
any pair ф{п), P<n) satisfy (3.4) then 0 (n+1), P<n+1>, if calculated by the iteration
procedure, should also satisfy (3.4). For graphical accuracy (0 .1 %) it was found
sufficient to take 101 integration points in the range 0 < s /s 0 < 1 and 51
integration points in the range 0 < A / A ^ 5.
5. T h e r a n g e s o f AK, 8 a n d В
The steepness A K of the longer waves will generally be limited by the fact th a t
the ‘ significant waves’, i.e. those with amplitude 2 A, cannot have a steepness much
exceeding the limiting steepness A K = 0.443 for steady progressive waves. Thus
we would expect л __ „ __ 4.
^ 2 A K ^ 0 .4 4 , A K < 0.22 (5.1)
approximately. The above estimate is consistent with the laboratory wind-wave
data of Lake & Y.uen ( 1 9 7 8 ), who found a mean value of A K for the dom inant
waves not exceeding 0.28. For a Rayleigh distribution this would correspond to
{\n)*AK; hence A K ^ 0.224.
We note th at in experiments with unsteady, plunger-generated waves, Ochi &
Tsai ( 1 9 8 3 ) found signs of breaking in individual waves for which A K = 0.35, less
than the limiting value 0.443, but this is not inconsistent with (5.1).
In a similar way, if we consider the short waves as a narrow-band process in
which the local wave steepness a = ak has a Rayleigh distribution,
р(л) = (ot/82) e-at' 2at, (5.2)
2031
26 M. S. Longuet-H iggins
it is reasonable to suppose th at local breaking of the short waves imposes a limit
5 < e0 = 0.22. (5.3)
We note th a t according to (5.2) and (5.3) the local mean-square slope surface slope
a2 satisfies ^ Q Q5 ^ 4j
which is not far from the limiting value 0.04 + 0.02 found by Plant ( 1 9 8 2 ).
For the rate of growth of surface slopes under the action of the wind, Plant
( 1 9 8 2 ) has shown th at the formula
P = 0.02(w*/c)2 o’ cos в (5.5)
fits many different observations. Here u+ denotes the friction velocity, taken to
equal 0.04 times the wind speed U at a standard elevation, с and a are the phase
speed and the radian frequency of the short waves, and в is the angle between
the directions of the short waves and the wind. Because our analysis refers to the
short-wave slopes rather than to their spectral density, we have here divided the
right-hand side of Plant’s equation (1) by a factor 2 . In the special case в = 0> when
wind and waves are in the same direction, (5.5) reduces to
/?= 3.2x10 ~*(U/c)2<r. (5.6)
For example, if we consider short-gravity waves of length I = 20 cm, we have
or = (2ng/l)t = 1 7 .6 rad s -1 and с = д/o' = 0.56 m s_l. Hence with U = 6 .0 m s-1
we find p = 0.065 s_1. With the period of the long waves about 7.5 s (Evans &
Shemdin 1 9 8 0 , table 2), we have В = 0.49 and eB = 1.63. With the long waves
having period 2 s, we have B = 0.129 and eB — 1.13 only. Thus swell and
wind-waves may correspond to rather different values of the wind amplification,
even for the same wind-speed.
With shorter (X-band) short waves, В is generally much greater, but the
amplification of the short waves is then more limited by breaking.
6. R e s u l t s
To explore the behaviour of the solutions, calculations were at first carried out
with representative values of A K and В , and with к = 0. The latter condition
implies zero correlation between successive waves, so th at р(А х\Аг) is independent
of A 2 and is, in fact, given by the Rayleigh density p(A) of equation (3.2), with
A = AV
Figure 5 shows the solution when A K = 0.1 and В = 0.1. I t will be seen th at
as the height A of the waves increases, so the curves of ф(з, A) tend to move to
the right, th at is to say the short waves on the whole become steeper. The
probability P{A) of the short waves attaining their limiting steepness also increases
monotonically with A. For the lowest values of A / A the probability^of finding
breaking waves a t the crests of the longer waves is quite low; when A / A = 0.5 the
probability is less than 20% . When А / A = 2, P(A) is more than 90%.
The probability density ф(з, A) does not always increase monotonically with 8.
In fact, when A / A = 0 and 0.5, the density has maxima near s/80 = 0.83 and 0.92
respectively, as well as the delta-function peak a t s /s Q= 1. The distribution of
2032
Sea-surface roughness 27
F i o u r e 5. (a) The calculated probability density ф{в. A). (6) The probability o f breaking P{A)
when A K = 0.1, В = 0.1 and к = 0.
s/80 A lA
F ig u r e 6 . As fig u r e 5 , w h e n A K = 0 .1 , В = 0 .2 , к = 0 .
2033
28 M. S. Longuet-Higgins
L0
(«)
/
AlA-0.0/ /Г \
/ / \ p 05
//
/ / A X
/
S i i i
0 0£ LO О 2 4
s f Sq A./A.
F ig u re 7. As figure 5, when A K = 0.2, В = 0.1, к = 0.
lJO
4
P 0L6
waves, say А / A < 1 , are determined by the fact th at the short-wave steepness is
limiting on the higher long waves and is then reduced by the subsequent fall in
long-wave height. The regeneration of short-wave steepness by the wind is
insufficient to overcome this effect.
In figure 8 , however, when В is increased to 0.2, the curves for ф(8, A) are shifted
back again toward the right.
Variation of к
In the results so far, the correlation parameter к has been set equal to zero.
Figures 9, 10 and 11 show the effect of increasing к by stages, while keeping A K
and В constant a t 0.2 as in figure 8 .
2034
Sea-surface roughness 29
L0
(a) (b) ' -----’
/
,^05
>4/4=00// ' V t
/ 0& / /
/ / /
1 1 _l
о 05 1.0 • 0 2 4
e/s0 ■ A /A
F ig u r e 9. As figure 8, when к = 0.5.
30 M. S. Longuet-Higgins
8/Sq А /A
F i g u r e 11. A s f ig u r e 8 , w h e n к = 0 .8 5 .
variability of the long waves, and so will force the short waves up to the breaking
steepness a t every long-wave crest. Thus, we shall have
ф(в,А)-> 0, P(A)-+1. (6.1)
This is the situation th at has been assumed, in effect, by previous authors (Phillips
& Banner 1 9 7 4 ).
Generally, however, we see that this condition is by no means attained, and that
the variation of long-wave heights is an important factor in reducing the probable
short-wave steepness, even a t the long-wave crests.
7. E ffect of b a c k g r o u n d ‘n o ise ’
So far we have assumed th at the short waves are present initially and, in spite
of some dissipation through breaking, are always regenerated by the wind. In other
words, we have a ‘boot-strap’ situation, the regeneration of the waves depending
on their initial presence. However, apart from the exponential growth rate of the
waves, as expressed through the factor in (2 .8 ), there may in fact be other sources
of short-wave energy, arising from splashing, nonlinear wave interactions, etc.
W ithout specifying the exact mechanism, we may represent such a noisy source
by a constant input of probability density to the small slopes, a sort of ‘probability
rain ’. For such an input, Аф, a plausible form is
Аф = N(s/s*) e-**'2**, (7.1)
Sea-surface roughness 31
The iteration was found to converge just as before (although a t a slower rate,
comparable to the initial conditions ( 4 . 1 ) ) . The resulting steady solution in the case
N = 0.05 is shown in figure 12. This may be compared with figure 8. Instead of
the almost vanishing values of ф found previously for low slopes (a/e0 < 0.25, say)
we now see a substantial ‘background probability ’ extending over the whole range
of low slopes, well beyond the original r m s value 5 of the noise input ( 7 . 1 ) . This,
of course, is caused by amplification of the noise by the wind stress.
Figure 13 shows the result when N is increased from 0.05 to 0.1. Now the two
maxima in ф are of comparable magnitude, when A / A ^ 2 . A t the larger values
of A / A the short waves are almost uniformly distributed over 0 < 8 < 80. The
probability P of breaking is correspondingly reduced.
e/e* A /A
F ig u r e 12. As figure 8, but showing the effect of added ‘noise’; N = 0.05.
32 M. S. Longuet-Higgins
8. D iscu ssio n
9. C o n clu sio n s
Sea-surface roughness 33
the probabilities of low short-wave slopes. Paradoxically, the probability P{A) of
breaking is reduced.
In a further paper (in preparation) we shall show how the analysis can be
extended so as to calculate the probable distribution of surface slope, s, over the
complete profile of the long waves. This leads to expressions for the mean-square
short-wave slope, the average phase lag between the surface slope and the long
waves, and other observed quantities.
R eferences
Evans, D. A. &, Shemdin, О. H. 1980 An investigation of the modulation of capillary and short
gravity waves in the open ocean. J. geophys. Res. 85, 5019-5024.
Hoogeboom, P. 1985 The determination o f modulation transfer functions from the ‘ Noordwijk ’
radar measurements in 1978 and 1979. TNO Physics and Electronic Lab. Report no. FEL-TNO
1985-70.
Keller, W. C. & Wright, J. W. 1975 Microwave scattering and the straining o f wind-generated
waves. Radio Sci. 10, 139-147.
Kimura, A. 1980 Statistical properties of random wave groups. In Proc. 17th Int. Conf. on
Coastal Engng, Sydney, pp. 2955-2973. New York: American Society of Civil Engineering.
Lake, В. M. & Yuen, H. C. 1978 A new model for nonlinear waves. Part 1. Physical model and
experimental evidence. J. Fluid Mech. 8 8 , 33-62.
Longuet-Higgins, M. S. & Stewart, R. W. i 960 Changes in the form o f short gravity waves on
long waves and tidal currents. J. Fluid Mech. 8 , 565-583.
Longuet-Higgins, M. S. 1984 Statistical properties of wave groups in random sea state. Phil.
Trans. R. Soc. Lond. A 312, 219-250.
Longuet-Higgins, M. S. 1987 The propagation of short surface waves on longer gravity waves.
J. Fluid Mech. (In the press.)
Ochi, М. K. & Tsai, C.-H. 1983 Prediction of breaking waves in deep water. J. phys. Oceanogr.
13, 2008-2019.
Phillips, О. M. &. Banner, M. L. 1974 Wave breaking in the presence of wind drift and swell.
J. Fluid Mech. 6 6 , 625-640.
Plant, W. J. 1982 A relationship between wind stress and wave slope. J. geophys. Res. 87,
1961-1967.
Plant, W. J. & Keller, W. C. 1983 Parametric dependence o f ocean wave-radar modulation
transfer functions. J. geophys. Res. 88 , 9747-9756.
Rice, S. O. 1 944~S The mathematical analysis of random noise. I. Bell. Syst. tech. J. 23, 282-332
and 24, 46-156.
Rice, S. O. 1956 Distribution of fades in radio transmission - gaussian noise model. Bell. Syst.
tech. J. 37, 581-635.
Smith, J. A. 1986 Short surface waves with growth and dissipation. J. geophys. Res. 91,
2616-2632.
Wright, J. W., Plant, W. J., Keller, W. C. & Jones, W. 1980 Ocean wave-radar modulation
transfer functions from the West Coast Experiment. J. geophys. Res. 85, 4957-4966.
2 Vol. 410. A
2039
M ic h a e l S. L o n g u e t - H ig g in s
Center for Studies o f Nonlinear Dynamics, La Jolla Institute. La Jolla. California and Department of
Applied Mathematics and Theoretical Physics. Cambridge, England
ABSTRACT
Laboratory observations of short gravity waves of length about 1 m by Koga show the occurrence of a
downward flow separation near the wave crests, making angles of 10® to 50° with the horizontal. Koga's
observations are compared with a theoretical model of flow separation suggested by Longuet-Higgins. Some
features of agreement are found.
is assumed constant To satisfy both Laplace’s equation and dissipation o f the short waves. Here we shall as
and the condition o f constant pressure when fi = a w e sume that the flow is so strongly turbulent that it is
take as stream function dominated by Reynolds stresses which can be repre
sented by a constant eddy viscosity N , say. (This is a
A ri a sin | (в - a ) (2. 1)
where
- g cosa. ( 2. 2 )
- = gr cosd - - q1 (2.3)
D L
(2.4)
It will be seen that both p and q are proportional to
the radial distance r. There is a stagnation point at r
= 0.
In the shaded region to the left, where - у < в < - 0 ,
the flow is assumed to be turbulent with a constant R e . 2. Estimated velocities V and directions в o f the (low ncai
density p'. The main source o f the turbulence is as the crest of wind waves, as calculated from observations of entrained
sumed to be the free surface, where there is breaking bubbles. (From Koga 1982).
2041
(•)
RG 4. (from Longuet-Higgins 1973). Examples of particular flows when p - p' and C '> 0.
(a) a = 60°; ( b ) « = 75e; (c) 6 = 90».
2042
W e m ay m ention that when p '/p < 1 there exists a applicable to the im mediate neighbourhood o f the
second class o f solutions in which the flow velocity in point (or lin e) o f separation. Like the Stokes 120° cor
the left-hand sector - у < в < ~(3 is relatively small. ner flow, it does not aim to describe the total flow in
Indeed when we let С 0 the fluid on the left becomes the wave, or even in the crest. Nevertheless a complete
static. The inclination o f the interface в = —0 is then description o f the flow must take into account the spe
given by cial conditions existing near the separation point, which
may largely determine the global characteristics o f the
tan/3 = (3 p /p ' - l ) / 3 1' 2. (2 .1 3 ) flow.
If p '/p = 0, we see that the configuration is that o f the Perhaps the most controversial aspect o f the model
sym m etric, 120° com er-flow. If on the other hand p '/ is its m odeling o f the turbulence by a uniform eddy
p = 1, it is found that the flow velocity vanishes ev viscosity. Such an assumption can be expected to yield,
erywhere. at best, only approximate, or qualitatively correct, pre
dictions. Underlying the assumption is that the tur
3. Comparison with observation bulence arises only partly from the shear layer along
the boundary (в = - /3 ) o f the laminar flow. A much
In Fig. 2b, which shows the direction o f the flow as stronger source o f turbulence is the dissipation of en
indicated by the bubble clusters (but allowing for the ergy by short capillary-scale features near the wave
rate o f rise o f the bubbles relative to the surrounding crest. Such dissipation occurs naturally as a result of
fluid; see Koga 1982, section 2 ) the direction varies the breaking o f capillary waves riding on the longer
from 190° to 2 30°, i.e. 10° to 50° downwards from gravity waves. Such capillary waves are blocked by the
the horizontal, with a mean value 39°. This compares gradients o f velocity on the forward face o f the gravity
with the theoretical value 49° shown in Fig. 4 in the waves; see Phillips (1981). The turbulcnce so created
case p' = p, С -► 0. If we assume p' < p the theoretical will tend to be circulated by the mean flow in the tur
angle lies between 49° and 30°. bulent sector o f the fluid.
The magnitude o f the velocities shown in Fig. 2a We have mentioned that the turbulence will help to
above range from 0.4 to 1.2 m s 4 , with a mean value support the inclination o f the free surface to the left
0.8 m s" 1. In the theoretical model the flow velocity q (i.e., up-wave) o f the point o f separation. In addition
in the laminar region is given by equations (2 .4 ) and we may expect a contribution from the radiation
(2 .2 ), hence stresses in the short capillary waves. The effect would
be similar to the wave “set-up” maintained by waves
q 2 = 2 grcosa. (3 .1 )
in a coastal surf zone (Longuet-Higgins and Stewart
This depends both on r and a, but if as representative 1963).
values we take r = 0.05 m , a = 39°, we obtain q = 0.87 We note that Koga (1 9 8 2 ) has proposed an expla
m s -1, within the range o f observation. nation for the entrainment o f bubbles at the point o f
Consider now the inclination o f the free surface. The separation. He suggested that it was analogous to the
photographs in Fig. 1 were taken from an angle slightly entrainment o f air by a jet entering the surface from
oblique to the line o f the wave crests on the left. It can above. The presence o f such jets is not evident in the
nevertheless be seen that the oblique line o f bubbles photographs o f Fig. 2. Rather, we believe the entrain
separates from the surface at a point slightly forward ment o f air is to be associated with the stability o f the
o f the wave crest. In Figs. la and lc the crest profile is flow near the point o f separation. However the inves
fairly sharply curved, so that the inclination o f the tan tigation o f this problem is beyond the scope o f the
gent is not well defined. Nevertheless it can be seen present paper.
that the oblique angle between the line o f bubbles and
the sm ooth part o f the free surface ahead o f the point
o f separation is not far from 120° as predicted by the 5. Conclusions
theoretical model.
In Fig. lb there appears to have been more than one Several features o f the flow separation observed by
separation event. The crest may also have become less Koga (1982) near the crests o f steep gravity waves arc
steep since the earlier events. Nevertheless the angle in agreement with the simple theoretical model pro
between the lines o f bubbles and the tangent to the posed earlier by Longuet-Higgins (1 9 7 3 ). The latter
mean surface is about 135°, only slightly greater than assumes that on one side o f the separating streamline
the predicted value. the flow is irrotational, and on the other side it is tur
From these comparisons we may conclude that there bulent. The turbulence, however, arises mainly from
is a rough, but quantitative agreement between the ob dissipation o f energy at the free surface, and not from
servations and the predictions o f the model. shearing at the separating streamline. The flow is es
sentially nonlinear, and is not contiguous to a slate of
4. Discussion rest. Aeration of the fluid by the entrainment o f bubbles
We emphasize first that the model o f flow separation has only a small effect on the bulk density o f the fluid.
which has been suggested is only a local model, possibly A possible difference in density between the laminar
2043
and turbulent regions has been included in the model Assuming i Ф 0, we have, to lowest order
in order to allow for the bulk effect o f bubbles pene
3 V3 c o s t + sin ? = 0 (Al l )
trating the free surface. However on the small scales
considered here such density differences will probably whence
be slight. Only in the case o f large-scale spilling breakers
where foaming is evident do we expect significant den 7 = - r + arctan3 "3/2 100°54'. (A12)
sity differences (Longuet-Higgins and Turner 1974).
We shall denote this value o f у by 70 .
APPENDIX Now from equations ( A l ) we have
The Evaluation of E and F 3C<2
E- - X<2 ( A13)
In section 3 o f Longuet-Higgins (1973) it is shown sin 35 ' 37 ~ 0
that the boundary conditions (2 .6 ) and (2.7) lead to 3 CQ
the equations ( AI 4)
sin 6 sin 7r/ 3
£ sin 36 + F sin 6 = 0
( A l) as t -*• 0. Moreover from (A 4)
E sin 36 + ^ F sin 5 ■2Cej
(!« •+ « -* )
and
E cos36 + F c o sS (A2) ~ -jj ( g/ W) c o s ( v - 70)
where
6 = y - 0 (A3)
(A15)
Hence the limiting value o f E is
(A 4)
<3 = J2 co s
M - Eo = Q - 0.0156(£/Л 0 ( A16)
Hence
as stated in section 2.
cot 35 (A 5) REFERENCES
Banner, M. L., and О. M. Phillips, 1974: On the incipient breaking
of small scale waves. J. Fluid \fcch.. 65, 647-656.
csc 35 - csc 5 (A 6) Crapper, G. D. 1957; An exact solution for progressive capillary waves
3С
cos of arbitrary amplitude. J. Fluid Mech., 96, 417-445.
Koga, M. 1982: Bubble entrainment in breaking wind waves. Tellus.
34, 481-489.
We exam ine the solution as С -+■ 0 and 5 т /3 . Longuet-Higgins, M. S. 1973: A model of Oow separation at a free
Writing surfacc. J. Fluid Mech.. 57, 129-148.
------ , 1987: Mechanics o f wave breaking in deep water. Proc. NATO
3 С = A, 6 = ir/3 + 5', (A7) Adv. Workshop on Natural Mechanisms o f Surface Generated
Noise in the Ocean. Lerici, Italy, 15-19 June 1987.
we find from (A 5 ), to order <2, that ------ ■, 1989a: Monopole emission of sound by asymmetric bubble
oscillations. I. Normal modes. J. Fluid Mech., 201, 525-541.
J_ 1 1 1 ------ , 1989b: Monopole emission of sound by asymmetric bubble
(A 8)
3« V3 + 3 e ' x - oscillations. II. An initial-vaiue problem. J. Fluid Mech.. 201,
543-565.
Now substituting into (A 6 ) and expanding similarly ------ , and R. W. Stewart, 1963: A note on wave set-up. J. Mar. Res..
in powers o f « we obtain 21,4-10.
------ , and J. S. Turner, 1974: An “ entraining plume" model of a
spilling breaker. J. Fluid Mech., 63, 1-20.
(l + 2Уз -^ < 2| co s ( 7 - с) Phillips, О. М., 1981: The dispersion of short wavelets in the presence
of a dominant wave. J. Fluid Mech., 107, 465-485.
= (1 - V3< + <2 ) c o s 7 (A 9) Toba, Y., 1961: Drop production by bursting of air bubbles on the
sea surface. III. Study by use o f a wind flume. Mem. Coll. Sci.
whence to order e2 Univ. Kyoto. Ser. A. 29, 313-344.
Turner, J. S.. 1969: Buoyant plumes and thermals. Ann. Rev. Fluid
(3V3< - 2e2) COS7 + (« + 2l/3<2) sin7 = 0. (A10) Dyn., 1, 29-44.
2044
1. Introduction
The hydrodynamical interaction of short waves with longer waves on the sea surface
has important consequences both for air-sea transfer processes and for the imaging
of the ocean surface by certain types of remote sensors which depend upon Bragg
scattering from the shorter waves. References are given in Longuet-Higgins (1987;
referred to as part I). In that paper, a model of the short-waves dynamics was
presented which included the following features: (a) modulation of the short-wave
steepness by orbital straining in the longer waves; (b} dissipation of short-wave
energy by breaking, particularly near crests of the longer waves; (c) regeneration of
the short-wave energy by the wind; (d ) a random distribution of long-wave
amplitude.
Particular attention was paid to the behaviour of the short waves at the long-wave
crests, and it was shown how an integral equation could be formulated for the
probability density p(s | A ) of the short-wave steepness s at the crest of a longer wave
of given amplitude A. In part I the author looked forward to a study of the
characteristics of the short waves over the whole profile of the longer waves. From
these one could obtain several parameters of interest such as the mean-square short
wave steepness s2 over the whole wave surface, and the phase and amplitude
modulation of s2 relative to the long waves.
Proc. R. Soc. Lond. A (1991), 435, 405-422
Printed in Great Britain 405
2045
406 M. S. Longuet-Higgins
The present paper gives the results of this study. To present and interpret them
we first describe a simplified model which includes the elements (a), (b) and (c) above
but omits, for the time being, the random element (rf). This simplified model is
described in §§2 and 3 below, and its consequences are discussed in §§4 and 5.
Randomness is then introduced in two stages. A partially random model, in which
the long-wave amplitude has a Rayleigh density but the wave height is locally
uniform (a narrow-band spectrum) is studied in §6. Then in §7 we re-introduce the
full bivariate Rayleigh distribution for the long wave amplitude, and present the
numerical results. In §§8 and 9 we make comparisons with laboratory and field data.
A discussion follows in § 10.
2. Physical assumptions
We assume that the surface-slope spectrum may be effectively divided into two
parts as follows. First, a narrow, low-frequency part, to be called the ‘long waves’.
A representative amplitude and wavenumber for these being denoted by A and K,
the corresponding surface elevation is given by
ij = Acosd, 6 = a t+ K x + e, (2.1)
where a; is a horizontal coordinate, t is the time. A and e are functions of x and t which
vary slowly compared with the phase в. If cr and К are both positive, the waves move
to the left, as in figure 1.
Superposed on the long waves we assume a much shorter, wind-generated, ‘surface
roughness’ with local mean-square slope s 2. Following Plant (1982) we assume that,
below saturation, s grows exponentially with time:
5 oc e* (2.2)
say, where the growth-rate ft depends upon both the wind-speed W and on the
frequency of the short waves. Here we shall take /? to correspond to the mean short
wave frequency cr', say. Note that the empirical law (2.2) already takes some account
of the nonlinear interactions among the short waves themselves. However, the short
wave steepness is assumed to be limited by breaking at a critical steepness s0, say.
For the reasons given in part I we choose
s0 = 0.22, s* = 0.0484, (2.3)
which is consistent with the limits on s found by Plant (1982).
In the absence of wind, the relative steepening of the short waves due to the orbital
motion in the long waves has been shown to be given by
5 oc eyK* (2.4)
approximately, where у = 2.08 (see Longuet-Higgins 1986). Hence (when the short
waves are not limited by breaking) we combine (2.2) and (2.4) to give
s „ efit+ук^ (2.5)
rj being given by (2.1). Thus
3« = \(р+уК.Ъг)/Ы)в, a < s0l 26
d< to, s = «oJ
The case s = s0 corresponds to the parts of the surface where the short-wave
steepness is limited by breaking.
Proc. R. Soc. Lond. A (1991)
2046
________ ►в
о ^ ____^
Figure 1. Time-history of short waves on a uniform train of long waves. The broken line
indicates intervals where the short waves are breaking, (a) Q < 1 , (6 ) Q ^ 1 .
408 M. S. Longuet-Higgins
0 0.4 0.8
Q
Figure 2. The critical phase-angles в ' and в” aa a function of Q = fi/yaA K .
Suppose Q to be given. If в > 1 then equation (3.2) has no solution. Physically this
means that the wind is blowing so strongly that the short waves are everywhere
breaking. If on the other hand Q ^ 1 then equation (3.3) may be solved for 6 \ and
from (3.7) we can then find в".
The results are shown in figure 2, where 6' and OFare plotted against Q throughout
the range 0 < Q ^ I. For example when Q 4 1 (a relatively light wind) then в' is
small and 0" approaches 2n. Thus P and Q approach the crest from either side. This
means that short-wave breaking is confined to the neighbourhood of the long-wave
crests.
On the other hand when Q approaches 1, then both 6' and 6" approach 90°, so P
and Q approach the p o in tif (в = £л). Physically, the short waves are breaking almost
everywhere, and only near the mean water level do we have 5 < s 0. By expansion of
the left-hand sides of (3.2) and (3.7) in powers of the small angles
а.' = \ п - в \ <х" = в " - \к (3.8)
we find a ' = ( l — Я 2)*, a " = 2(1 - Q 2) l (3.9)
In other words, Q is twice as far from Ж as is P.
4. Statistical properties
For comparison with observation, three numerical quantities are specially useful,
namely the mean-square slope
(4.1)
(4.7)
and ? = ^ [ 1 + ^ -Г (2 F + -f’2)d(?
where X =
2
ГW ) соавйв,
7 2
Y = J - Г F(d) sin вАв.
71Jp
(4.9)
Equation (4.8) shows that C(6) takes its maximum value when ф = ф0, where
cos фи = X/{X* + Г)1, sin ф„ = У/(Хг + Г)* (4.10)
and th at Стм = а\А (Х г + Т ) У (4.11)
Hence also c*M = ( ^ + n V 2 te (4.12)
and C*(0) = X /2 iz, (4.13)
We consider now some limiting values. When Q approaches 1 from below, then
[$* —6')-+0 and from equation (4 .6 ) we have
52/ s J - > 1 . ( 4 .1 5 )
Also from (4 .9 ) and (4 .1 0 ), we see that Z - > 0 and 0 (since F < 0) and Y/X -+ 0 ,
hence
0 O-> — 9 0 °, (4 1 6 )
410 M. S. Longuet-Higgins
AK
Figure 3. The relative mean-square slope s*/5o and correlation coefficient С
as functions of AK in the limiting case 12->-0.
Figure 4. Statistical parameters for short waves on a uniform train of long waves, (a) s2/ 5o» (6) Стлх, (с) ф0, (d) C(0).
2051
412 M. S. Longuet-Higgins
might expect. More interestingly, for constant wind, s2/s\ is a decreasing function of
the long-wave steepness AK. This is due to the increased amount of short-wave
stretching, away from the long-wave wrests.
As /?/(г-*0, the limiting values of s2/ s 02 and Cmax are given by (4.19) and (4.23)
respectively. Since for small values of z
I0(z) = 1 +i*2, 1 ,(2 ) = \z( 1 + \ z 2) (5 . 1)
II
long-wave amplitude.
EW с.
C>I
414 M. S. Longuet-Higgins
in which < ) denotes _the statistical average. Similarly we may define the average
quantities <C'*(^)), <s4> and <^2)0 , and finally, as in equation (4.3),
C(0) = «T*(0)>/[<?><7‘>]». (6.4)
The angle ф, is defined as previously, as the angle for which С{ф) takes its maximum
value Cmax.
The results are shown in figure 5 a-d as functions of A*К and where
A* = <Л> = (1я)!Л0 (6.5)
denotes the average long wave amplitude. Figure 5a showing <s2>/s;j, may be
compared with figure 4 a. The form of the contours is very similar, except that the
diagonal contour in figure 4a corresponding to Q = 1 is now smoothed over.
Similarly figure 56 showing ф corresponds closely to figure 46, and figure 5c
showing Cmax corresponds closely to figure 4c. In both figures 4a and 5a, for example,
the contours of constant ф0 have a slope of nearly unity, indicating that 0O is a
function of A*K/{cr/B) very nearly.
ASt°Cha°4crnod e l o f
f ^ r M e г0ид/1Пш ц
SK W - A 0991)
2055
416 M. S. Longuet-Higgins
For general values of к, when the amplitudes Ax and A2 are unequal, the
calculation of C(82,7f) requires that we define the elevation 7) over the interval
0 < в < 4л. We assumed that the amplitude varied ‘piecewise linearly’, that is
([А1+ (Аг —Ах)6] costf, О < 0< 2я,
* “ 1 [Аг + (А3—Аг) (в—2ji)]cos0, 2ti < в < 4л,
where А3 denotes the wave amplitude following A2. To take full account of the
variability of A3 would be impractical computationally, but in the case к = 0 we can
reasonably replace Аъ by its constant mean value <^4> = A*.
The results when к = 0 are shown in figure 6 a~d. Again it will be seen that the
contours of s2/sj, ф, Cmax and C0 are qualitatively similar to those in figure 4.
Table 2. Amplitude and slope of the Fourier harmonics of the mechanically generated wave
n л пк п
1 4.11 0.043
2 0.40 0.015
3 0.46 0.038
4 0.04 0.013
5 0.01 0.008
6 0.02 0.003
The next columns of table 1 show /p, the peak frequency of the slope spectrum and
ft, the growth-rate of the wind-generated waves, as calculated from Donelan’s (1987)
formula
0 = 2.05 x 10-b{ULI2/c-X )2A3(o. (8 . 1 )
Here ULI2 denotes the wind-speed at a height of half a wavelength above the waves.
We equate this to W. We also set w = 2 nfp and с = g/o). If we equate sx to s0, then
we may compare the fourth and the last columns of table 1 . The agreement is closer
at the higher wind-speeds, presumably because the short waves are closer to
equilibrium with the wind and the underlying long waves, as assumed in the theory.
It has, however, been noted by Miller et al. (1991) that the mechanically
generalized waves contained higher harmonics which, while contributing negligibly
to the surface elevation, nevertheless produced significant components of the surface
slope. This is shown in table 2, where the amplitudes An of the various harmonics are
given. The harmonic n = 3, for example has an amplitude A3 which is only one-tenth
of the fundamental Ax. But because the frequency crn is proportional to n, the
wavenumber K n is proportional to n2. Hence А3К г is comparable in magnitude to Ax
K x. The ratio Р/<тп, however, is proportional only to n , and so varies less
dramatically.
In this paper the wavenumber К of the long waves has been assumed to be
constant, or nearly so. To take full account of the higher harmonics would require
development of the theory beyond our present limits. However figure 3 suggests that
if the values of AK and f}/cr for the third harmonic were used instead of those for the
first harmonic, then s2/s\ would be somewhat less than the values given in table 1 .
Presumably the effect of both the first and third harmonics acting_together is to
reduce s2/s\ still further, bringing it closer to the observed value of s2/s\.
It should be emphasized that in these experiments the higher harmonics in the
waveform were practically invisible to the naked eye, yet their effects on the mean-
equare slope s2(6) were quite marked (see Miller et al. 1991). Similar effects may have
been present in the previous experiments mentioned above.
Proc. R. Soc. Lond. A (1991)
2057
418 M. S. Longuet-Higgins
Table 3. Comparison of theoretical and observed values of s2/slfor the Pierson-Moskomtz spectrum
W/(m s-1) fil*9 s2/*l s2= s2„+sl s2 (obs)
6 0.100 0.906 0.051 0.030
8 0.154 0.949 0.053 0.041
10 0.215 0.976 0.054 0.051
12 0.282 0.993 0.055 0.061
14 0.356 0.998 0.056 0.071
9. Field observations
We have seen that the orbital contraction of the long waves is measured not by the
long-wave amplitude A alone, but by the long-wave steepness AK. Thus the
requisite information for the long waves is not the surface elevation spectrum E(a)
which is usually measured, but the slope spectrum
S(cr) = кгЕ(<т)> кг = (т2/д. (9.1)
Often a low-frequency, or swell, component which is visible in the elevation spectrum
E(cr) may not be significantly present in the slope spectrum S{a).
Commonly used expressions for E(<r) such as proposed by Pierson & Moskowitz
(1964) are valid only at frequencies below about 0.5 Hz, and tend to underestimate
the spectral density at higher frequencies. Direct measaurements of the slope
spectrum are rare. Those by Shemdin & Hwang (1988) show a single broad peak, with
a high-frequency behaviour like f~l &approximately.
Field observations of sea surface slopes are known to be strongly influenced by
other parameters such as atmospheric instability, as measured by the air-sea
temperature difference (Hwang & Shemdin 1988). In fact, after removal of the
temperature instability effect from the available data the residuals show a large
scatter and no obvious dependence on the steepness of the dominant waves. (The
value of A used seems to have been calculated from the significant wave height, not
the mean wave height.) It is also unclear whether the ‘significant waves’ tabulated
for example by Cox & Munk (1954) and used in this dataset were sufficiently
separated from the short, steep waves to be counted as swell in the sense of the
present paper.
10. Discussion
The present simplified model of long-wave/short-wave interaction, while including
the features listed in § 1 , has neglected others that may be significant in some
circumstances; particularly the wind-drift surface current discussed by Phillips &
Banner (1974) and the weak nonlinear interactions between the short waves
(Valenzuela & Laing 1972). The importance of the latter relative to the wind
generated growth rate has been discussed by Miller et al. (1991), in relation to the
laboratory experiments described there. The two growth rates were shown to be of
comparable magnitude. It is, however, arguable that the empirical formulae for the
growth-rate derived by Plant (1982) or Donelan (1987), which we have used, do in
fact include the short-wave interactions, so that these are already accounted for,
though roughly.
A further deficiency of the model is the use of a single frequency to characterize the
short-wave spectrum; especially since the growth-rate due to the wind depends
Proc. R. Soc. Lond. A (1991)
A stochastic m odel o f sea-surface roughness. II 419
that frequency. However, it does appear from figures 3-5
strongly on the (nrT S П 10 2’ b°th S^ S2° and °™x and °* dePend much more
short г ng-wave steepness AK than on f$/a. Hence the dependence on
Р е г Ы ^ ^ иеПСУ iS n0t aS critical “ first ^PPears.
cases when th m?St ser*ous limitation of the model is that it applies strictly only in
wind-wave ' 6ft ШГ °*ear seParation between the long waves (swell) and the shorter
laboratorvS |П t и S °^в sPectrum- Such a separation can certainly be achieved in the
tpmnprof П f ocean’ weH-documented cases of such spectra in neutrally stable
temperature conditions are still rare.
Wa^ to аРР^ ^ e theory, which suppresses the assumption of
wavp о П11S ^lven *n fcbe Appendix. In this we consider simply an equilibrium wind-
FUm a s*ngle peak frequency or The low-frequency part of the
hicrh f m 1S assumec^ to be given by the Pierson & Moskowitz (1964) formula. The
loft " ГеяиепсУ Part follows the laser observations of Shemdin & Hwang (1988). The
hicrhi* 1^ a^Prc^^matec^ by a single band of frequencies at the mean frequency of the
, Г, an 2 6 *ow'frequency part of the spectrum makes a small but constant
oht£»n л ^°П S л t 0 mean"s4uare slope. The high-frequency contribution s H 2 is
f ITt к " " ^ure above, with the assumption that s%= 0.0484. Whereas s 2L is
•° 6 a c<^ns^?,nt independent of the wind-speed, s2Hincreases with W on account
? M 8 1j1^rease in s2/sl with ft/cr in figure 6 a. The sum (s\ + s2H) is compared with Cox
un s empirical formula for the total mean squared slope in table 3. It will be
Ti6n ^ 616 a 8 reement between theory and observation at about W — 1 0 m s-1.
so e rate of increase in (s^ + s2,) with W is of the correct sign but too low in
a so u va ue. ГЧо doubt the reason is that much of the observed increase in mean-
square slopes with increasing wind-speed is due to the broadening of the slope
spectrum due to the progressive lowering of the low-frequency cut-off. The highest
requencies in the spectrum are affected less by the ‘swell’ than by the short waves
of intermediate frequency on which they ride; it is only the intermediate waves
which are reduced by the ‘swell’.
Indeed, in a broad spectrum of wind-waves we may expect that reduction of the
energy in the peak frequencies by interaction which swell waves will lead to an
increase of energy in the highest frequencies. Evidence of this effect can be clearly
seen m the spectra recorded by Donelan (1974). To represent the phenomenon
satisfactorily an 7i-scale model is required, with n > 2.
T he calculations in §5 were begun while the author was visiting the California In stitu te o f
T echnology J e t Propulsion Laboratory m 1987. The author thanks particularly Victor Z lotnieki for
technical assistance and Omar Shemdin for general discussions. The completion of the work was
made possible by the Office of Naval Research under Contract N00014-86-C-0303 to Ocean
Research and Engineering, Pasadena, California, U.S.A.
420 M. S. Longuet-Higgins
W being the wind-speed. This has a maximum when
or/cr0 = [4/0/5]* = 0.877. (A 3)
The slope spectrum corresponding to (A 1 ) is
SL(a) = c u r l erW*>\ (A 4)
which has a maximum when
<r/<r0 = (4# = 1.312, (A 5)
differing from (A 3) by a factor 1.5, approximately. We shall denote the frequency
corresponding to (A 5) by crp, so
o-p = 1.312g/W. (A 6 )
Suppose that equation (A 4) adequately represent the low-frequency part of the
spectrum, that is in the range
0 < cr < cr1} a l = n ap, (A 7)
where я. is a factor of order 2. Then in the two-scale model it is reasonable to assume
that the long wave has a mean steepness A*К given by
(A*K)* = \K((AK)*> = iKs[, (A 8 )
where -
4 = J sL(a)A(r. (A 9)
Thus we shall assign to the short waves in the model a mean frequency
o) = (n(7po-q)i (A 14)
where <rp is the peak frequency (A 6 ) and «rq is a cut-off frequency: 2n x 16 rad s
Proc. R. Soc. Land. A (1991)
2060
References
Cox, C. S. & Munk, W. H. 1954 Statistics of the sea surface derived from sun glitter. J. Mar. Res.
13, 198-227.
Donelan, M. A. 1987 The effect of swell on the growth of wind waves. Johns Hopkins Univ. APL
Tech. Dig. 8, 18-23.
Hwang, P. A. & Shemdin, О. H. 1988 The dependence of sea surface slope on atmospheric
stability and swell conditions. J. geophys. Res. 9 3 , 13903-13912.
Jahne, B. 1989 Energy balance in small-scale w av es-an experimental approach using optical
slope measuring technique and image processing. In Radar scattering from modulated wind-waves
(ed. G. J . Komen & W. A. Oost), pp. 105-120. Dordrecht: Kluwer.
Longuet-Higgins, M. S. 1980 On the distribution of the heights of sea waves: some effects of
nonlinearity and finite band width. J. geophys. Res. 85, 1519-1523.
Longuet-Higgins, M. S. 1987 A stochastic model of sea surface roughness. I. Wave crests. Proc. R.
Soc. Lond. A 410, 19-34.
Miller, S. J . & Shemdin, О. H. 1991 Measurement of the hydrodynamic modulation of centimeter
waves. J. geophys. Res. 9 6 , 2749-2759.
Miller, S. J ., Shemdin, 0 . H. & Longuet-Higgins, M. S. 1991 Laboratory measurements of
modulation of short wave slopes by long surface waves. J. Fluid Mech. (In the press.)
Mitsuyasu, H. 1966 Interactions between water waves and wind. I. Rep. Res. Inst. Appl. Mech.
Kyushu Univ. 14, 67-88.
Phillips, О. M. & Banner, M. L. 1974 Wave breaking in the presence of wind drift and swell.
J. Fluid Mech. 66, 625-640.
Pierson, W. J . & Moskowitz, L. 1964 A proposed spectral form for fully developed wind seas based
on the similarity theory of S. A. Kitaigorodskii. J. geophys. Res. 6 9 , 5181-5190.
Plant, W. J. & Wright, J . W. 1977 Growth and equilibrium of short gravity waves in a wind-wave
tank. J. Fluid Mech. 82, 767-793.
Proc. R. Soc. Lond. A (1991)
2061
422 M. S. Longuet-Higgins
Plant, W. J . 1982 A relationship between wind stress and wave slope. J. geophys. Res. 87,
1961-1967.
Shemdin, О. H. & Hwang, P. A. 1988 Comparison of measured and predicted sea surface spectra
of short waves. J. geophys. Res. 93, 13883-13890.
Valenzuela, G. R. & Laing, М. B. 1972 Nonlinear energy transfer in a gravity-capillary wave
spectrum, with applications. J. Fluid Mech. 54, 507.520.
Wenz, F. J . 1976 Cox and Munk’s sea surface slope variance. J. geophys. Res. 81, 1607-1608.
Received 20 February 1991; accepted 25 March 1991
1. Introduction
Some interesting examples of very high accelerations occurring in surface gravity
waves were recently reported by Cooker & Peregrine (1990); see also Peregrine &
Cooker (1991). Their numerical calculations of steep waves meeting a vertical wall (or
of two such opposite waves meeting) show that the formation of a wave trough near
the wall is sometimes followed surprisingly by a thin vertical jet which shoots
upwards with accelerations typically of 20-50 g, and occasionally as great as 1 0 0 0 g
(Peregrine & Cooker 1991). The jet is accompanied by correspondingly large
pressures and pressure gradients at the wall itself.
Upward-pointing jets induced by solitary waves meeting a wall have also been
recorded experimentally by Nishimura & Takewaka (1990). Such jets are often seen
against the bow of a ship. It is likely that the axial jets seen in standing waves and
collapsing bubbles (Longuet-Higgins 1983 c) are closely related.
In spite of these calculations and observations, there is little understanding of the
phenomenon, or of the range of conditions under which it occurs. For this purpose,
purely numerical calculations are inadequate, and cannot always describe accurately
the asymptotic form of the fluid flow when the free surface develops a sharp
curvature. What is required is an analytic solution which can unify the different
numerical computations and define the asymptotic behaviour of the flow.
Exact solutions for time-dependent flows with a free surface are rare. A flow which
has some of the appropriate features is the ‘Dirichlet hyperbola’ ; see Longuet-
Higgins (1972). In this flow the free surface has the form of hyperbola with varying
2063
450 M. S. Longuet-Higgins
angle в between the asymptotes. When в approaches 90°, the velocity and
acceleration become infinite, like Г 5 and Г* respectively, where t is the time measured
from the critical instant. A physical explanation has been given by Longuet-Higgins
(1983c). There are two weaknesses in this solution, however. One is the very
restricted form of the free surface, which does not allow the formation of a cusp -
only a thin, ultimately parabolic, jet as t-+ oo. Secondly, it is not immediately clear
how the solution is to pass through the infinity at time t = 0 .
These considerations prompted the author to search for a more general analytic
solution. In this and two companion papers use is made of a parametric
representation of irrotational flow first suggested by John (1953). The representation
is sometimes called ‘semi-Lagrangian \ The present author has already expressed the
Dirichlet hyperbola and its extension to a rotating free surface, in a semi-Lagrangian
form; see Longuet-Higgins (1983a, 6 ). With this representation, as is shown in the
present paper, it is a simple matter to generalize the Dirichlet hyperbola by adding
an extra term linear in w and t. This has a non-trivial effect; for example it makes
possible the formation of a cusp at the free surface. It also gives rise to a beautiful
family of solutions which have the principally desired feature, the occurrence of an
inertial shock at a certain instant. It can be shown, moreover, that by the addition
of two further terms the infinite pressure and acceleration are replaced by
momentarily large but finite values (see §13 below).
The plan of the present paper is as follows. Since it may be convenient for many
readers, a brief introduction to the semi-Lagrangian representation is given in §2. In
the following section we give the expression for the Dirichlet hyperbola in semi-
Lagrangian form, and in §4 state the first generalization. It is shown that this leads
to two coupled, nonlinear, ordinary differential equations for the coefficients^/) and
G(t) as functions of the time. One of these equations (equation (4.10)) follows from
the free-surface condition, the other, (4.11), from the condition that the flow be free
of singularities within the fluid. In §5 we see that in the simplest case, (4.10) leads to
a simple mapping of the trajectories in the (F, G)-plane. The time-dependence of the
solutions is described in §6 , including the mapping of the ‘shock line’. In §7 we
describe the behaviour of the solutions in the neighbourhood of the shock line, and
show that F and G behave like J* there, just as for the Dirichlet hyperbola. There is
one exceptional trajectory (§ 10) where the time-dependence is like F. Knowing this
behaviour one can integrate up to and across the shock line without difficulty (§8 ).
The mapping of the solutions in the general case is described in §§ 11 and 1 2 . Section
13 contains some discussion and an indication of further developments.
X= J 2*(w>0 0 (2-7)
452 M. S. Longuet-Higgins
where a prime denotes d/d/, while
zu = F cosh (o + iGsinh iо. (3-3)
Hence the boundary condition (2.3) is satisfied provided that
Fn = -RG, G" = RF. (3.4)
On eliminating R from these equations we have
FF" + GG" = 0. (3.5)
As for the interior condition, it is clear from (3.3) that zu will vanish on the axis
of symmetry (oj pure imaginary) when
(3.6)
and so (3.7)
However, we have also from (3.3) that
z j = F' cosh и) —iG' sinh to. (3.8)
The zeros of z*t therefore coincide with those of zu provided that
F'/F = - G'/G, (3.9)
that is F'G+FG' = 0 (3.10)
or FG = constant. (3.11)
From the two equations (3.4) and (3.11) a differential equation for R(t) and hence
F and G can be obtained (see Longuet-Higgins 19836). An alternative method is
given in Appendix A.
The free surface corresponding to (3.1) is clearly a hyperbola:
(3.12)
with semi-axes equal to F and G respectively. The branch points (3.7) are at the two
foci, and both are annulled by the condition (3.11).
We remark that since the fluid is on one side only of the free surface, it is really
unnecessarily to annul both of the branch points; only the branch point within the
fluid itself need be annulled. Hence we suspect the existence of a more general
analytic expression of this type in which only one branch point is annulled.
4. A first generalization
The key to a more general type of solution than that of §3 is to add to the right-
hand side of (3.1) a term linear in both ш and t. Thus let
z = Fsinh aj + iGcosh w + yw + i//, (4.1)
where у = a + /ft, (4.2)
2066
454 M. S. Longuet-Higgins
X
F i g u r e 1 . (a )
A typical profile of the free surface corresponding to (4.1), showing the two branch
points corresponding to zeros of гы. [F ——1.5, G = —0.8.) (6) Streamlines corresponding to the
solution of (a) when (2.10) is satisfied at the lower branch point. (F' = 1.0, G' = —2.0598.)
Provided (F2+ G2—y 2) > 0, z also is pure imaginary, and the branch points lie on the
imaginary axis (see figure la). This axis intersects the surface at oj —0, z = i((7+//).
Thus provided (F2—y 2) > 0 one of the branch points always lies ‘above’ the free
surface, and one *below ’ the surface, so it is the negative branch point that is to be
excluded.
2068
5. M ap of the solutions: /? = 0
To fix the ideas, let us take first the simplest case: /? = 0 . Equation (4.10) then
reduces to
Ф = {F'G+FG')2- a 2{F'2+ G'2) = 0, (5.1)
that is ( ^ - а ^ 'Ч д а 'С Ч ^ - а 2) ^ 2 = 0. (5.2)
For any given values of F and G this defines in general two directions in the (F , G)-
plane given by
F' -F G ± ol[F2+ G2- o.2)'
G' G2- a 2 ( *)
provided that F2+ G2—a 2 ^ 0. (5.4)
Assuming a to be positive, the set of directions corresponding to the positive sign in
(5.3) is shown in figure 2. The circular region in the centre is excluded by (5.4). The
pattern clearly has a 4-fold rotational symmetry about the origin.
The set of directions corresponding to the negative sign in (5.3) is exactly similar
to figure 2 , but with the opposite direction of rotation; the two patterns are mirror
images of each other.
Equation (5.2) or its solution (5.3) is in fact the condition that at least one of the
zeros of z w be annulled by one of the zeros of z * t . Now the zeros of z * t are given, as
in (4.12), by
ew= [ - p±i(F'2+ G'2- - \ G ' ) (5.5)
and when /? = 0 it is evident that the two roots cannot both annul roots of (4.12) (in
which у = a) except in the special case а = 0 . Thus in general, only one root of the
roots of zw= 0 is annulled. It turns out that choosing the plus sign in (5.3) always
annuls the singularity below the free surface (see figure 1 ) while choosing the minus
sign annuls the singularity above the free surface. We shall be concerned only with
the former case, for which the pattern of directions is as in figure 2 .
It should be noted that while the subsidiary condition (2.10) is satisfied exactly at
all times, this removes only the principal part of the singularity at zu = 0. In figure
1 (6 ) we have plotted the instantaneous streamlines for a solution in which the free
surface is as in figure 1 (a). This shows that on the part of the axis of symmetry below
the branch point the streamlines are very nearly, but not quite, parallel to the axis.
To accommodate the flow, we must assume a weak line sink along the y-axis between
the branch point and у = —oo. Thus, while the solution is not exact in this respect,
it appears as a very good approximation.
We are interested especially in flows that develop upward-pointing cusps, without
loops. This implies that, over some part of the trajectory, F < —a and G < 0. Clearly
the trajectories of interest are confined to the region — oo <F < — a , — oo <G < cl
shown in figure 3 .
2069
456 M. S. Longuet-Higgins
F/a
F ig u r e 2. Directions of the trajectories of the coefficients F and О in the (F, (?)-plane.
F/a.
F ig u r e 3. Trajectories of [F, 0) in the relevant part of the plane: F < —a ,0 < 1.
The trajectories are of three types. Type A starts from F = —oo, G —a and
becomes asymptotic to the cusp line F ——a, G < 0. Each trajectory crosses the
6?-axis at some point (JP0>0 ) between —oo and —a.
Type В also starts from F ——oo, G ~ a, but crosses the cusp line F = —cl at some
point where G > 0.
2070
458 M. S. Longuet-Higgins
Thus the critical circle is also part of the locus A = 0 or of R = oo. However, it lies
outside the region of interest for us.
Lastly we note that a general expression for the curvature к of the trajectory is
F'G"-F"G'
K~ + ' (6,11)
So on substitution for F" and G" from (4.5),
460 M. S. Longuet-Higgins
t
F ig u r e 4. The behaviour of A and A3 in the neighbourhood оf t = te for the typical trajectory
(F0, g0) = ( - 1.3,0), f ; = 1 .0.
-5
(9.1)
where an and bn are constants. Thus the gradient of the surface at w = ± oo, namely
G/\F\, tends to zero like When t = —oo the surface becomes flat.
As t increases from —oo to about —1 the value of G/\F\ increases to a maximum,
and the surface develops a hollow, as in figure 6 (a), with maximum gradient of about
13° when t = —1 . Then the gradient decreases to zero at t = 0 , when (F , G) crosses the
axis G = 0; the hollow ‘fills up’ and the surface is plane; see figure 6 (6 ). The particle
velocities, however, do not vanish. The velocity vectors are shown at various points
along the surface.
For positive values of t (figure 6 c) the surface becomes convex, and since G < 0 the
gradient at infinity is negative. The shock line is crossed at time tc = 0.14866. As
shown by the velocity vectors, the velocity undergoes a sharp kick, resulting in an
S-shaped trajectory for each particle (see figures 6 d and 126).
As t increases beyond tc the surface develops a sharp curvature at the crest; see
figure 6 (c). It may be shown that for large positive times t
F(t) ~ —a + axГ 1 + a 2 1~2+ ...,
(9.2)
G(t) ~ C+1+ 60 + 6 j t~l + 62 1~2+ ... j
462 M. S. Longuet-H iggins
F ig u r e 6 (a , 6 ). F o r c a p tio n see fa c in g p a g e .
The condition for a cusp, F = —a, is not attained in finite time, but only in the limit
as t ^ c o . Nevertheless the form of the surface is quite close to a cusp even at
moderately large times t.
Figure 7 shows a similar sequence of profiles for a typical trajectory of Type B,
which does not cross the axis G —0. Thus G is always positive and the surface is
always concave. The profiles were obtained by integrating backwards and forwards
H ighly a ccelerated, free-su rfa ce flow s 463
т---- 1---- j---- 1---- 1---- 1---- 1---- 1---- 1---- 1---- 1---- 1---- 1---- 1---- г
J___ I___ 1 , I___ I___ I___ I___ I___ I___ I___ I___ I___ I------1------L
-2 0 2
F ig u r e 6. A
sequence of surface profiles for the Type A solution (F0,G0) = ( —1.3,0), F'0 —1,
Я 0 = 0. H'0 =
10.0. (a) - 5 < * < - 1 ; (6) —1 ^ < 0, and (c) 0 ^ i < 0.4532. Velocity vectors are
marked for the particles at w = —5.0 (0.25) 5.0. (d) Typical paths of particles in the free surface.
from time /= 0 (say) with initial conditions (FyG) = ( —2.0,0 .8 ), F'0 = 1. The
behaviour as t ^ —oo is similar to that in Type A trajectory; see (9.1). As t increases
from —oo the surface gradient s^ at (o = oo again increases, so that a hollow
develops. The cusp line is always reached in finite time in this type of trajectory but
2077
464 M. S. Longuet-Higgins
F ig u r e 7. (a) A sequence of surface profiles for the T ype В solution (^0,(?о) = ( —2.0,0.8), F'0 = 1 ,
H0 = 0 ,H'0 = 3.0, when —5 < I < 1.0600. (6 ) The downward-pointing cusp in (a) a t tim e t = 1.0600.
the cusp points downwards. Beyond F ——a, the surface profile develops a loop, and
the solution becomes unphysical. What happens in practice may be that the sharp
cusp entrains some air below the water surface at this point.
The third type of trajectory, which is intermediate between Types A and В will be
discussed in the following section.
2078
466 M. S. Longuet-Higgins
F i g u r e 8. The surface profiles for the solution corresponding to the critical trajectory (Type C).
(a) / ^ 0, (6) 0.
Thus the approach to the shock line is still described by a power law, but the powers
are different from the general case.
The sequence of surface profiles is shown in figure 8 . Again, as /< —oo the surface
is plane. As t increases the surface becomes concave upwards, with a maximum
” 8™ ЬУ U0..7,
when t = —0.56. As £->0 the surface again reverts to a plane (figure 8b), but with a
non-zero velocity field. In this case, therefore, it appears that the cusp line is reached
in finite time.
2080
-(^ o + l ) - 1
F i g u r e 9. The ratio CJC_ shown as a function of — (F0+ 1)"1, when G0 = 0.
C. c+ cjc. -(^o + i r 1
—00 1 .0 0 0 0 - 1 .0 0 0 0 - 1 .0 0 0 0 .0 0
-2 .0 1.0 4 8 3 - 1 .1 8 8 3 -1 .1 3 4 1 .0 0
—1.5 1.0 9 9 6 - 1 .7 2 1 8 -1 .5 6 6 2 .0 0
-1 .3 1 .1 5 3 0 -2 .1 9 6 0 - 1 .9 0 5 3 .3 3
-1 .2 1 .2 0 5 6 -2 .6 2 7 7 - 2 .1 8 0 5 .0 0
-1 .1 1 .3 1 0 2 -3 .6 4 5 1 -2 .7 8 2 10 .0 0
-1 .0 5 1 .4 3 0 0 -5 .0 0 1 3 -3 .4 9 7 2 0 .0 0
T a b l e 1. Computed values of C+./C_ for various values of F0
For small negative values of ty the solution near the origin is self-similar. In fact
by setting ш = FQ, Z = ftz, с = —J and choosing a suitable expression for H{t) the
surface profile (4.1) reduces to the self-similar form
468 M. S. Longuet-Higgins
Table 1 shows the values of C_ and C+calculated by numerical time-stepping and
extrapolation of F'{t) and G'(t) to t ——oo and + oo respectively, for a sequence of
values of F0 approaching —a. When plotted logarithmically in figure 9 we see that
they follow rather closely a power law; in fact
C+/C_~(Fq+1)-* (1 0 .2 1 )
approximately. We conclude that the upwards velocity of the crest of the surface
profile should be considered as infinite, in the limiting case. In figure 8 (6 ) we show the
surface profiles developing in time, but scaled by this infinite velocity. Apart from
the scaling factor, the solution is quite similar to the finite solution shown in figure
6 (c), beyond about t = 0 .2 .
-5 -4 -3 -2 -1 0
Р
F i g u r e 10. (a ) Trajectories through the point ( —2,0) in the (P, (J)-plane, when P\ is positive.
(6) Trajectories through the point ( - 2, 0) in the (P,£)-plane, when P\ is negative.
2083
470 M. S. Longuet-Higgins
Now substituting for fP from ( 1 1 .4) we obtain
A = (PQ'+P'Q) (.PP' + QQ'/П [РГ - QQ'/t*)
t2 P'2+ Q'2/tA ' (И.
The factor (PP' —QQ'/t*) can be shown to vanish only on the ellipse (11.3), and the
shock line is given by
[PP' + QQ'/t*) = 0. (11.13)
Now (11.13) can be written
t2P' Q/t2
-Qr = - ^ j r - (1114)
Since t2P'/Q' is also given by (11.7), we see that the equation of the cusp locus is a
function only of P and Q/t2, independently of other variables. And it is the same
equation as in the case /? = 0 treated earlier, except that now P, Q/t2 and /? take the
place of F, G and a.
Suppose then that we set
Q = Q/t2. (11.15)
Then in the (P, §)-plane, but not in the [P, Q)-plane or the (F, (?)-plane, the equation
of the shock line is invariant with respect to the time. So also is the boundary ( 1 1 .8 ),
which becomes the ‘circle’
P 2+ Q2 = ft2, (11.16)
and also the cusp line, P = —/?. It is clearly advantageous, therefore, to map the
trajectories of (F, G) in the (P, $)-plane, where these crucial features stay put (see
figure 1 0 ), rather than in either of the other two planes.
point ( —ft, Q) on the cusp line (P = —Д). Each trajectory also crosses the shock line
at some instant t = tc , where P and Q behave like (t —£c)*, as described in §7. Since the
limit points of P and Q are not at the origin, F and G must behave asymptotically
like t. The free surface is then of the form
z ~ t(Lsinhoj + iMcosh (jj—fioj), (12.3)
where L and M are constants. This is a cusped curve of fixed shape, with dimensions
increasing linearly with time.
However, when о < p ; < P'.crit the trajectories do not cross the shock line, but
instead tend to limit points on the Q-axis. In the critical case P '^ P lc rit
trajectory touches the shock line.
The asymptotic behaviour of F and G at the limit points can be investigated by
Wntmg F = t(A+0, G = t(M+ v ), (12.4)
where A and fi are constants and £, tj tend to 0 in the limit. On substituting in ( 1 1 .2)
we find to lowest order ^ „
Ф=е4(А2 - Л / ^ 2> (12-5>
which must vanish for all values of t. Hence either A = —ft or fi = 0. The first case
describes the limit points on the cusp line. The second case describes the limit points
on the $-axis.
For sufficiently small values of PJ, the asymptote to the trajectory as /-►0 makes
an angle greater than 45° with the negative $-axis. Hence the trajectory must
intersect the second shock line, touching the circle £ at the point (0 , /?); see figure
1 0 (a).
On the other hand, when P\ < 0 (figure 106), each trajectory has an asymptote
P = P0 as J->0 . It then crosses the shock line and as t-+ oo tends to a limit point on
the negative Q-axis. The free surface is asymptotically flat.
Figure 11 shows the family of trajectories through a point [P^Qx) lying off the
Ф-axis. The family contains some curves which cross the cusp line above the point
2085
472 M. S. Longuet-Higgins
{—ft, 0). In these, the free surface develops downward-pointing cusps at a finite time
t,which subsequently develop into loops. There exists a family of such trajectories
passing through the point (—/?,0). The limiting form of the free surface is derived
below in Appendix B.
This work has been supported by the Office of Naval Research under Grant No.
N00014-91-J-1582.
f ’ =exp[±I(i^ )H (A10)
and that G = —1/F. The path of a particle at the free surface (w fixed and real)
relative to a frame moving downwards with speed Vis now given parametrically by
x = F s inhw, у = Vt—F~xcosh a). (A ll)
When В 4: 1 we have from (A 9) that t ~ IB3, so В ~ (30* and so from (A 10)
F ~ 1±|(30*- (A 12)
Thus the particle velocity becomes infinite like H and there is a typical weak ‘shock
Ifl FLM 248
2087
474 M. S. Longuet-Higgins
(*)
(b)
In the special case V= 0, the particle paths satisfy xy —constant and all are
hyperbolae; see figure 12(a). The large velocities near /= 0 are not immediately
obvious. However, if we view the trajectories in a moving reference frame by taking
V= 2, for example, we get figure 1 2 (6 ), in which the typical 5-shaped kinks appear.
This shows that the kinks are in a sense artifacts of the reference frame. On the
other hand they could be useful experimentally as a diagnostic for weak shocks.
2088
Appendix B. The lim iting form of trajectories through the point ( —/?, 0) as
t—>00
For large values of the time t, the coefficients F and 0 have the asymptotic
expansions
F = -/ ft(l+ a1*2*'+ a2*4*'+...)} G = pt(bxГ+ 6 2 Р + ...). (В 1 )
where v is a negative index. Substituting into ( 1 1 .2 ) and (4.1 1 ) we find to lowest order
Ф a [2(y + 2) (Зу + 2) а г 6 J —(2v)2a\] t*v+2 = 0 (B 2 )
and FF" + GG" a [2v{2v + 1 ) ax+ v(v + 1 ) b\] t2v = 0 (B 3)
respectively. On eliminating the ratio ax/b\ from these equations we obtain a cubic
equation:
7i'3 + 20y2+ 16^ + 4 sb 0 (B 4)
for the exponent v, which has only one real solution, namely
v ——1.7231. (B 5)
Correspondingly we find a 1/6 f = —0.1478. (B 6 )
When ш/Р <€ 1 it is easily shown that the free surface has the self-similar form
R E FE R E N C E S
C o o k e r, M. J . & P e r e g r i n e , D. H. 1990 Violent water motion at breaking-wave impact. In
Coastal Engineering (Proc. ICCE Conf., Delft, Holland), vol. 1, pp. 164-176. ASCE.
J o h n , F. 1953 Two-dimensional potential flows with a free boundary. Commun. PureAppl. Maths
6 , 497-503.
L o n g u e t - H i g g i n s , M. S. 1972 A class of exact, time-dependent, free-surface flows. J. Fluid Mech.
55, 529-543.
L o n g u e t - H i g g i n s , M. S. 1976 Self-similar, time-dependent flows with a free surface. J. Fluid
Meek. 73, 603-620.
L o n g u e t - H i g g i n s , M. S. 1983a On the forming of sharp corners at a free surface. Proc. R. Soc.
Lond. A 371, 453-478.
L o n g u e t - H i g g i n s , M. S. 19836 Towards the analytic description of overturning waves. In
Nonlinear Waves (ed. L . Debnath), pp. 1-24. Cambridge University Press, 360 pp.
L o n g u e t - H i g g in s , M. S. 1983c Bubbles, breaking waves and hyperbolic jets at a free surface.
J. Fluid Mech. 127, 103-121.
L o n g u e t - H i g g in s , M.S. 1993a Highly-accelerated, free-surface flows II. Space-periodic
solutions. (To be submitted.)
L o n g u e t - H i g g i n s , M. S. 19936 Highly-accelerated, free-surface flows. III. Inertial shocks of
bounded intensity. (To be submitted.)
N is h im u r a , Н. & T a k e w a k a , S. 1990 Experimental and numerical study on solitary wave
breaking. In Coastal Engineering (Proc. ICCE Conf., Delft, Holland), vol. 1 , pp. 1033-1045.
ASCE.
P e r e g r i n e , D. Н. & C o o k e r , M. J . 1991 Violent motion as near-breaking waves meet a wall.
Proc. IDTAM Symp. on Breaking Waves, Sydney, Australia, July 1991.
2089
It is shown theoretically that the crest o f a steep, irrotational gravity wave, considered
in isolation, is unstable. There exists just one basic mode of instability, whose
exponential rate of growth fi equals 0.1 23(g/R)\ where g denotes gravity and R is the
radius o f curvature at the undisturbed crest. A volume o f water near the crest is shifted
towards the forward face o f the wave; the ‘ toe’ of the instability is at a horizontal
distance 0.45Л ahead o f the crest. The instability may represent the initial stage of a
spilling breaker. On small scales, the ‘ toe’ may be a source o f parasitic capillary waves.
1. Introduction
Despite much interest during the last twenty years, the problem of how and why
gravity waves break in deep water has remained incompletely solved; for recent reviews
see Longuet-Higgins (1988) and Banner & Peregrine (1993).
As pointed out previously (Longuet-Higgins 1981), a breaking wave does not
necessarily pass through the limiting configuration for steady, steep irrotational waves
first calculated by Michell (1893), with a sharp angle of 120° at the crest (the ‘ Stokes
comer-flow’). The actual flow being unsteady, this limiting configuration is by-passed,
by a less or greater margin. Thus from observation it is found that the mean ratio H /L
o f height H to wavelength L in a breaking wave is usually less than the value 0.141 for
limiting waves. The latter would correspond to a ‘steepness’ parameter ak — nH /L =
0.443. For breakers resulting from the fastest-growing Benjamin-Feir instability it is
found that ak « 0.38 (see Longuet-Higgins & Cokelet 1978, figure 22(c)). In a typical
sea-state, ak may lie between 0.05 and 0.20 (see Holthuijsen & Herbers 1986).
The Benjamin-Feir instability also does not account for the marked horizontal
asymmetry o f steep waves prior to breaking (Kjeldsen & Myrhaug 1980; Bonmarin
1990), although Tulin & Li (1991) have shown that this type of instability does produce
an increased supply of energy to the forward face of the wave.
The general approach o f regarding the initial stages of wave breaking as being due
to an instability o f a steady flow is attractive, however. In the present paper we consider
first the instability o f a single, steep wave crest, as described by the theory of the
‘ almost-highest’ wave (Longuet-Higgins & Fox 1977, 1978). This theory describes the
flow near the crest o f any steady, irrotational wave having slightly less than the
maximum height. The curvature o f the surface profile is everywhere finite. In the ‘ inner
flow’ the surface slope tends to 120° in each direction. We here consider time-
dependent perturbations of this flow. If the crest is perturbed by the addition o f extra
energy, how does the deformation develop? What are the normal modes of
2090
117
F igure 1. Profile of the almost-highest wave, showing coordinates and the 30° asymptotes
(from LHF1).
by (2.2). For large values o f z the inner flow tends to Stokes’s 120° comer flow,
z ~ ( fi* ) ’ as z-> oo. (2.7)
To calculate the complete inner flow LHF1 make the transformation
p-'\X (2 .8)
l* = /?T W
where fi is a positive constant, and then suppose that inside the circle \<o\ = 1 (which
corresponds to the fluid domain xjr < 0) the solution can be expressed in the form
z = (8+iX)l B(a>), (2.9)
where 8 is a positive constant and
B((j) = BQ4- (о + В 2uP + •. • (2.10)
is a power series in u with real coefficients Bn. It is convenient to take 8 = /?. To ensure
that the free surface passes through the wave crest, where x = 0, <o = 1, we must specify
that я ( 1) = г 0+ г , + г г+ ... = r * . (2. 11 )
Since the coefficients Bn are all real, the flow will be symmetric about the line x = 0.
Now to satisfy the free-surface condition (2.4), we have from (2.9)
= K8+ix)-bH(wl (2 . 12 )
where H(oj) = fB(oj) —( 1 + o>) B'{oj) (2.13)
and a prime denotes d/dw. The boundary condition (2.4) then becomes
+ = 1, (2.14)
to be satisfied on the circle M = 1 when r = argw lies in the open interval ( —я, я).
Since w- * may be expanded in a Fourier series:
(2.15)
—oo
sin(rt + i)7C
where = n = 0 , ± 1, ± 2 , (2.16)
(л+|)я
we can substitute into the boundary condition (2.14) and by equating coefficients o f
cos пт obtain an infinite set o f cubic equations for the coefficients Bn.
In LHF1 this system o f equations was solved numerically, together with (2.10),
taking 8 — 10*. On successive truncation o f the system at increasing values o f л a
solution was found which converged satisfactorily. The profile o f the free surface is
plotted in figure 9 o f LHF1.
2094
In n
F igure 2. Plot of ln|2?J vs. In я when n = 1, ...,9 0 , showing the asym ptotic behaviour o f the
coefficients.
Equating coefficients of cosw and sin пт in each of these equations gives four
simultaneous equations for the vector
x —(Aq, aXi..., bQ,bXi..., Cj, c2, . •., dxi d2i ...), (5.24)
see the Appendix. These four equations all have real coefficients. We now seek normal-
mode solutions in which Ъ/bt ~ ft. Thus equations (5.23) yield a matrix equation of the
form , л
PAx = Bx, (5.25)
where >5 and В are matrices of infinite order. Truncating the vector x by setting
Bn = 0 when n > N, say, and letting ant bn, cn, dn also vanish when n > N, we can
choose (4N+2) of the equations (5.25) to determine a sequence of eigenfrequencies or}
and the corresponding eigenvectors xy Further details are given in the Appendix.
6. Results
The computations were programmed in FORTRAN (double or quadruple
precision), with the aid of a CLAMS library routine SGEEV for extracting the
eigenvalues and eigenvectors of a real, non-symmetric square matrix, of order (2 N+ 1)
(see the Appendix). The lower eigenvalues, for N = 30(2)50 are shown in table 1.
They appear to have converged reasonably well. No use has been made of Pade
approximants.
As expected, there is one (double) root p —0, for, from equations (A 7) and ( A ll)
of the Appendix, we see that the elements of the first column of A22 are all zeros. Hence
by (A 1) and (A 2) there is a solution in which p —0, a 0 Ф 0 and all the other
coefficients vanish. This corresponds to a purely horizontal shift of the surface profile
through a distance a0, without change of form.
Table 1 shows further that p has just one positive value, corresponding to an
exponential rate of growth. All the remaining eigenvalues p are pure imaginary,
corresponding to oscillatory normal modes.
The most interesting mode is clearly the exponentially growing mode. Table 1 shows
that by N = 50 the corresponding value of p has converged effectively to
P = 0.0544.... (6.1)
The form of this mode is shown in figures 3 and 4, at successive times /, illustrating
the exponential growth. The stability consists of a forward displacement of the fluid,
confined mainly to the wave crest. The displacement begins rather abruptly at the ‘ toe’
of the instability, which is located on the forward face of the unperturbed wave. The
location of the toe can be defined as the point of maximum upwards curvature of the
perturbation profile. This occurs at a horizontal distance у = —2.3 from the centreline.
2097
У
F ig u re 3. Profile of the unstable mode /? = 0.05442 at times t = 0, 10, and 20. The amplitude a
at t = 0 is 0.01.
У
F ig u r e 4. Close-up of the profile in figure 3.
У
F ig u r e 5. As in figure 3, but at times t = 20, 25 and 30, indicating qualitatively the later
development of the instability.
The tangent to the unperturbed surface at this point makes an angle of 21° with the
horizontal.
It will be realized that the present analysis is valid for a small perturbation only. If
the amplitude of the perturbation were increased, the theoretical profile would develop
as in figure 5, so that the crest would overturn. But one criterion for the validity of the
perturbation expansion is that the corresponding perturbation in the surface slope
2099
У
F ig u r e 6. Profiles of the neutrally stable mode a = 0.263 59 at four different phases, (a) a t = 0,
(fc) a t = (с) a t = n, ( d ) a t = 2тс.
7. Discussion
The existence of an inherent instability in the crest of a steep gravity wave has
important implications for the way in which two-dimensional gravity waves actuall>
break. From our results it seems extremely unlikely that a steep wave will attain the
Stokes limiting form; in practice it must topple over before that steepness is attained
In previous theoretical work on periodic, irrotational waves (Tanaka 1983; Longuet-
2 101
and similarly
J Re [Q(w) E(w) - R(w) C(w)) + S(r) F(r),
iC0S"T
smm- J
L (A 4)
21 ‘ —2 ( 1 + cos t) d/ydr,
fRe[/>MC(w)]-i?(T)F(T), *:cosnT), (A5)
21 *|lm [/’(w) C(>v)] sin m J
f —2(1 + cost) d/ydr
^ C0S"r] , (A 6 )
: {Re [Q(w) E{w) - R{w) C(w)] 4 - S(t) F(t), Sin ЛТ J
2102
p</*">
+) i R f
-R<+)\|
\ 0 \ J' J
jl 0 : - j \
A2l — 1 » ^22= 1 ... , (A 8 )
_/*-> : I1 \
T
cb
b
fc
0
1
1
where in the case N = 3, for example,
( (.P0+ P0.) (Р-г + Р- J (P- 2 + P- 2) (Р-3 + Р - Л
(/>.,+Л) (P- 2 +P0 ) (P-3+ P -J (P-, + P~d \ (A9)
о ч б з + е ,) i - ( f i . + e j 2 ( e , + e 2) 3 ( f i .+ e ,) /
Q' is the same as Q but without the first row. Also Я(+) is similar to Я<+) but without
the first column while
/ (Л ..- Л ,) (Л-2 -Л „) (Л -з-Л -,) (Л-4 -Л _ 2)\
« (->= I (Л -.-Л .) (Л -з-Л .) ( * - ,- * „ ) (Л -5-Л .,) • (А 12)
\(Л-з~Лз) (Л .4 - Л 2) (Л_5 - Л ,) (Л_в- Л 0) /
Я denotes Я(_) without the((S_t-st ) ( S Then
first column. .,- S .,)
S (+) (isS similar
. , - S . 2)\
to Я(+>, while
s<-> = (S- 3 - S .) (S . 4 - S 0) (S_s-S _ ,) . (A 13)
\ (5- 4 - 5 2) (S_6 —5,) (S _ ,-S 0) I
Lastly, •/ = I t t 3 1* (A 14)
while J ' is the same as J but without the first row. Note that the first rows of J and
Q are double the values corresponding to the pattern in the matrices as a whole.
From (A 2) one obtains immediately
f i u - B 1v, ftv = B 2u, (A 15)
2103
REFERENCES
B an n er, M. L. & P e regrin e, D. H. 1993 W ave breaking in deep water. Ann. Rev. Fluid Mech. 25,
373-397.
Bonm arin, P. 1989 Geometric properties of deep-water breaking waves. J. Fluid Mech. 209,
405-433.
C le a v e r , R. P. 1981 Instabilities o f surface gravity waves. PhD thesis, University o f Cambridge,
UK, 224 pp.
H o lt h u u s e n , L. H. & H e r b e r s , Т. H. C. 1986 Statistics of breaking waves observed as white-caps
in the open sea. J. Phys. Oceanogr. 16, 290-297.
K je ld s e n , S. P. & M y r h a u g , D. 1980 Wave-wave interactions current-wave interactions and
resulting extreme waves and breaking waves. Proc. 17th Inti Conf. Coastal Engng, pp. 2277-2303.
ASCE.
L o n g u e t - H i g g i n s , M. S. 1978a Some new relations between Stokes’s coefficients in the theory of
gravity waves. J. Inst. Maths Applies. 22, 261-273.
L o n g u e t - H i g g i n s , M. S. 19786 The instabilities of gravity waves of finite amplitude in deep water.
I. Superharmonics. Proc. R. Soc. Lond. A 360, 471-488.
L o n g u e t - H i g g i n s , M. S. 1981 On the overturning of gravity waves. Proc. R. Soc. Lond. A 376,
377-400.
L o n g u e t - H i g g i n s , M. S. 1986 Bifurcation and instability in gravity waves. Proc. R. Soc. Lond. A
403, 167-187.
L o n g u e t - H i g g i n s , M. S. 1988 Mechanisms of wave breaking in deep water. In Sea Surface Sound
(ed. B. R. Kerman), pp. 1-30. Kluwer. 639 pp.
L onguet-H iggins, M . S., C leaver, R. P. & Fox, M . J. H. 1994 Crest instabilities of gravity waves.
Part 2. Matching and asymptotic expansion. J. Fluid Mech. 259, 333-344.
M. S. & C o k e l e t , E. D. 1976 The deformation of steep gravity waves. I. A
L o n g u e t - H ig g in s ,
numerical method of computation. Proc. R. Soc. Lond. A 350, 1-26.
L onguet-H iggins, M . S. & C okelet, E. D. 1978 The deformation of steep gravity waves on water.
II. Growth of normal-mode instabilities. Proc. R. Soc. Lond. A 364, 1-28.
M .S. & Fox, M.J . H. 1977 Theory of the almost-highest wave: the inner
L o n g u e t - H ig g in s ,
solution. J. Fluid Mech. 80, 721-741 (referred to herein as LHF1).
L o n g u et-H ig g in s, M . S. & Fox, M . J. H. 1978 Theory of the almost-highest wave. Part 2. M atching
and analytic extension. J. Fluid Mech. 85, 769-786.
M i c h e l l , J. H. 1893 The highest waves in water. Phil. Mag. 36, 430-437.
T a n a k a M. 1983 The stability of steep gravity waves. J. Phys. Soc. Japan 52, 3047-3055.
T u l i n , M. P . & Li, J. J. 1991 A mechanism for wave deformation and breaking intermediated by
resonant side-bands. Proc. IUTAM Symp. on Breaking Waves, Sydney, Australia July 1991 (ed.
M. L. Banner & R. H. J. Grimshaw), pp. 21-37. Springer.
2104
1. Introduction
The present paper is a sequel to a recent investigation (Longuet-Higgins & Cleaver
1994, referred to herein as LHC) in which the problem of wave breaking was tackled
by considering the stability of the flow in an ‘almost-highest wave’, that is to say the
asymptotic form assumed by the crest of a steep, irrotational gravity wave as the
limiting Stokes corner flow is approached, but while the crest is still rounded (see
Longuet-Higgins & Fox 1977, referred to herein as LHF1). It was found that there
exists just one unstable normal-mode perturbation of this flow, with a form resembling
the initial stages of a spilling breaker.
The question then arises: how is this local instability of a gravity wave crest related
to the known unstable modes of perturbation of a progressive gravity wave? For
example, it was shown numerically by Tanaka (1983) and Longuet-Higgins (1986) and
analytically by Saffman (1985), that the lowest superharmonic perturbation becomes
unstable at a value of the wave steepness parameter ak = 0.4292, corresponding to the
first maximum in the wave energy density as a function of ak. Is this superharmonic
instability essentially the result of the local flow near the wave crest?
Now the ‘almost-highest wave’ considered in LHF1 and LHC is the inner flow near
the wave crest, in terms of a small parameter e which measures departure of the flow
from an ideal Stokes corner flow. Thus we may define
e = q J2 c, 0-1)
2105
T tanhf = 2 7 2 - <2'4>
viz. fi = 0.7143— As я ,-* 0 0 , so clearly /?->0 and
*1 ~ (§i*A (2.5)
representing the Stokes 120° comer-flow.
On the other hand in Zone III we know from LHF2 (see also Grant 1973) that in
the limit e -> 0 (a sharp-crested, limiting wave), and as ^ - > 0 (the neighbourhood of the
sharp crest), so
2~ (ii*)J[l+ 7(i*)A- 1], (2 .6 )
2106
F ig u re 1. (after LH F2). Sketch showing regions of validity for the inner solution (Zone I), outer
solution (Zone III) and matching of the two solutions (Zone II).
(43)
and from (3.2)
dz = ( J + i ^ « N + ? ( i + i * 1)*-S» M , (4.4)
d(iXi)
where Н(ш) = Щш)-(1+ш)В'(ш) (4.5)
as in LHC §2, while _ _
H(w) = (Л- i) B(to) - H(w) B'(w). (4.6)
Thus from (3.2) and (4.2) we have
z, = Дг( 1+ <o)~>B(ca) + r/A2i-\ I + w)-J-i B(<o), )
dz , 1 (4.7)
- i p - = д - 4 1 + «)! а д + , “ - , (1 +«)-•'-> HM,
dXi J
and A = (2S)\
It is convenient to write the boundary condition for z1 as
<4-*>
where c.c. denotes the complex-conjugate expression. Then on substituting for z, and
dzj/d^, and considering the terms in i/° and t}' respectively we obtain
ш-'>В(ш)Н(<о)Н(а>-1)+ c.c. = 1, (4.9)
and (1 + a>y~* ot~* B(tit) H(o>) H(a>~l)+c.c.
+ (1 + ш)‘-дйГ! B(w) Н(ш) H(aj~')+c.c.
+ (1 + <o)'-A« Н B(u) Н{ш) Й(оГ') + С .С. = 0. (4.10)
On extracting the real factor
[ (l+ w )( l+ & rl)]i(,_A> (4.11)
2108
z
S
n N = 50
II
о
и
0 0.047765 0.047761 0.047758
1 -0 .0 7 0 7 1 1 -0 .0 7 0 7 0 6 -0 .0 7 0 7 0 2
2 0.014920 0.014921 0.014921
3 0.003529 0.003531 0.003532
4 0.001474 0.001475 0.001476
5 0.000858 0.000858 0.000859
6 0.000570 0.000571 0.000571
7 0.000371 0.000371 0.000372
8 0.000292 0.000292 0.000293
9 0.000192 0.000192 0.000193
10 0.000167 0.000167 0.000167
U 0.000108 0.000108 0.000108
12 0.000101 0.000101 0.000101
13 0.000063 0.000062 0.000062
14 0.000064 0.000064 0.000064
15 0.000037 0.000037 0.000037
16 0.000041 0.000041 0.000041
17 0.000022 0.000022 0.000022
18 0.000028 0.000027 0.000027
19 0.000013 0.000013 0.000013
T able 1. Coefficients B n in the expansion o f the modified inner solution, equations (3.2) and (3.4)
У\
У
F ig u r e 2. Free-surface profiles for the cases e = 0, 0.1, 0.2 and 0.3: (a) on the scale of
zx = e*z and (b) on the scale of z.
The corresponding surface profiles are shown in figure 2 for e = 0, 0.1, 0.2 and 0.3.
In figure 2(a) the profiles are plotted with the scaled coordinate zx = e- 2z. In figure 2(b)
they are plotted against z. It should be borne in mind that these represent only the inner
solution near the crest of the wave, so that in figure 2 (6 ) the region of validity is only
within a distance 0(e2) from the wave crest.
The relation between the parameter e, defined by equation (1.1), and the steepness
2110
ak
F igure 3. R elation between the param eter e and the wave steepness ak. The broken
line shows the asym ptote (4.16).
(5.5)
(5.8)
where
(5.9)
6( oj) = f i ( a O - f ( l + « ) # ( « ) - ( 1 + w ) 2J > ) ,
and a prime denotes d/dw. Similarly if
(5.10)
then C, = -iA -3(I + «)£(«), (5.11)
where £(w) = (l+ w )C /(w). (5.12)
then on substituting in equation (5.4) we obtain
to be satisfied on
In equation (5.14), P, Q, R, S are precisely the same functionals of B^H and G as they
are of By H and G in equation (5.16) of LHC1. Moreover, since B(o)), like B(oj), is
expansible in positive powers of w, all the remainder of §5 of LHC1 applies also. Thus
we may seek normal-mode solutions in which £ and F have a time dependence like e^,
and proceed to determine fi in a similar manner.
6. Results
Calculations of the eigenvalues were carried out with 8= 10* as in LHC1 and e in
the range 0 to 0.105. It was found that all the eigenvalues fi2were real. Moreover, only
the lowest eigenvalue was positive, corresponding to an exponential growth or decay
rate /?0. In figure 4 is plotted as a function of e, and will be seen to be a steadily
decreasing function of e, diminishing from = 0.00292 at e = 0 to zero at about
e = 0.102, apparently. The computation is valid, however, only for sufficiently small
values of e.
It is more revealing to consider the unsealed growth rate fi/ae, relative to the radian
frequency o' of a progressive deep-water wave of low amplitude (o'2 = gk, where к is the
wavenumber). Figure 5 shows ( f i j e o ) 2 plotted against the wave steepness ak. The
curve on the upper right of the diagram corresponds to that in figure 4, the same
portions being shown by solid or broken curves. The remaining plots in figure 5 are of
the frequencies of the lowest (non-trivial) superharmonic mode of a progressive wave
of finite amplitude, as calculated by Tanaka (1983) and Longuet-Higgins (1978, 1986).
The quantity —a\/<r2 is plotted from figure 7 of Tanaka (1983) and table 2(b) of
Longuet-Higgins (1986). These calculations demonstrated that the lowest mode
becomes unstable at around ak —0.4292, where the energy density has a local
maximum (see also Saffman 1985). This point is marked by a cross (x ) in figures 4
and 5.
2112
e
F ig u r e 4. The rate of growth /?0 of the unstable mode, is shown as a function of e.
ak
F ig u r e 5. Growth rates of the crest instability compared with those of the lowest superharmonic
instability in a progressive, irrotational wave in deep water: О ------О O V 0”)2» from ^ 8ure 4 >
• ----- • —(<r,/<r)2 from Longuet-Higgins (1978, 1986); +------ b —{vjo)* from Tanaka (1983).
From figure 5 it appears that the calculations of the present paper provide an
asymptote for the previous calculations of the lowest superharmonic instability, in the
limit as e->-0 and ak-^(ak)m&x. In other words, the superharmonic instability is
associated primarily with the instability of the flow near the wave crest; it is a ‘crest
instability’.
2113
У\
F ig u r e 6. Form of the free-surface profile in the growing instability, when e = 0.08.
Secondly, the asymptotic analysis, so far as it has been carried in the present paper,
is valid only for sufficiently small values of e (as one would expect). Figure 5 suggests
that the limit of validity is at about e = 0.08 (ak = 0.440). More accurate calculations
of this instability from the present point of view would have to take account of the flow
not only in the inner zone I (see §2 above) but also in the matching zone II and hence
also the flow in zone III. This would imply corrections depending upon higher powers
of e.
The displacement of the free surface corresponding to the growing unstable mode is
shown in figure 6 , for the case e = 0.08. It will be seen to be practically identical in form
to the case e = 0, shown in figures 3 and 4 of LHC. That is, the fluid near the crest is
displaced forwards, and the ‘ toe’ of the breaker occurs at about y x = —2.3. The chief
discernible difference is in the rate of growth /?0, which is 0.0544 when e = 0 and 0.0421
when e = 0.08. According to figure 4, fil decreases almost parabolically with e, so that
$ = 0.00296-О.ОЗОе2. (6.1)
Hence in the unsealed time frame the rate of growth /?' = fio/ea relative to the radian
frequency cr of a progressive wave in deep water (and of small slope) may be written
P = 0.0544e_1(l -0.05e2), (6.2)
generalizing equation (6.5) of LHC. This is valid only when 0 ^ e ^ 0.08. Since over
this range e2 a 2Aak, where Aak is the difference between the steepness parameter ak
and its limiting value 0.4432, we have also
P % 0.0385(Aafc)“* (l-OAOAak) (6.3)
over the same range.
Tanaka (1983) does not describe the surface profile of the instability in his
calculations, so that changes in the form of the instability in the range
0.4292 < ak< 0.44 cannot be ruled out.
2114
R E FE R E N C E S
C le a ve r ,R. P. 1981 Instabilities of surface gravity waves. PhD thesis, University of Cambridge.
D o l d , J. W. & P eregrine , D . H. 1987 Water-wave modulation. In Proc. 20th Inti Conf. on Coastal
Engng, Taipeh, Nov. 1986, pp. 163-175. ASCE.
Fox, M .J .H . 1977 Nonlinear effects in surface gravity waves. PhD. thesis, University of
Cambridge.
G r a n t , M. A. 1973 The singularity at the crest of a finite amplitude Stokes wave. J. Fluid Mech. 59,
257-262.
H o l t h u ijse n , L . H . & H e rbe rs , Т . H . C . 1986 S ta tis tic s o f b r e a k in g w a v es o b serv ed a s w h ite-ca p s
in th e o p en se a . J. Phys. Oceanogr. 16, 290-297.
L o n g u et -H ig g in s , M. S. 1986 Bifurcation and instability in gravity waves. Proc. R. Soc. Lond. A
403, 167-187.
L o n g u et -H ig g in s , M. S. 1993 The crest instability of steep gravity waves or How do short waves
break? Proc. Air-Sea Interface Symp. June 1993, Marseille, France (ed. M. A. Donelan) (to
appear).
L o n g u et -H ig g in s , M. S. & C le ave r , R. P. 1994 Crest instabilities of gravity waves. Part I. The
almost-highest wave. J. Fluid Mech. 258, 115-129 (referred to herein as LHC).
L o n g u et - H ig g in s , M. S. & C o kelet , E. D. 1978 The deformation of steep surface waves on water.
II. Growth of normal-mode instabilities. Proc. R. Soc. Lond. A 364, 1-38.
L o n g u et-H ig g in s, M .S . & Fox, M .J.H . 1977 Theory of the almost-highest wave: the inner
solution. J. Fluid Mech. 80, 721-741 (referred to herein as LHF1).
L o n g u et - H ig g in s , M .S . & Fox, M . J . H . 1978 Theory of the almost-highest wave. Part 2.
Matching and analytic extension. J. Fluid Mech. 85, 769-786 (referred to herein as LHF2).
S a f f m a n , P. G. 1985 The superharmonic instability of finite amplitude water waves. J. Fluid Mech.
159, 169-174.
T a n a k a , M. 1983 The stability of steep gravity waves. J. Phys. Soc. Japan 52, 3047-3055.
T a n a k a , M. 1985 The stability of steep gravity waves. Part 2. J. Fluid Mech. 156, 281-289.
2116
1834 J O U R N A L OF P H Y S I C A L O C E A N O G R A P H Y V o l u m e 24
ABSTRACT
Fractal models of breaking waves in a random surface should preferably describe dynamical as well as geo
metrical properties. This becomes feasible if there is a wide separation between the length scales of component
waves. Using this idea, a simple model of breaking waves is constructed, which shows that whereas the downward
acceleration of particles at a wave crest is limited to g, the upward accelerations in a wave trough are unbounded.
Owing to tangential stretching or contraction, certain phases of a progressive or standing wave can be identified
as being stable or unstable. The most striking instabilities are expected on the forward slopes of progressive
waves, and in the troughs of steep waves meeting a vertical wail.
In a p rogressive wave (see Fig. 5) the surface is con done by radiation stress, and reduction in the effective
tracting on the forw ard face of the wave, and stretching gravity. All these effects will be cumulatively greater
horizontally on the rear face. Thus, the forward face on the forward face of the longest waves.
of the wave is unstable to short-wave disturbances, and
the rear face is stable. This is reflected in the common 6. Surface stability in standing waves
observation that the forward face of a steep wind wave
is generally rough, while the rear face is smooth. Consider on the other hand a standing surface wave
The wavelength of the short waves is also affected (Fig. 7). In such a wave the amplitude of the crests
by the contraction of the free surface, and the amplitude and troughs are either increasing simultaneously as in
of the short waves is further increased by the relatively
smaller value of the apparent gravity g' on the upper
part of the long waves (Longuet-Higgins and Stewart
1960).
The extent of the short-wave steepening on long
waves of finite amplitude is, in fact, greater than would .stable
be expected on a linear theory for the long waves. Exact
calculations (see Fig. 6 ) show that even on a long wave
with steepness parameter AK = 0.3, the short waves
steepen between the long-wave trough and long-wave
crest by a factor as great as 4. When AK = 0.4 the
steepening factor exceeds 10. b
Thus, it is most probable that the steepest short waves
will be found near the crests of the long waves, and unstable unstable
preferably on their forward face. For, if the short waves
break ahead of the long-wave crests, they will not regain
enough energy from the wind to break on the rear face
(cf. Longuet-Higgins 1991). unstable
Still shorter waves riding on the short waves will be Flo. 7. Stable and unstable phases in a standing gravity wave: (a)
subject to the similar processes of steepening by con crests and troughs increasing in amplitude; (b) crests and troughs
traction of the free surface due to the short waves, work diminishing.
2120
1838 J O U R N A L OF P H Y S I C A L O C E A N O G R A P H Y V o lu m e 24
Fig. 7a or decreasing simultaneously (Fig. 7b). In the enon. But even there the local contraction of the free
first case the free surface is everywhere being stretched. surface in the unperturbed flow is a significant factor
Hence, it is everywhere stable to short-wavelength per in the instability.
turbations. In the second case the free surface is every A satisfying fractal model of a breaking sea that fully
where contracting and so it is unstable to short-wave- incorporates the dynamics of a free surface under the
length perturbations. The latter case, when the wave action of gravity still awaits development. The present
trough is rising, corresponds to the situations found by paper draws attention to this challenge and makes a
Cooker and Peregrine (1991) to give rise to very high modest start.
upward accelerations.
REFERENCES
Banner, M. L., and D. H. Peregrine, 1993: Wave breaking in deep
7. Discussion water. Annu. Rev. Fluid Mech.. 25, 313-397.
Cooker, M. J., and D. H. Peregrine, 1991: Violent motion as near-
We have shown how, by incorporating dynamical breaking waves meet a wall. Proc. IUTAM Symp. on Breaking
considerations into a simple fractal model of surface Waves. Sydney, Australia, Springer, 291-297.
waves it is possible to understand some observed fea Huang, N. E., S. R. Long, С.-С. Tung. M. A. Donelan, Y. Yuan,
and R. J. Lai, 1992: The local properties of ocean surface waves
tures of the sea surface. The model is highly simplistic by the phase-time method. Geophys. Res. Lett.. 19, 685-688.
and is made possible only by the assumption of a wide Kerman, B. R., 1990: Fractal geometry and sound generation of a
separation of scales between the component waves. wind-swept sea. Proc Conf. on Natural Physical Sources of Un
This enables us to assume that the short waves are in derwater Sound. Cambridge, England, KJuwer Academic, 467-
a sense controlled by the longer waves on which they 482.
------ , 1993: A multifractal equivalent of the Beaufort scale for sea-
ride and to ignore the reaction of the short waves on state. Geophys. Res. Lett.. 20, 297-300.
the much more energetic long waves. Longuet-Higgins, M. S., 1985: Accelerations in steep gravity waves.
Nevertheless it is worth noting that the short-wave J. Phys. Oceanogr , 15, 1570-1579.
steepening shown in Fig. 6, for example, is nearly in ------ , 1987: The propagation of short surface waves on longer gravity
waves. J. Fluid Mech.. 177, 293-306.
dependent of the wavelength of the short waves when ------ , 1988: Mechanisms of wave breaking in deep water. Sea Surface
the ratio of the wavelengths exceeds about 2 ( see Lon Sound. B. R. Kerman, Ed., Kluwer Academic, 1-30.
guet-Higgins 1987, Figs. 4 and 5). Thus the aspects of ------ , 1991: A stochastic model of sea surface roughness. II. Proc.
wave steepening discussed here may not be very de Roy. Soc. London A, 435, 405-422.
------ , and R. W. Stewart, 1960: Changes in the fonn of short gravity
pendent on the assumption concerning the component waves on long waves and tidal currents. J. Fluid Mech.. 10,
length scales. 529-549.
It should also be noted that the horizontal contrac ------ , and M. J. H. Fox, 1977: Theory of the al most-highest wave:
tion, which destabilizes certain portions of the free sur The inner solution. J. Fluid Mech.. 80, 721-741.
face, is merely the mechanism that initiates the break ------ , and R. P. Cleaver, 1994: Crest instabilities of steep gravity
waves. Part 1. The inner solution. J. Fluid Mech., 258, 115-
ing. Another such mechanism is the instability of the 129.
crest of a steep wave ( see Longuet-Higgins and Geaver Phillips, О. М., 1958: The equilibrium range in the spectrum of wind
1994). It appears the latter is mainly a local phenom generated waves. J. Fluid Mech.. 4, 426-434.
2121
2458 J O U R N A L OF P H Y S I C A L O C E A N O G R A P H Y Volume 25
Reprinted from J ournal of Physical Oceanography. Vol. 25, No. 10, October 1995
A m m a n MnextJopoJ Society
ABSTRACT
An inviscid mechanism is proposed foe the breakup of the jet in a plunging surface wave. Streamwise per
turbations of the original surface are shown to grow rapidly owing to stretching of the thin jet and to drastic
reduction in the normal pressure gradient. This converts transverse gravity waves into almost pure capillary
waves. Conservation of wave action for the perturbations then implies a strong increase in the perturbation
amplitude.
Fig. I. Successive profiles of a plunging breaker in water of finite depth, calculated by numerical time stepping.
(After Vinje and Brevig 1981.)
(b)
F ig . 2. Sketch o f short transverse waves on a plunging breaker, Fig. 3. Sketch of waves on a thin sheet of fluid: (a) symmetric and
at an early stage o f development. (b) antisymmetric (after Taylor 1959).
2123
2460 J O U R N A L OF P H Y S I C A L O C E A N O G R A P H Y V o lu m e 25
ф= aa and
cosh kz sin (ky - at) (3.1)
к sin bkh E = 2V = T(ak)2. (3.10)
in our notation, z = 0 being the central plane. The dis The action density A is then given by Eta. For the
persion relation is symmetric waves this leads to
Tky tanh kh (3.2) A = (T lh ),/2a 2. (3.11)
as in pure capillary waves on water of finite depth A.
In case (b) the displacements on each side are of the We then have two relations, derived from the conser
sam e sign, corresponding to a velocity potential vation of mass and of wave action, respectively. From
mass conservation
фш sinh kz sin(ky - at) (3.3) hAs = const, (3.12)
к cosh kh
and from action conservation
and dispersion relation
( a 2/hU2)As = consL (3.13)
a 2 = Tk* cosh kh. (3.4)
On dividing (3.13) by (3.12) we obtain
It is obvious that an initial perturbation confined to
one side of the jet will be resolved into the sum of a a 2/h: const, (3.14)
symmetric and an antisymmetric perturbation, which
then will be propagated independently, each with its
own phase and group velocity. a a /r (3.15)
When the sheet is thin compared to the length of the
perturbation, that is, kh < 1, the above expressions are
greatly simplified. For the symmetric perturbations we
have
aa .
= j^ s m i k y - a t)
(3.5)
= Thk*
giving
с = (T lh )m . (3.8)
These waves are therefore nondispersive and, in fact, (b)
(b)
Fig. 5. Sketch showing the change in orientation o f obliquely propagating waves on a sheet o f water,
when the sheet is stretched in the x direction.
2462 J O U R N A L OF P H Y S I C A L O C E A N O G R A P H Y V o lu m e 25
237
TH E C R E S T IN ST A B IL IT Y OF ST E E P G R A V IT Y W A V E S,
O R H O W DO SHO RT W AVES B R E A K ?
M ic h ae l S . L o n g u e t-H ig g in s
I n stitu te f o r N on lin ear S cie n ce , U n iversity o f C aliforn ia, San D iego, La J o lla , C a liforn ia
92093-0402, U.S.A.
A b s t r a c t . T h e local, normal-mode perturbations of the crest of a steep g rav ity wave (an
“alm o st-h igh est” wave) have been calculated . T here is ap paren tly only one un stab le m ode,
which is sta tio n ary and has an exponential grow th-rate 0 = 0.0544, on the local tim e-scale. In
th is in stab ility the fluid is displaced forwards, creatin g a region of sh arp (concave) cu rv atu re,
or “toe” , on the forward face of the wave.
T h e in sta b ility is shown num erically to be an asym ptotic form of the low est un stab le
superh arm on ic in stab ility of a finite-am plitude Stokes wave.
It is suggested th at in short gravity waves (w avelength 5 to 50 cm ) th is in sta b ility is the
source of p arasitic c ap illary waves, which originate a t the “toe” of the in stab ility rath er than at
the wave crest. T he p arasitic capillaries produce strong vorticity, leading to flow sep aratio n and
the generation of a roller beneath the gravity-w ave crest. T he roller then becomes turb ulent
and the crest crum ples. T h is prediction is supported by recent labo rato ry observations d ue to
Professor J.H . Duncan.
1 Introd uction
A steep g ra v ity w ave, w ith wavelength in the range 5.0 cm to 50 cm ap p ro x i
m ately, is com m only observed to support a train o f “p arasitic” c a p illa ry w aves
(C o x , 19 5 8 ). T h e capillaries not only contribute stron gly to the d am p in g o f
th e g ra v ity w ave (see Longuet-Higgins, 19 6 3 ) b u t in so doing th e y g en erate
enough v o rtic ity to contribute su b stan tially to the vo rtex o r “ro lle r” which is
g en erally found beneath the crest o f such waves; see Ebuchi, K a w am u ra and
T ob a (1 9 8 7 ), Longuet-Higgins (19 9 2 ). In early theories o f p arasitic cap illaries
(Longuet-H iggins, 1963; C rap p er, 1970) it was supposed th a t the source o f the
c ap illa ry waves was the sharp curvature at the crest o f the g ra v ity w ave. B u t
close observation suggests th at the capillaries originate from some p oint on the
forw ard face o f the wave, rath er than at the crest itself. W h y is th is so?
To answ er this question, we have investigated carefully the form o f th e in sta
b ility o f th e crest o f a very steep g ravity wave. In a recent pap er by L onguet-
Higgins and C leaver (19 9 3 ), here referred to as LHC, we took as o u r m odel
th e crest o f an “alm ost-highest wave” (see Fig. 1 ), whose asym p to tic form was
first found by Longuet-Higgins and Fox (19 7 7 ), (or LH F 1 ), and we found the
n orm al-m ode pertu rb ation s of this flow, th a t is to say those p ro p o rtio n a l to
etct, where a is the radian frequency and t is the tim e. A m ong the sp ectru m
o f such oscillations we found ju s t one mode w ith an im aginary frequ en cy cry
th a t is w ith a positive exponential growth having /3 = - io > 0. T h e form o f
the in stab ility is shown in Fig. 2, and it will be seen th at it does indeed have
a sh arply concave “toe” on the forw ard face o f the crest, which could becom e
the source o f cap illary waves. (These m ust ride ahead o f the source, since th e ir
group velocity exceeds th eir phase velocity; see Lamb, 19 3 2). An o u tlin e o f the
calculations leading to this result is given in Sections 2 to 4 below.
238 The Crest Instability of Steep Gravity Waves, or How Do Short Waves Break?
Fig. 1 . (from L H F l). Profile of the alm ost-highest wave, showing coordinates and 30°
asym ptotes.
» +5?2 =o (1 )
precisely. (The acceleration o f gravity g is taken to be u nity). If ф and ф
denote the velocity potential and the stream -function in the steady flow and if
X = ф + хфу z - i(x + iy), then equation ( 1 ) becomes
(z + z-)zx (zxy = l (2 )
у = - 1 , q = q0 = V2 (4)
To calculate the form of the inner flow, LHF1 made the transformation
1 -w 6 - i\ /,ч
= <5—— , и = - (5)
1 +w <5+ *X
where 6 is a suitable constant. This maps the flow onto the interior of the circle
|u;| = 1. Then the solution was expressed as
« = (< + ! # % ) (6)
where
B(u) = Bo + B\ lj + J?2 ^ Hh ■ (7)
The coefficients Bn (which are real, by symmetry) were determined by substi
tuting ( 6 ) into the boundary condition (2) and equating coefficients ofu;n. This,
together with the necessary condition that
where Z = i(X + iY) denotes the unperturbed flow found in Section 2 and
С = *(£ + irj) is a small perturbation, LHC show that the kinematical and the
2129
2 40 The Crest Instability of Steep Gravity Waves, or How Do Short Waves Break?
dynam ical boundary conditions can be combined as the real and im aginary parts
o f the single equation
to be satisfied on Im (x) = 0 .
Now z = Z is alread y expanded in a known power series in u ; see equation ( 6 ).
T his can be expanded in a Fourier series on the unit circle и = e t r , -тг < r < гг.
W e m ay assume sim ilarly the expansions
( = E ~ o(4-. + '« n K
(11)
/ = EST=o (cn cos nT + <*» Sin nr)
w here the coefficients a n , 6n , c„ and dn are functions o f the tim e t only. Then
su b stitu tin g in equation ( 10 ) and equating coefficients o f cos n r and sin n r in
both real and im aginary parts we obtain a system o f equations (with real coef
ficients) for the vector
To seek norm al-m ode instabilities we w rite d/dt = (3 and hence obtain a
m atrix equation o f the form
M x = Bx (13 )
w here A and В are infinite m atrices. These can be solved num erically by setting
B n = 0 for n > N , say, and also an = bn = c„ = dn = 0 when n > jV, and
then choosing the lowest (AN + 2 ) equations to determ ine the eigenfrequencies
<jj = i/3j and the corresponding eigenvectors X j .
A n im portant im provem ent in accuracy is attained by noting th at the form
o f A is such th at the vector x can be split into two parts:
u = (ао,вь
(1 4 )
v = ( 6o , 6b .. . 6# ; cb c2l ...c;v)
P A U u = A 12v }
(15)
(3A2i v = A 22u J
w here Д ц , etc. are subm atrices of A. Then we have
pu = B jv , B i = А 7/ A j2
(16)
p\ - B 2u, B 2 = A ji1 A 22
Michael S. Longuet-Higgins 241
and hence
^ 2u = B i B 2u (17)
T his reduces th e ord er o f the m atrices by one-half, for a given value o f N , and
leads to a considerable im provem ent in accuracy for a given available m achine
tim e. F u rth er details o f the calculation are given in LHC.
242 The Crest Instability of Steep Gravity Waves, or How Do Short Waves Break?
the flow becomes highly nonlinear, so th at subsequent profiles in Figure 2 are
at best q u alitative. In fact from previous studies of breaking gravity waves
(Longuet-Higgins and Cokelet, 1978) we expect to see the eruption o f a strong
je t from th e tip of the overturning wave - unless the scale is so sm all th at the je t
is contained by surface tension. It is clear th a t the sharp concave curvature at
the toe o f the instability, which precedes the overhanging je t, is a likely source
for the generation o f cap illary waves when surface tension is taken into account.
T his is m ore in accord w ith observation (see below) than previous theories in
which it was assumed th a t the source o f parasitic capillary waves was the sharp
cu rvatu re a t the wave crest itself.
W e note th at by reversing the sign o f each component in the eigenvector x
in equation ( 12 ), and w ith the same value o f /?, we obtain a pertu rb ation which
p rojects m ainly from the rear face of the wave. This is less frequently observed.
A possible explanation is th at the original stim ulus for the p ertu rb ation is likely
to come from shorter waves which are swept backwards up the forw ard face
o f the advancing wave, and are increased in steepness by the o rb ital straining
combined with the radiation stress. These disturbances will excite the instability
on the forw ard face. By contrast, on the rear face straining tends to reduce the
steepness o f any shorter, superposed waves. Thus the sign o f the eigenvector x
is m ore likely to be determ ined by disturbances on the forward face than on the
rear face. Once the sign o f x is determ ined, the instability will grow accordingly.
On the present scaling, the ratio of the length-scale to th at of a gravity wave
o f wavenum ber к = 1 and wavelength 2тг is equal to ( 2, w here e is the param eter
defined by equation (3). The tim e-scales are in the ratio c. Thus the rate of
grow th /3 to the radian frequency a of a sm all-am plitude wave in deep w ater
(<72 = gk) is given by
P!<j = 0.0544 c"1 (IS)
This shows th at in very steep waves (if th ey are attain ab le) the grow th-rate
o f the instab ility becomes infinite like (0.4432 - ak)~1/2. Thus it appears most
unlikely th a t a steep gravity wave will ever attain the lim iting Stokes form w ith
a sharp 120° corner. In practice it must topple over, probably forw ards, before
the lim iting form is attain ed .
ak
Fig. 3. T he dim ensionless grow th-rate 0 ' = 0/c<r of the unstable mode on the outer scale. 0 *
is plotted as a function of the wave steepness ak. Also shown are the values of w here
<72 is the radian frequency of the lowest normal-mode perturbation of a deep-w ater Stokes wave
of finite am p litude: ( • ) from Longuet-H iggins (1978, 1986), (+ ) from Tanaka (1983).
crest. In LH CF we calculated the norm al-m ode p e rtu rb ation s o f this corrected
flow. It turns out th at the slightly less steep waves are still unstable, b u t w ith a
reduced rate o f grow th /3. M oreover the rate o f grow th can be com pared to the
rate o f grow th o f the lowest unstable (superharm onic) mode o f a fin ite-am p litu d e
Stokes wave, which was calculated by Longuet-Higgins (1 9 7 8 , 1986) and T anaka
(19 8 3 ). This shows th a t the crest in stab ility is sim ply an asym p totic form o f
the la tte r in stab ility; see Figure 3. T he norm al m ode retu rns to s ta b ility when
the wave steepness ak is less than 0.4292.
6 D iscussion
W e have shown th a t for very steep Stokes waves, w ith ak > 0.40, th ere is ju s t
one local in stab ility o f the wave crest, and th at it is a lim iting form o f the
superharm onic in stab ility discovered earlier by Tanaka (198 3 ). In its lim itin g
form , the in stab ility produces a ra th e r sharp concavity in the free surface on the
forw ard face o f the wave. In large-scale gravity waves the in stab ility p ro b ab ly
develops a forwards je t, which falls forw ard and traps air, as in the in itia l stage
of a spilling or plunging breaker.
2133
244 The Crest Instability of Steep Gravity Waves, or How Do Short Waves Break?
W e have of course assumed ra th er special initial conditions, nam ely the weak
p ertu rb ation o f a perfectly periodic train of progressive waves. We may expect
a sim ilar crest in stab ility to be found in a wider class o f non-periodic waves.
W e emphasize, however, th at this is alm ost certainly not the only type o f in sta
bility which will lead to wave breaking, since in random seas gravity waves are
com m only observed to break at much lower wave steepnesses; see for exam ple
Holthuijsen and Herbers (1986).
From an oceanographic point o f view the most interesting consequences o f
th e crest in stab ility axe likely to be for very short g ravity waves, with wave-
lengths between 5 and 50 cm, say. In these, the instability may be prevented
by surface tension from developing a sharp-pointed je t. Instead, the sharply
concave curvature at the toe o f the instab ility will become a point of origin
for capillary waves propagating down the forward face o f the wave: the well-
known “parasitic capillaries 4 (Cox, 1958). Such capillaries are known to damp
th e gravity wave rapidly, through th eir increased dissipation of energy due to
th eir short wavelength (Longuet-Higgins, 1963, Longuet-Higgins and S tew art,
19 6 4). A more interesting consequence has been dem onstrated by Longuet-
Higgins (199 2 ), nam ely th at the parasitic capillaries will shed enough vorticity
(associated with the energy dissipation) to produce a vo rtex, or “capillary ro ller”
in the crest of the gravity wave. This roller not only helps to sustain the capillary
waves, but it gives rise to turbulent motion in the wave crest, with consequent
increase in the tran sfer o f heat and gases between ocean and atm osphere. It
m ay also happen th at the roller is so strong as to advance upon the capillaries
and overtake them , producing a quasi-steady flow as at the toe o f a spilling
breaker, where air is entrained in term itten tly in the form of bubbles (see Koga,
1982; Longuet-Higgins, 1990).
The sequence o f events — from the developm ent o f the instability, to the
form ation of the parasitic capillaries, to the shedding o f vo rticity and form ation
o f the roller, to the generat ion o f turbulence in the wave crest — m ust take place
very rapidly, and be fully visible only with the aid o f high-speed photography.
5 cm 5 cm
5 cm
8 Conclusion
W e have described a mechanism, inferred from both an alysis and la b o ra to ry
observations, whereby short g ravity waves can essentially d isintegrate. T h is
form o f breaking is to be distinguished both from the spilling or plunging o f
large-scale gravity waves and from the pinching o f air bubbles by the trou g h s
of very short capillary waves. A descriptive name for this in term ed iate form o f
breaking would be “crum pling” . B y generating turbulence in the crests o f sh o rt
gravity waves it may contribute significantly to the transfers o f heat and gases
2135
246 The Crest Instability of Steep Gravity Waves, or How Do Short Waves Break?
between ocean and atmosphere at low and moderate wind-speeds.
This work has been supported by the Office of Naval Research under Contract
N0014-91-J 1582.
References
C rapper, G. D.: 1970, ‘Nonlinear capillary waves generated by steep gravity waves’, J. Fluid
M ech. 40, 149-159.
Ebuchi, N., Kawamura, H. and Toba, Y.: 1987, ‘Fine structure of laboratory wind-wave
surfaces studied using an optical method’, B oun dary-L ayer M et. 39, 133-151.
Holthuisen, L. H. and Herbers, Т. H. C.: 1986, ‘S tatistics of breaking waves observed as
w hite-caps in the open sea’, J. P hys. O ceanogr. 16, 290-297.
Koga, М.: 1982, ‘Bubble entrainm ent in breaking wind waves’, Tellus 34, 481-489.
Lamb, H.: 1932, H ydrodynam ics, 6th ed., Cambridge Univ. Press, 738 pp.
Longuet-Higgins, M. S.: 1963, ‘The generation of capillary waves by steep gravity waves’, J.
F luid M ech. 16, 138-159.
Longuet-Higgins, M. S.: 1978, ‘The instabilities of gravity waves of finite am plitude in deep
w ater, I. Superharm onics’, P roc. R. Soc. Lond. A 36 0 , 489-505.
Longuet-Higgins, M. S.: 1986, ‘Bifurcation and instab ility in gravity waves’, P roc. R. S oc.
Lond. A. 403, 167-187.
Longuet-Higgins, M. S.: 1990, ‘ Flow separation near the crests of short gravity waves’, J. P hys.
O ceanogr. 20 , 595-599.
Longuet-Higgins, M. S.:1992, ‘ C apillary rollers and bores’, J. Fluid M ech. 2 4 0 , 659-679.
Longuet-Higgins, M. S. and Cleaver, R.P.: 1994, ‘Crest instabilities of gravity waves, I. The
inner solution’, J. Fluid M ech. 258, 115-129.
Longuet-Higgins, M. S., Cleaver, R. P. and Fox, M. J. H.: 1994, ‘Crest in stab ilities of gravity
waves. II. M atching and asym ptotic expansion’, J. Fluid M ech. 2 5 9 , 333-344.
Longuet-Higgins, M. S. and Cokelet, E. D.: 1978, ‘The deformation of steep surface waves on
w ater. II. Growth of normal-mode in stab ilities’, P roc. R. S oc. Lond. A 3 64 , 1-38.
Longuet-Higgins, M. S. and Fox, M. J. H.: 1977, ‘Theory of the alm ost-highest wave: the inner
solution’, J. Fluid M ech. 80, 721-741.
Longuet-Higgins, M. S. and Fox, M. J. H.: 1978, ‘Theory of the alm ost-highest wave. II.
M atching and an alytic extension’, J. Fluid M ech. 85, 769-786.
Longuet-Higgins, M. S. and Stew art, R. W .: 1964, ‘Radiation stresses in w ater waves. A
physical discussion with applications’, D eep-Sea Res. 1 1 , 529-562.
Tanaka, M. 1983: ‘The stab ility of steep gravity waves’, J. P hys. Soc. Japan 5 2 , 3047-3055.
2136
The ‘ almost-highest wave’, which is the asymptotic form of the flow in a steep
irrotational water wave of less than the limiting height, was recently shown to be
unstable to infinitesimal disturbances (see Longuet-Higgins & Cleaver 1994). It was
also shown numerically that the lowest eigenfrequency is asymptotic to that of the
lower superharmonic instability of a progressive wave in deep water (Longuet-Higgins,
Cleaver & Fox 1994). In the present paper these calculations are revised, indicating the
presence of more than one such instability, in -agreement with recent calculations on
steep periodic and solitary waves (Longuet-Higgins & Tanaka 1997).
The nonlinear development of the fastest-growing instability is also traced by a
boundary-integral time-stepping method and the initial, linear growth rate is
confirmed. The subsequent, nonlinear stages of growth depend as expected on the sign
of the initial perturbation. Perturbations of one sign lead to the familiar overturning
of the wave crest. Perturbations of the opposite sign lead to a smooth transition of the
wave to a lower progressive wave having nearly the same total energy, followed by a
return to a wave of almost the initial wave height. This appears to be the beginning of
a nonlinear recurrence phenomenon.
1. Introduction
In a recent paper (Longuet-Higgins & Cleaver 1994) it was shown that the local flow
near the crest of a steep, progressive gravity wave on water is itself subject to an
exponentially growing instability. To distinguish it from instabilities involving the
whole wave, this was called a ‘ crest instability’. In a second paper (Longuet-Higgins,
Cleaver & Fox 1994) it was further shown numerically that the crest instability is the
asymptotic form of the lowest superharmonic instability of a complete, periodic gravity
wave in deep water when the wave steepness ak tends to its limiting value
Сak)maz - 0.4423. As shown by Tanaka (1983) this instability first occurs when
ak = 0.4292, the value at which the total energy E attains its first maximum (see also
Saffman 1985). For solitary waves, a parallel result was demonstrated numerically by
Tanaka (1986).
In the above work all of the calculations, whether analytic or purely numerical, were
carried out on the basis of linear perturbation theory, and the departures from a
steady, nonlinear wave were assumed infinitesimally small. An interesting paper by
Tanaka et al. (1987) traced the subsequent nonlinear development of the fastest-
growing instability for the solitary wave by numerical time stepping. Depending on the
2137
where w= ^ (2. 7)
8 + ix
and 8 is a constant at our disposal (<$was taken to be 103/2). The series (2.6) is assumed
to converge inside and on the circle M = 1 which corresponds to the unperturbed free
surface. By applying suitable conditions on the boundary, the coefficients B0, B19..., B90
were determined to a high degree of accuracy.
The normal-mode perturbations of this flow were calculated in Longuet-Higgins &
Cleaver (1994). We write
z = Z + £, (2.8)
where Z is the previously determined flow, given by equation (2.6) and £ is a small
perturbation. The free surface is defined by \fr = F(0, t) and only terms linear in £ and
Fare retained. The kinematic and dynamical boundary conditions combined then lead
to a single (complex) condition to be satisfied on \jr = 0. The perturbations £ and Fare
expanded in series:
n -0
(2.9)
F — 2 (cncosm- + rfnsinrtr),
n -l
where e‘T= w and the coefficients an,bnicn,dn depend on the time t only. Assuming a
time dependence e* and equating coefficients of cosm- and sin/rr in the boundary
conditions as far as n = TV, say, one is led to a set of (4N4-2) simultaneous linear
equations for the vector
л: = (aQ,ai i ...iaN't b0,bl9... ybN' c19...,c N; dx, . . . 9dN) (2 .Ю)
in the form
(ЗАх = Bx (2.П)
where A and В are matrices of order (4ЛГ+2). The numerical solution of these
equations, for a sufficiently large value of the truncation parameter N, then yields the
required set of eigenvectors (3 and the corresponding eigenvector x , which in turn yields
a normal-mode perturbation. There is always one mode in which /? = 0, a0 Ф 0 and all
the other coefficients an vanish. This solution corresponds to a perturbation consisting
of a pure phase shift. In Longuet-Higgins et al. (1994), the results of Longuet-Higgins
& Cleaver (1994) for vanishingly small e were extended to positive values of e by
matching the ‘ inner’ solution for the flow near the crest (that is to say when z = 0(e2))y
to an ‘ outer’ flow representing the rest of the wave. The matching process requires a
modification to the inner flow by an amount of order e3(A-1), where Л is the lowest
positive root of
— tan— ____ (2.12)
2 2 2V 3
2139
F igure 1. The positive growth rate /?, of the lowest crest instability, plotted against e. The broken
curve represents the asymptotic formula (3 .3).
for the remaining coefficients. This yields xfr) and 1\ф) by equations (2.7). However,
since the free surface is given by \Jr = Ft not \jr = 0, we have for the surface profile to
calculate
у=У(ф,0) + [г{ф)^(ф,0) + ч(ф, 0 ) p ‘. (3.5)
We write this as
y —Y+aAyc^\ (3-6)
2141
ak
Figure 2. Growth rates Ля of the two lowest superharmonic instabilities of steep periodic waves, as
calculated by Tanaka (1995) (crosses). The circles are values of (ije from figure 1 of this paper.
where Ay denotes the expression in square brackets on the right o f (3.5), normalized
so that
max|A>»|=l. (3.7)
Essentially Ay gives the form o f the perturbation. In figure 4, Ay is plotted against the
unperturbed coordinate x/l with ф as parameter. The curves are shown for € = 0 and
0.08. Each curve has a relatively sharp maximum when x < 0 (on the forward face o f
2142
X
- -
X
-
-0.001 1 __ 1__ —1------ 1------ 1____I. 1 .1 1 1
0.1 0.2
F igure 3. Graph of f t (circles, this paper) and (еЛ„)2 (crosses, T anaka 1995) against e.
the wave) and a less pronounced minimum on the rear face. There is little difference
between the curves, however.
Figure 5 shows a comparison of the eigenfunction corresponding to e = 0.08 with
T an aka’s calculated eigenfunction when o j = 0.92 (e = <?/\/2c0 = 0.07533). The curves
are similar but by no means identical. It should be remembered, however, that the
eigenfunction corresponding to e = 0.08 is that of the inner solution, which is valid
over a distance x = 0(e2) only. For any non-zero value of e there has to be a matching
zone of width O(e) where the solution differs from the solution in the inner zone by a
proportional difference which is 0(1). In the region where x /e 2 is about 10, say, x /e is
about 0.8 so that some significant differences are to be expected.
x/l
F igure 4. Vertical displacement Ay of the free surface plotted against the scaled horizontal
coordinate x in the cases e = 0 and 0.08. The wave travels to the left.
so that x is an analytic function of z. At the free surface then the kinematic condition
can be written
D z _ fd * y
(4.2)
Dt \dz) ’
where D/D/ denotes (0/6/ + V0-V) and a star denotes the complex conjugate. The
condition of constant pressure can be written
Dф _ 1
(4.3)
Df gy 2
where g denotes gravity. We shall take g —1 and also in this application the
wavelength 1 = 1 . The mapping w = exp(—2niz) maps one wavelength onto the
interior of a closed contour C(t) in the w-plane, and we may apply Cauchy’s theorem
(4.4)
&У
F igure 5. Comparison o f the lowest normal-mode instability o f the almost-highest wave when
e = 0.08 (solid curve) with the lowest instability o f a periodic wave when e = 0.0753 (dashed curve).
At any instant /, the boundary conditions (4.2) and (4.3) provide new values of the
coordinates z and velocity potential фon С at the time (/ + dr). Then a discretized form
of (4.4) is used to find the unknown streamfunction xfr on C. On C, piecewise-linear
functions are used to represent фand \{r; the resulting influence coefficients are as given
in Vinje & Brevig (1981). To advance in time we use a fourth-order Runge-Kutta
scheme. In order to prevent the growth of saw-tooth instabilities, five-point smoothing
and re-gridding was applied after every ten time steps.
As a check on the accuracy of the method we first applied it to a steep Stokes wave
with steepness parameter
ak —0.43982. (4.5)
Here 2a denotes the crest-to-trough wave height and к = 2n/L is the wavenumber.
(This corresponds to H/L = 0.14.) The wave profile was found by expanding the
surface elevation and velocity potential in Fourier series with 4000 terms, and adjusting
the coefficients so as to satisfy boundary conditions (for steady flow) at 4000 equally
spaced grid points on the surface. To proceed with the time stepping, we used 2000
unequally spaced points; near the wave crest the spacing was reduced to provide better
resolution. The smallest spacing Ax was 0.61 x 10"4 at the crest and 0 .4 3 x 10"3 in the
trough.
A measure of the numerical error in the Bernoulli equation is provided by the
expression
\В+ г1+ №фУ + сфх\ (46)
\В\+ Н+\ФФУ + с\фх\
which was found to be less than 3 x 10-6. (Here В is the Bernoulli constant, the mean
level being taken as zero.)
2145
X
F igure 6 . Form o f the initial perturbations when a = 1, as functions o f the unsealed horizontal
coordinate x: ( a) Ay, (b ) Дф.
2147
x/L
F igure 7. Initial development of the positive perturbation, when a = 0.2, in a moving
reference frame.
Figure 12 shows £ as a function o f the wave steepness ak. Thus the initial wave, for
which ak = 0.4398 and e = 0.08487, has an energy E = 0.07306, which is the same as
that o f a lower wave with e = 0.2239 and hence ak = 0.4182 (H/L = 0.1331).
The actual long-term development o f the negative instability is shown in figures 13
and 14. Figure 13 shows the crest-to-trough wave height 2 a as a function o f the time,
and figure 14 shows the phase difference, as measured by the position o f the maximum
elevation o f the crest. From figure 13 it appears there is indeed a transition to a lower
wave o f height 0.1303, approximately as expected, but that the wave then returns
almost to the original height, in a kind o f recurrence, and then oscillates. The difference
in phase speed relative to the original wave may be measured by the gradient o f the
2148
F igure 8 . Range o f the perturbations as a function o f the time /. Positive and negative
perturbations (a = ± 0.2) are marked by crosses and square plots, respectively.
x/L
F igure 9. Development o f the surface profile after the positive perturbation (a = 0.2) at time intervals
A/ = 0.05, as seen in a reference frame moving with the phase speed o f the unperturbed wave.
2149
x/L
Figure 10. Development of the negative perturbation, when a = -0 .2 , in the moving reference
frame of figure 9.
(x+ct)/L
F igure 11. Development of the surface profile after the negative perturbation (a = —0.2) at time
intervals Л/ = 0.2, as seen in the moving reference frame.
ak
Figure 12. The total energy E and the phase speed с as functions of the wave steepness ak.
t
F igure 13. Crest-to-trough height of the perturbed wave, as a function of the time /.
Figure 15 shows the complete surface profiles at times t = 0,8.4 and 16.0, the vertical
scale being exaggerated. The recurrent character of the transition is clearly different
from the corresponding behaviour in the case of the solitary wave (Tanaka et al. 1987)
where the wave height tends to its lower value exponentially in time. As a check on the
accuracy of the numerical time stepping in figures 13 and 14, we note that the total
energy at time /=19.6 was 1.85022 x 10~3 compared with the initial energy
1.850 31 x 10~3 at time /= 0, a proportional difference of less than 5 x 10"5.
2151
t
Figure 14. Phase of the perturbed wave, as a function of the time t.
x/L
Figure 15. Surface profiles at times t = 0, 8.4 and 16.0. The vertical scale is enlarged.
e ak H/L E c/ c0
0.08487 0.43982 0.1400 0.07306 1.09262
0.22400 0.41818 0.1331 0.07306 1.08894
Table 3. Comparison of wave parameters
REFEREN CES
C leaver, R. P. 1981 Instabilities o f gravity waves. PhD thesis, University o f Cambridge.
D uncan , J. H., Philomin, V., Q iao, H. & K immel, J. 1994 The formation o f a spilling breaker. P hys.
Fluids 6 , S2.
J illians, W. J. 1989 The superharmonic instability of Stokes waves in deep water. J. Fluid M ech.
204, 563-579.
Longuet-H iggins, M . S. 1992 C apillary rollers and bores. J. Fluid M ech. 240, 659-679.
Longuet-H iggins, M. S. & C leaver, R. P. 1994 Crest instabilities of gravity waves. Part 1. The
almost-highest wave. J. Fluid M ech. 158, 115-129.
L onguet-H iggins, M. S., C leaver, R. P. & Fox, M . J. H. 1994 Crest instabilities of gravity waves.
Part 2. M atching and asymptotic analysis. J. Fluid M ech. 259, 333-344.
Longuet-H iggins, M. S. & C okelet, E. D. 1976 The deformation of steep surface waves on water.
I. A numerical method of computation. P roc. R. S oc. Lond. A 350, 1—26.
Longuet-H iggins, M .S . & Fox, M . J . H . 1977 Theory of the almost-highest w ave: the inner
solution. J. Fluid M ech. 80, 721-741.
Longuet-H iggins, M. S. & Fox, M. J. H. 1978 Theory of the almost-highest wave. Part 2.
Matching and analytic extension. J. Fluid M ech. 259, 333-344.
Longuet-H iggins, M. S. & T anaka , M . 1997 On the crest instabilities of steep surface waves.
J. Fluid M ech. 336, 51-68.
2153
1. Introduction
The stability of a train of steep, progressive, irrotational waves in water of finite or
infinite depth is of some interest in connection with the theory of wave breaking.
Subharmonic instabilities, having wavelengths larger than the basic wavelength 2n/k,
can occur at relatively low wave steepness. But if we confine attention to the
superharmonic instabilities, that is to say two-dimensional instabilities having the same
space-periodicity as the unperturbed wave, then it is found that only very steep gravity
waves are appreciably unstable.t In deep water it was shown numerically by Tanaka
(1983) and analytically by Saffman (1985) that gravity waves first become unstable in
this way when the steepness parameter ak attains the value 0.4292, corresponding to
the first maximum in the total energy density. Similarly, for solitary waves Tanaka
(1986) showed that instability first occurred at a height-to-depth ratio a/h —0.7824,
which again corresponded to the first maximum in the total wave energy E.
The wave steepnesses just mentioned are already high, being close to the limiting
steepnesses: ak = 0.4432 for a wave in deep water and a/h = 0.8332 for the solitary
wave. Calculation of the properties of such steep waves by ordinary methods presents
increasing difficulties as the limiting wave is approached. It was therefore of some
interest when Longuet-Higgins & Fox (1977) showed that the crest of any steep Stokes
wave (not the steepest) approached a certain asymptotic form characterized by a length
scale
/ = <?72g, (11)
t We are excluding the ‘ bubbles’ of non-stationary instability pointed out by MacKay & Saffman
(1986) and Kharif & Ramamonjiarisoa (1990).
2155
= gl/4 (1.2)
to be small, c0 = (g/k:)1/2 being the linear phase speed. A corresponding theory for
solitary waves, in which c0 = (gd)1/2, has been given by Longuet-Higgins & Fox (1996).
Now in two recent papers (Longuet-Higgins & Cleaver 1994; Longuet-Higgins,
Cleaver & Fox 1994) it was suggested that for very steep waves the lowest
superharmonic instability of a progressive wave should correspond, in the limit as
e-*0, with an instability of the inner flow of the almost-highest wave. The lowest
instability of the inner flow was therefore calculated, also the first-order correction to
this instability resulting from the matching process. The growth rates of the instability
apparently agreed with those found numerically by Tanaka (1983) for progressive
waves of finite amplitude. Tanaka’s calculations, however, at that time extended only
as far as ak = 0.439, and no eigenfunctions were available. One of us (M.S.L.-H.)
therefore proposed to Tanaka to extend his calculations (with a more modem
computer) to higher values of ak and to calculate if possible the corresponding
eigenfunctions. It is these calculations, first presented last year (Tanaka 1995), that are
reported fully in the present paper. We report also similar calculations for solitary
waves which are more accurate than those for periodic waves.
The calculations by Tanaka (1995) showed the need for a re-examination of the
numerical calculation reported in Longuet-Higgins & Cleaver (1994) and Longuet-
Higgins et al. (1994). This has led to a significant improvement in their accuracy (see
Longuet-Higgins & Dommermuth 1997, referred to herein as LHD). As will be shown
here, the revised values of the rate of growth as € - * 0 agree satisfactorily with those
reported by Tanaka (1995), especially for the lowest mode {n = 1). The revised
calculations also show the presence of a second mode (n = 2 ) as was originally
suggested by Cleaver (1981); this also is consistent with Tanaka’s (1995) calculations.
The plan of this paper is as follows. In §§2 and 3 we describe the method of
computation, and the numerical results, for steep periodic waves in deep water, and in
§§4 and 5 we present corresponding results for the solitary wave in shallow water. In
§ 6 we compare the two sets of results with one another and with the eigenfunctions
obtained by LHD for the limiting case e-*0. A discussion follows in §7.
N ak с error (%)
1[ 64 0.4267367789 1.0915699348 1.805 x 10-“
w = 0.80 , 128 0.4267368619 1.0915699350 3.820 x 10“10
1[ 256 0.4267368619 1.0915699350 1.602 x Ю" 10
512 0.4426166715 1.0922690822 5.260 x 10°
w = 0.97 - 1024 0.4427541436 1.0922769604 5.666 x 10-3
2048 0.4427542224 1.0922769604 1.258 x 10-®
T able 1. Convergence of the Stokes wave in deep water
(O ak £ c/c0 Rk/e2
0.80 0.42674 0.18858 1.09157 5.3274
0.81 0.42855 0.17917 1.09201 5.3063
0.82 0.43022 0.16974 1.09235 5.2869
0.83 0.43173 0.16030 1.09261 5.2692
0.84 0.43312 0.15086 1.09279 5.2529
0.85 0.43437 0.14141 1.09290 5.2382
0.86 0.43551 0.13196 1.09295 5.2248
0.87 0.43655 0.12251 1.09294 5.2128
0.88 0.43749 0.11307 1.09289 5.2021
0.89 0.43835 0.10362 1.09282 5.1926
0.90 0.43912 0.09419 1.09272 5.1843
0.91 0.43983 0.08475 1.09261 5.1772
0.92 0.44047 0.07533 1.09251 5.1711
0.93 0.44106 0.06591 1.09242 5.1659
0.94 0.441 58 0.05649 1.09235 5.1618
0.95 0.44204 0.04707 1.09231 5.1585
0.96 0.44244 0.03766 1.09228 5.1560
T able 2. Calculated values of the parameters for periodic waves in deep water
Table 1 shows the convergence of the steepness parameter ak and the phase speed
с as the number N of Fourier modes used to express the flow is increased, for the cases
0) = 0.80 and и) = 0.97. The maximum deviation of the Bernoulli constant \q2 + gy
along the free surface is also shown. The limiting value of ak is known to be 0.4432,
hence the ak of the two cases shown in the table are about 96.3% and 99.9% of the
limiting value. The table shows that N = 128 is sufficient to get the convergence of с
to the ten decimal places for w = 0.80, while more than N = 1024 is necessary to get
the same convergence when w = 0.97. The speed of convergence of the series
representing the Stokes wave also determines the speed of convergence of the solution
of the eigenvalue problem in the stability analysis, and this rapid slowdown of
convergence for larger values of шcauses serious difficulty in the stability calculation
later.
The range 0.80 ^ cj ^ 0.97 corresponds to 0.1886 ^ e ^ 0.0282, see table 2. Within
this range of e the asymptotic relations for ak and с derived by Longuet-Higgins & Fox
(1978) coincide with our numerical results up to four significant figures. The last
column of table 2 shows the radius of curvature R at the crest converging to the value
5.15c2 predicted in Longuet-Higgins & Fox (1977). According to the asymptotic
theory for the deep-water wave given in Longuet-Higgins & Fox (1978), the lowest-
2157
- -
X
“
5.30
- X
-
- -
- X -
RJl 5.25 — X
—
- X -
- -
X
- -
X
- _
5.20 ___ X
_
_ * _
X
X -
_ X _
X
X
order correction to the inner solution is proportional to e3(e-1) where a is the lowest
root of the transcendental equation
that is to say a = 1.8027. In figure 1 we have plotted R/l against e3<e-1>, that is e2,4081
and it can be seen that the plots lie almost on a straight line. (A small oscillation about
this line is also to be expected from the analysis.) From figure 1 we may deduce the
more accurate expression
R/l ^ 5.1525 + 9.5 e3^ 1* (2.2)
for the crest curvature R.
К A? At A?
10 -0.0 2 2 7 6 w = 0.97 70 2.94771
15 -0 .0 1 8 2 0 120 2.63295
20 -0 .0 1 5 0 6 200 2.45113
30 -0 .0 1 4 8 8 300 2.40413
50 -0 .0 1 4 8 8 400 2.39750
70 -0 .0 1 4 8 8 500 2.39659
T able 3. Convergence of the eigenvalue of the lowest mode
tO 90
OJ E A? (eAj)2 (eA2)2
0.80 0.073991 -0 .0 14 8 8 -0.000529 -0 .53 7 23 -0 .0 1 9 1 0 5
0.81 0.074032 -0 .00 3 99 -0.00 0 12 8 -0 .5 0 8 17 -0 .0 1 6 3 12
0.82 0.074026 0.00766 0.000221 -0.4 7 8 6 4 -0 .0 13 7 9 0
0.83 0.073979 0.02026 0.000521 -0 .44 8 60 -0 .0 1 15 2 8
0.84 0.073898 0.03411 0.000776 -0 .4 1 8 0 0 -0 .0 0 9 5 13
0.85 0.073791 0.04957 0.000991 -0.3 8 6 7 5 -0.00 7 73 4
0.86 0.073668 0.06718 0.001170 -0.3 5 4 8 0 -0.00 6 17 8
0.87 0.073534 0.08774 0.001317 -0.3 2 2 0 6 -0.004834
0.88 0.073399 0.11238 0.001437 -0.28 8 49 -0.003688
0.89 0.073268 0.14288 0.001534 -0.25 4 06 -0.00 2 72 8
0.90 0.073149 0.18197 0.001614 -0 .2 18 8 3 -0 .0 0 19 4 1
0.91 0.073046 0.23405 0.001681 -0 .18 2 9 4 -0 .0 0 13 14
0.92 0.072964 0.30643 0.001739 -0 .14 6 5 7 -0.00 0 83 2
0.93 0.072904 0.41179 0.001789 -0 .10 9 8 1 -0.000477
0.94 0.072867 0.57390 0.001831 -0 .0 7 2 19 -0.000230
0.95 0.072849 0.84216 0.001866 -0 .0 3 19 7 -0.000071
0.96 0.072847 1.33468 0.001893 0.01593 0.000023
0.97 0.072852 2.39659 0.001912 0.08742 0.000070
T able 4. Calculated values of the total energy density E and of the squared eigenfrequencies of the
normal-mode perturbations of a deep-water wave
Table 3 shows the convergence of the square of the eigenvalue of the mode
n = 1 for the deep-water waves with o>= 0.80 and o) = 0.97. In these tables, Ne is
the number of Fourier modes which are used to express the eigenfunctions and
does not have any direct relation with N of table 1 besides the trivial relation Ne < N.
In terms of Ne the size of the matrix whose eigenvalue problem we need to solve is
(2Ne+ 1) x (2Ne+ 1 ). When o>= 0.80 we can have convergence of A* to five decimal
places with Ne = 30, while for o j = 0.97 we get the convergence to only two decimal
places even when Ne is increased to 400. When o j = 0.97 and Ne —400, the total
CPU time used in the construction of the matrix and the calculation of all the
eigenvalues and the eigenvectors was about 12 s on VP2600 of the Computation
Center of Nagoya University.
In the analysis of the local flow around the crest by LHD time and space are
normalized in such a way that g = 1 and qc = \/2, while in our calculations they are
normalized such that g = к = 1 for the Stokes waves in deep water. The propagation
speed of the basic steady wave c0 in the limit of infinitesimal amplitude equals 1. In
order to compare our results with those of LHD, we need to multiply our quantities
by e~2 for each dimension of space and e"1 for each dimension of time.
2159
ез<«-1)
F igure 2. Squares of the growth rates A„ of the two lowest modes of instability of the deep-water
wave, scaled by the parameter e, shown as a function of e3'*-1’. Crosses denote present calculation;
circles, the asymptotic theory of the almost-highest wave, from Longuet-Higgins & Dommermuth
(1997).
The square of the scaled eigenvalues еЛ1 еЛ2 of the two lowest eigenmodes n = 1 and
n = 2 are shown in table 4 and plotted in figure 2. The time-dependence of the
perturbation is assumed here to be like eA‘, hence Л2 > 0 implies instability. On the
other hand the asymptotic theory given in Longuet-Higgins et al. (1994) and LHD
indicates that the lowest-order corrections to the eigenfrequencies of the inner flow
vary as e3(a_1>. In figure 2 we therefore plot the results as functions of e3(e_1). As
expected we see that the plots indeed fall approximately on a straight line, for small e.
A slight oscillation about the straight line is also seen.
Figure 2 also shows the corresponding results from LHD, and it will be seen there
is close agreement over the expected range of e.
Figure 3 shows the profile of the lowest unstable mode (n = 1) as a function of the
horizontal coordinate x for oj = 0.84, 0.88 and 0.92. It can be seen that as ш increases
the perturbation progressively shrinks and is confined to a narrower region around the
crest (which is at x = 0). The arrow in the figure indicates the direction of propagation
of the Stokes wave. The amplitude of the normal mode, which is arbitrary within the
framework of the linear stability analysis, is normalized in such a way that the
maximum deviation of the profile is equal to unity, and this convention will be used
throughout this paper.
Figure 4 shows the profiles of the same normal mode as those shown in figure 3, but
as functions of the scaled horizontal coordinate x/e2. The widths of the curves are now
2160
F ig u r e 3. Vertical displacement rjl in the lowest unstable mode of a deep-water wave as a function
of the horizontal coordinate x , when <
j = 0.84, 0.88 and 0.92.
F ig u re 4. Vertical displacement tj1 in the lowest unstable mode of a deep-water wave as a function
of the scaled coordinate x/e2, when w = 0.84, 0.88 and 0.92.
2161
than the deep-water wave case for our present purpose which is to clarify the asymptotic
behaviour of unstable normal modes as e-^O.
Table 5 indicates that the scaled radius of curvature R/l approaches the theoretical
value 5.15 as before. According to the asymptotic theory for steep solitary waves
(Longuet-Higgins & Fox 1996) we expect the lowest-order corrections to the inner
solution to be of order e'[3<e-1)], as for the deep-water wave. Accordingly in figure 5 we
have plotted R/l against e't3<«-D3 an(j we obtain as previously an almost straight-line
plot.
Figure 5. The scaled crest curvature R/l for solitary waves, as a function of e',3<e_l>1.
2164
q/c / E A? A’
0.01 2.54347 0.97268 13.89290 0.71497
0.02 2.54347 0.97268 3.45947 0.15537
0.03 2.54337 0.97260 1.52518 0.05006
0.04 2.54315 0.97246 0.84689 0.01488
0.05 2.54286 0.97233 0.53209 0.00242
0.06 2.54261 0.97228 0.36050
0.08 2.54275 0.97275 0.188 86
0.10 2.54449 0.97439 0.10866
0.12 2.54839 0.97738 0.06489
0.14 2.55467 0.98164 0.03875
0.16 2.56312 0.98684 0.02242
0.18 2.57325 0.99248 0.01211
0.20 2.58438 0.99795 0.00577
0.22 2.59567 1.00260 0.00215
0.24 2.60620 1.00578 0.00043
0.25 2.61089 1.00661 0.00008
0.26 2.61506 1.00685
0.28 2.62138 1.00527
0.30 2.62436 1.00057
T able 7. C alculated values o f the im pulse I and total energy E o f the so litary w ave, and o f the
squared grow th rates of the two lowest instabilities
e'[3(a-D)
Figure 7. Scaled values of the squared growth rates of the two lowest modes of instability o f the
solitary wave, as a function of
2165
[(d)
x x
Figure 8. Profile of the first unstable mode o f the solitary wave (solid line). The dashed curve shows
the unperturbed wave profile: (a) q/c = 0.25, (b) q/c = 0.20, (c) q/c = 0.10, ( d ) q/c = 0.02.
F igure 9. Profiles of the first unstable mode of the solitary wave plotted against the scaled
coordinate x/e'2.
2166
x
F igure 10. Comparison of the first unstable mode of the solitary wave when q/c = 0.25 (solid curve)
with a pure phase shift (dashed curve).
Numerical values of the total impulse and energy are given in table 7, along with the
squared growth rates of the two lowest instabilities.
In figure 7 we show the square of the scaled eigenvalues e'An of the two lowest
unstable modes plotted as functions of e,[3(a-1>1. As yet there is no corresponding
asymptotic theory for the eigenvalues at general values of e, but the points approach
the vertical axis along a straight line, as expected. The limiting values of e\n as e-*0,
according to the asymptotic analysis of LHD, were shown in figure 2 above. It will be
seen that these are in reasonable agreement with the present calculations.
Figure 8 (a-d) shows the profiles of the lowest unstable mode (n = 1) for q/c = 0.25,
0.20, 0.10 and 0.02. The dashed line in each figure shows the profile of the unperturbed
solitary wave for the corresponding value of q/c. The horizontal axis is the original
horizontal coordinate x for which g = 1 and d = 1 .It can be seen clearly that the length
scale of the normal mode gradually diminishes as q/c^0 and the perturbation is
confined to a narrower region around the crest in a similar manner as that we observed
for the first unstable mode of the deep-water wave (see figure 3). When we replot in
figure 9 those profiles as functions of the scaled coordinate x/l (that is x/e'2 in our
scaling), there appears to be a convergence of the profile for sufficiently small values
of q/c (but see § 6 below). Figure 10 compares the profile of the first unstable mode
when q/c = 0.25 (solid line) with the profile of the trivial neutral mode corresponding
to a horizontal shift (dashed line); this is simply the derivative of the profile of the
steady solitary wave, for this value of q/c. According to Saffman (1985) the trivial
neutral mode and the non-trivial mode should become identical at the point where the
non-trivial mode exchanges its stability. As the critical point for the lowest non-trivial
mode, which is equal to the first maximum of the total energy, is q/c « 0.259 as figure
6 indicates, the case shown in figure 10 (q/c = 0.25) is supercritical (in this sense) only
marginally. The closeness of the two profiles shown in figure 10 can be regarded as an
indication that our numerical results are in accord with Saffman’s theory.
Figure 11 (a-d) shows the profile ij2 of the second unstable mode as a function of *
for q/c = 0.05, 0.04, 0.03 and 0.02, respectively. The first three curves are also plotted
2167
: Ф)
: ^
x x
F ig u re 11. Profile of the second unstable mode of the solitary wave (solid line). The dashed curve
shows the unperturbed wave profile:(a) q/c = 0.25, ( b ) q/c = 0.20, (c) q/c = 0.10, (d ) q/c = 0.02.
F ig u re 12. Profiles of the second unstable mode of the solitary wave plotted against the scaled
coordinate х/е'г.
2168
depend only on the velocity q, for fixed g, and not on e, so that the scaled frequencies
( 6 .2)
g
can be compared in the limit as e-+0.
For the limiting deep-water wave we have from figure 13, since q = \/2e,
Axq
= 0.0621, ^ = 0.016. (6.3)
g g
This can be compared with the limiting solitary wave for which, from figure 7,
A, q Aj? (6.4)
= 0.0621, ^ = 0.014.
g S
On the other hand the limiting values according to the asymptotic theory of the almost-
highest wave (LHD) are
e 3 ( « - i)
F ig u re 13. Extrapolation of the eigenfrequencies eA, and еЛ2 for the deep-water wave, from the three
lowest calculated values of e. The fitted lines are given by у = 0.0019305-0.10138 * when n = 1 and
by у = 0.0013251 -0.3147 7 л: when n = 2.
F igure 14. Comparison of the first unstable mode of a deep-water wave (w = 0.84, solid curve) with
that of a solitary wave (q/c = 0.16, dashed curve).
2170
when x fl is of order 50. However, when x/l is of order 5, say, we would expect closer
agreement than is evident in the figure.
The eigenfunction for the solitary wave at q/c —0.02 shown in figure 15 is very close
to the solitary-wave eigenfunction at q/c = 0.16 shown in figure 14, which suggests that
it has effectively converged, especially in the inner zone where x/l is of order ■
Probably, therefore, the present determination of the lowest mode is more accurate
than that presented in LHD.
REFERENCES
C leaver, R. P. 1981 Instabilities of surface gravity waves. PhD thesis, University of Cambridge.
J illians, W. J. 1989 The superharmonic instability of Stokes waves in deep water. J. Fluid Mech.
204, 563-579.
K h a r if, C. & R am am onjiarisoa, A. 1990 On the stability of gravity waves on deep water. J. Fluid
Mech. 218, 163-170.
Longuet-H iggins, M. S. & C leaver, R. P. 1994 Crest instabilities of gravity waves. Part 1. The
almost-highest wave. J. Fluid Mech. 258, 115-129.
Longuet-H iggins, M. S., C leaver, R. P. & Fox, M. J. H. 1994 Crest instabilities of gravity waves.
Part 2. Matching and asymptotic analysis. J. Fluid Mech. 259, 333-344.
Longuet-H iggins , M. S. & D ommermuth, D. G. 1997 Crest instabilities of gravity waves. Part 3.
Nonlinear development and breaking. J. Fluid Mech. 336, 33-50 (referred to herein as LHD).
L onguet-H iggins, M. S. & Fox, M. J. Н. 1977 Theory of the almost-highest wave: the inner
solution. J. Fluid Mech. 80, 721-741.
Longuet-Higgins , M .S. & Fox, M .J.H . 1978 Theory of the almost-highest wave. Part 2.
Matching and analytic extension. J. Fluid Mech. 85, 769-786.
Longuet-Higgins, M. S. & Fox, M. J. H. 1996 Asymptotic theory for the almost-highest solitary
wave. J. Fluid Mech. 317, 1-19.
M a c K a y , R. S. & S affman, P. G. 1986 Stability of water waves. Proc. R. Soc. Lond. A 406,
115-125.
S affman , P. G. 1985 The superharmonic instability of finite-amplitude water waves. J. Fluid Mech.
159, 169-174.
T a n aka , M. 1983 The stability of steep gravity waves. J. Phys. Soc. Japan 52, 3047-3055.
T a n aka , M. 1986 The stability of solitary waves. Phys. Fluids 29, 650-655.
Ta n a k a , M. 1995 On the “crest instabilities” of steep gravity waves. Workshop on Mathematical
Problems in the Theory of Nonlinear Water Waves, CIRM, Luminy, France, May 1995 (Abstract
only).
T a n a ka , М., D old , J. W., L ew y , M. & Peregrine, D. H. 1987 Instability and breaking of a solitary
wave. J. Fluid Mech. 185, 235-248.
Z ufiria, J. A. & S affman , P. G. 1986 The superharmonic instability of finite-amplitude surface
waves on water of finite depth. Stud. Appl. Maths 74, 259-266.
2172
High-speed photographs of a gently breaking water wave have shown the partic
ular instability of the wave crest predicted by Longuet-Higgins et al. (J. Fluid
Mech. 259, 333-344 (1994)) and Longuet-Higgins & Cleaver ( J. Fluid Mech. 258,
115-129 (1994)), with a ‘bulge’ on the forward face of the wave and the genera
tion of parasitic capillaries ahead of the instability, emanating from the ‘toe’ of
the bulge. The photographs also show the unexpected occurence of longer (type
2) capillary waves above the toe of the bulge. In this paper it is shown that the
type 2 capillaries are probably shear-flow instabilities arising from the vorticity
shed by type 1 (parasitic) capillaries.
At large amplitudes the initially linear shear instabilities must break up into
discrete vortices. By formulating a simple model of the shear instabilities as
vortex waves it is shown that the wavelength of the type 2 capillaries should
be about 4 times the wavelength of the type 1 capillaries, a result in agreement
with observation.
When the flow separates at the toe of the bulge, the use of Prandtl’s rule to
estim ate the strength of the vortices confirms the above relation. At larger length-
scales when surface tension is negligible, the argument leads to a simple rule for
predicting the wavelength of the initial surface roughness of a spilling breaker.
1 . In tro d u ctio n
Gently breaking water waves (‘spilling breakers’) have long been only partly un
derstood. Although a physical model for the quasi-steady flow, in which a tur
bulent whitecap propagates steadily down the forward face of wave, was given
some years ago by Longuet-Higgins & Turner (1974) and a simple model of the
forward tip of the wave was proposed by Longuet-Higgins (1973, 1990), the ques
tion of how a spilling breaker is initiated remained obscure until recently when
Duncan et al. (1994 a, b) published high-speed observations that showed that the
process begins with the formation of a ‘bulge’ or instability on the forward face
of the wave, in accordance with theoretical calculations by Longuet-Higgins &
Cleaver (1994) and Longuet-Higgins et al. (1994). In fairly short gravity waves,
with length 1 m or less, the sharp curvature at the ‘toe’ of the bulge produces a
train of ‘parasitic’ capillary waves (see Cox 1958) propagated ahead of the bulge,
owing to anomalous dispersion. Such capillary waves have been shown to be a
surprisingly strong source of vorticity, capable of producing a ‘roller’, or closed
circulation in the crest of the gravity wave (see Longuet-Higgins 1992). However,
t This paper was produced from the author’s disk by using the Tfe* typesetting system.
Proc. R. Soc. Lond. A (1994) 446, 399-409 © 1994 The Royal Society
Printed in Great Britain 399
2173
400 M. S. Longuet-Higgins
-4 --------------
type 2
Figure 1 . Sketch of the crest instability of a steep gravity wave, slowing the initial bulge, with
type 1 (parasitic) capillaries ahead of the toe and type 2 capillaries above the toe. The gravity
wave travels to the left.
the photographs by Duncan et al. also showed the unexpected growth of a train
of longer capillary waves above the toe of the bulge, as in figure 1 . The latter cap
illary waves (which we call type 2 to distinguish them from the type 1 parasitic
capillaries) grew in amplitude until they broke and collapsed into turbulence.
The purpose of this paper is to suggest an explanation for the type 2 capillaries.
In short, it is proposed that they are shear instability waves arising from the
shear layer due to vorticity generated by the type 1 capillaries. The theory of
the instability of horizontal shear flows with a free upper surface was initiated by
Stern & Adam (1973) and has been developed by Kawai (1977), Voronovich et
al. (1980), Caponi et al. (1991), Morland et al. (1991), Shrira (1993) and others.
Stern & Adam (1973) consider a simple layer of constant vorticity and uniform
depth overlying a stationary, infinitely deep fluid. Their model leads to a cubic
equation for the dispersion relation, which was analysed in detail by Caponi et
al. (1991). The theory has been developed in connection with the generation of
surface waves by wind. Kawai (1977) showed that the neglect of viscosity had little
effect on the growth rates. Morland et al. (1991) have compared the ‘piece-wise
uniform’ shear model with three smooth distributions of vorticity and have shown
that the wavelength of maximum growth is not sensitive to the model, although
the rates of growth are. Voronovich et al. (1980) added a thin, discontinuous shear
layer at the base of the uniform layer, and showed that this had a de-stabilizing
effect on the rates of growth. They also discussed the possibility of an ‘explosive
instability’ arising from the nonlinear interaction of the resonant triads of linearly
unstable modes.
In §§2 and 3 below we calculate the wavelengths predicted by the model of
Stern & Adam (1973) and show they are in good agreement with the type 2
instabilities observed by Duncan, on the present hypothesis. On the other hand a
much simpler model is proposed in § 4 for the later (nonlinear) stages of growth.
It is argued that the shear layer will tend to break up quickly into an array
of discrete vortices, whose strength is determined by the strength of the initial
shear-layer and the wavelength of the instability. The model leads to a very simple
Proc. R. Soc. Lond. A (1994)
2174
Figure 2. A surface current with uniform shear, in the model proposed by Stern & Adam
(1973). In this diagram, the x-axis is directed to the left.
prediction, namely that the wavelength of the type 2 capillaries will be about four
times that of the type 1 capillaries, if there is no flow separation at the Чое’ of
the crest instability. If there is flow separation, as predicted in Longuet-Higgins
(1992) then the application of Prandtl’s rule allows a sim ilar prediction to be
made (§5).
Such predictions apply in the first place only to the breaking of rather short
gravity waves, with lengths less than about 1 m. The most prominent whitecaps
in the ocean occur on larger scales. It is therefore of some interest that the use of
Prandtl’s rule enables one to generalize the earlier results to much larger spilling
breakers and to estimate the horizontal scale of their surface roughness (see § 6 ).
and the stream-functions in and beneath the shear layer have the forms
= \Пу2
2 + (ПН - c)y + (АекУ + Ве~кУ)ёкх
(2.3)
Ф2 = -\П Н 2 - су + Dekyeikx, '}
where а ь а2, А ,£ and D are constants, possibly complex. On applying the con con
ditions of constant pressure at у = 771, continuity of pressure across у = r)2y and
Proc. R. Soc. Lond. A (1994)
2175
402 M. S. Longuet-Higgins
the kinematic boundary conditions, one obtains a dispersion relation in the form
of a cubic equation for
в = (с - tis)/c0, where c0 = (g/k + Tk)1/2 (2.4)
and g and T denote gravity and surface tension. (The density is taken as 1 .) The
equation for 9 is
в3 + (F + 2q)62 - ( 1 + 0F )6 + F = 0, (2.5)
where
q = u./c„ 0 = q/kH, F = 10(1 - e~2kH) - q. (2.6)
After solving (2.5), the radian frequency a of the waves relative to the stationary
axes is found from
а = ck = (в + q)c0k. ( 2 .7 )
If any of the roots в of equation (2.5) are complex, then <
7 * = Im(cr) is interpreted
as a time-rate of the growth of the perturbation.
It is convenient to take units such that
<7 = 1 , T = 1, (2 .8 )
so that the wavenumber km and speed Cm of ordinary surface waves of minimum
phase speed are 1 and y/2 respectively. The corresponding wavelength Am is 27r.
In c.g.s. units, Am and cm would be 1.74 cm and 23 cm s-1, for clean water at
20 °C.
In analysing equation (2.5) Caponi et al (1991) showed that instability occurs
only if us > cm and if q and /? lie within a certain wedge-shaped zone in the ( q, /?)
plane. Morland et al (1991) showed that the rate of growth satisfies the Howard
semi-circle theorem, namely that с lies inside a circle with the line joining (0 , 0 )
and (us,0) as diameter. For a typical dispersion relation derived from equation
(2.5) see figure 3 below.
Figure 3. Dispersion relation according to the model of figure 2, when uB= 4.0(gT)l/4 and
H = 0.4(p/T)1/2. (a) crr as a function of к and (6) (7* as a function of к in dimensionless units.
us = 4.0, H = 0.4 in dimensionless units. In the range 0 < A < 1.31 there are
three stable solutions, the uppermost curve corresponding to the surface gravity-
capillary mode, and the lower two curves mainly waves on the interface у — —H.
Sim ilarly when к > 7.35. In the intermediate range, however, cr has a non-zero
im aginary part, and one mode is unstable. Prom figure 3 6 , which shows o* plotted
against fc, it can be seen that the maximum rate of growth occurs when fc = 3.41,
corresponding to a (dimensional) wavelength of 1.74/3.41 = 0.51 cm. This is not
fax from the observed value of 0 .5 5 cm.
The corresponding growth-rate in figure 3 6 is <
7 * = 0.965, or dimensional units
4. Vortex waves
Consider a horizontal array of vortices, having circulation 27Г7, each at a depth
h below the near surface level and separated by a distance 2тг/fc (see figure 5).
We shall later take h = \H and fc = fc2. If the free surface were a perfect plane,
there would be an image array of vortices with circulation —27Г7 at a height h
above the surface. Under the influence of the image array the whole system would
move to the left with speed
Proc. R. Soc. bond. A (1994)
2177
404 M. S. Longuet-Higgins
- X
Figure 4. Steamlines in the most unstable mode corresponding to figure 3 a, b. The broken
curves indicate the boundaries of the vortical layer. (The я-axis is to the left).
Q -y 0 " y Q
G О G
5 ук coth kh
Figure 5. Sketch of vortex waves, due to an array of vortices travelling to the left.
sinh kh
u — \v = c + £c6e_lfc(z“ ct) + 7 к (4.8)
cos k ( z - ct) - cosh kh
ak = € (4.10)
с2 - с Г
d = ^ + Tk. (4.11)
From the equation (4.10) it follows th a t when cj > c2, th a t is when the surface is
relatively ‘stiff’, the surface elevation 77 im mediately above the vortices is negative,
i.e. the surface is drawn down by the Bernoulli effect. On the other hand when
cj < c2, th a t is for a relatively relaxed surface, the sign of a k is positive and the
surface above each vortex is raised.
When с = c0 the model does not apply, but equation (4.10) suggests th a t as
с approaches Co there is a kind of ‘resonance’, when the response of the fluid
becomes large, possibly causing the waves to break.
Proc. R. Soc. Lond. A (1994)
2179
406 M. S. Longuet-Higgins
5. Vortex ripples
From the above simple model we may predict a relation between the wave
lengths of type 1 and the type 2 capillaries. (We denote corresponding quantities
by suffixes 1 and 2.) For, the total vorticity in one wavelength of the shear layer
is given by
2tt7 = (2тг/ к 2) П Н = ( 2 п / к 2)ив. (5.1)
So
7^2 = иь « Ci (5.2)
by §3. B ut we saw th at c2 = §7 &2. Hence
сг « Jci. (5.3)
B ut for capillary waves of wavenumber к we have in general th at c2 = T k . Thus
if the type 2 waves are in resonance we have
k i / k 2 = c j/c j ~ 4. (5.4)
In other words, the wavelength of the type 2 capillaries should be about four
times th a t of the type 1 capillaries. The observed wavelengths being 0.55 cm and
0 .1 1 cm we see th at this relationship was approximately satisfied.
Equation (5.4) is valid only approximately, depending as it does on the relation
u s « Ci, which in turn depends on the observation ( a k ) 0 « 1. If for example, in
equation (3.2), (a k 0) > 1.07 we shall have u8 > Ci, so th at the surface velocity
above the toe of the instability will be negative relative to the stream below, and
mass will be transported down the forward face of the wave towards the toe. At
still larger values of ия this may cause the bulge itself to advance furthur down
the forward face, as is observed. Another reason may be th at, as the wave crest
becomes flatter owing to dissipation of the wave energy or to the advance of the
wave through the wave group, the horizontal scale of the ‘crest instability’, which
is roughly proporational to the radius of the curvature at the crest of the gravity
wave, m ust grow in time, and so advance down the forward face of the wave.
An approxim ate rule for the length of the most unstable waves in a shear flow
can also be deduced. Thus in §3, if we equate 7 к with f l H then the speed of
the instability waves is | u s. If the waves are in resonance with the free capillary
wave, then ,
\ u > * c 0 = (g /k + T k ) l' \ (5.5)
For a given value of u B, this gives us a quadratic equation to estim ate k. For
example, when u s — 4.0 then к = 2 + \/3 = 3.73, compared with the value
к = 3.41 in figure 36. However, the estim ated speed of the instability is \ u t = 2.0,
compared with the calculated value a T/ k — 0.82.
6. Flow separation
W hen the steepness (a k ) 0 of the parasitic (type 1) capillaries approaches its
limiting value 2.29 the flow will tend to separate from the sharp troughs of the
capillaries (see Longuet-Higgins 1992). Flow separation can also be expected on
the forward face of the larger gravity waves, when the tendency of the ‘crest
instability’ to overturn (see Longuet-Higgins & Cleaver 1994; Longuet-Higgins et
al. 1994) can no longer be prevented by the action of the capillarity near the toe
2 тг7 = Г \ U 2d t = \ U 2t 2. (6.3)
Jo
Hence
7 fc2 = \ U 2t 2/ A2 = \ U 2/ c2, (6.4)
where c2 is th e speed of the shear waves. B ut we saw th a t if the shear waves are
in resonance w ith th e free surface waves then | 7 k 2 — c2. So from equation (6.4)
c* = itfVc*. (6.5)
Therefore
Cj = i U, (6 .6 )
as in § 3.
Now equation (6 .6 ) is applicable not only to ripples bu t to all types of gravity-
capillary waves, including large-scale gravity waves. Since for such waves c | =
p'/fc2, where g' is the effective value of gravity allowing for the tilt of the unper
tu rb ed surface and the real acceleration of the unperturbed flow (c.f. Longuet-
Higgins 1987), we have the result th a t
A2 = n U 2/ 2 g r (6.7)
relating the wavelength A2 of the shear instabilities to the speed U of the flow a t
th e point of the flow separation. More generally, if capillarity is included, equation
(6.7) becomes
A2 = ^ [ U 2 ± (U 4 - 64g ' T ) 1' 2] (6.8)
provided th a t U 2 > 8 (s 'T )1/2.
To apply equation (6.7), for example, we have only to determ ine the velocity U
at the toe of the spilling breaker. This may be done through Bernoulli’s equation,
if we know the speed с of the gravity wave, together with the elevation Y of th e
Proc. R. Soc. bond. A (1994)
2181
408 M. S. Longuet-Higgins
toe of the spilling breaker above the mean surface level. As the particle speed at
the mean level is equal to с (see Lamb 1932, §250), U 2 may be found from
U 2 = c2 - 2g Y } (6.9)
g being the true value of gravity.
7. Discussion
We have shown th at the type 2 capillary waves observed by Duncan et al.
(1994 a, 6 ) in their experiments are very probably instabilities in the shear flow
induced by the type 1 capillaries, i.e. those propagated ahead of the toe of the crest
instability. The type 2 capillaries quickly grow and collapse into turbulence. We
may infer the existence of such a phenomenon for most gently-breaking surface
gravity waves with wavelengths between about 10 cm and 1 m. For the shear
instability to occur, the speed of the type 2 capillaries, which is about half the
speed of the type 1 capillaries, must be greater than the minimum velocity of
capillary-gravity waves (23 cm s-1). Hence the speed of the type 1 capillaries,
which equals the particle speed U in the gravity wave before instability, must
exceed about 46 c m s '1.
In §3 we considered instabilities th at were uniform in space and growing in
time. In some circumstances it may be more appropriate to consider instabilities
th a t гиге stationary in tim e but growing exponentially in a horizontal direction.
Their rate of growth will be described by a complex wave number к = kr -1- ikit
the imaginary p art k{ being related to the growth-rate cr* by k{ = сг{/ с д where cg
denotes the group-velocity dcrr /dA:. This is readily determined from graphs such
as figure 3 a.
Instabilities above the toe of the bulge are not limited to capillary waves. On
a larger scale, particularly when there is flow separation at the toe, there may
be type 2 instabilities which are essentially gravity waves. B ut their phase-speed
will be about | С/, and will be close to the speed of free surface waves of the same
wavelength.
In §4 we have of course assumed th a t the discrete vortices, though possibly
unstable themselves, remain effectively stable and two-dimensional for a certain
period of time.
It will be understood th at all of the analysis in this paper applies solely to
the initial stages of a spilling breaker. Later stages have been partly modelled
by Longuet-Higgins (1973, 1990) and by Longuet-Higgins к Turner (1974). From
qualitative observations Peregrine & Svendsen (1978) have suggested th at the
flow separation seen to occur at the toe of a well-established spilling breaker or
bore can be considered as a free shear layer or mixing layer. The break-up of such
a layer into vortices, and the subsequently thickening of the layer by amalgama
tion of pairs or triplets of vortices, was beautifully dem onstrated by W inant &
Browand (1974); see also Brown & Roshko (1974). Such mixing layers may not
be too far from the free surface to be associated directly with surface waves. Al
though some preliminary observations of the turbulence in spilling breakers have
been made by Longuet-Higgins and Donelan (see Longuet-Higgins 1974) and by
B attjes & Sakai (1981), further experim ental studies of mixing layers with a free
surface are desirable.
Proc. R. Soc. Lond. A (1994)
2182
R e fe re n c e s
Battjes, J. A. & Sakai, Т. 1981 Velocity field in a steady breaker. J. Fluid Mech. I l l , 421-437.
Brown, G. L. & Roshko, A. 1974 On density effects and large structure in turbulent mixing
layers. J. Fluid Mech. 64, 775-816.
Cox, C. S. 1958 Measurements of slopes of high-frequency wind waves. J. Mar. Res. 16, 199-225.
Caponi, E. A., Yuen, H. C., Milinazzo, F. A. & Saffman, P. G. 1991 Water-wave instability
induced by a drift layer. J. Fluid Mech. 222, 207-213.
Duncan, J. H., Philomin, V., Behres, M. & Kimmel, J. 1994 a The formation of spilling water
waves. Phys. Fluids. (In the press.)
Duncan, J. H. et al. 1994 6 Gallery of fluid motions. Phys. Fluids 6(9). (In the press.)
Karman, T. Von 1911 Fliissigkeits-u. Luftwiderstand. Phys. Zeitschr. 13, 49.
Kawai, S. 1977 On the generation of wind waves relating to the shear flow in water. A preliminary
study. Sci. Rep. Tohoku Univ. ser. 5, Geophys. 24, 1-17.
Lamb, H. 1932 Introduction to hydrodynamics, 6th edn. Cambridge University Press.
Longuet-Higgins, M. S. 1973 A model of flow separation at a free surface. J. Fluid Mech. 57,
129-148.
Longuet-Higgins, M. S. 1974 Breaking waves - in deep or shallow water. In Proc. 10th Syrup,
on Naval Hydrodynamics, Cambridge, Mass., pp. 597-605. Arlington, VA: Office of Naval
Research.
Longuet-Higgins, M. S. 1987 The propagation of short surface waves on longer gravity waves,
J. Fluid Mech. 117, 293-306.
Longuet-Higgins, M. S. 1990 Flow separation near the crests of short gravity waves. J. Phys.
Oceanogr. 20, 595-599.
Longuet-Higgins, M. S. 1992 Capillary rollers and bores. J. Fluid Mech. 240, 659-679.
Longuet-Higgins, M. S. 1993 The crest instability of steep gravity waves, or How do short waves
break? In Proc. Symp. on the Air-Sea Interface, Marseille, France, 24-30 June 1993.
Longuet-Higgins, M. S. & Cleaver, R. P. 1994 Crest instabilities of gravity waves. I. The inner
solution. J. Fluid Mech. 258, 115-129.
Longuet-Higgins, M. S. & Turner, S. J. 1974 An entraining plume model of a spilling breaker.
J. Fluid Mech. 63, 1-20.
Longuet-Higgins, M. S., Cleaver, R. P. & Fox, M. J. H. 1994 Crest instabilities of gravity waves.
II. Matching and asymptotic expansion. J. Fluid Mech. 259, 333-344.
Morland, L. C., Saffman, P. G. & Yuen, H. C. 1991 Waves generated by shear layer instabilities.
Proc. R. Soc. Lond. A 433, 441-450.
Peregrine, D. H. & Svendsen, I. A. 1978 Spilling breakers, bores and hydraulic jumps. In Proc.
16th Conf. on Coastal Energy, pp. 540-550. New York: ASCE.
Prandtl, L. & Tietjens, O. G. 1934 Fundamentals of hydro- and aeromechanics (translated by
L. Rosenhead). New York: Dover.
Stern, М. E. & Adam, Y. A. 1973 Capillary waves generated by a shear current in water. Mem.
Soc. R. Liege 6, 179-185.
Shrira, V. I. 1993 Surface waves on shear currents: solution of the boundary-layer problem.
J. Fluid Mech. 252, 565-584.
Voronovich, A. G., Labanov, E. D. & Rybak, S. A. 1980 On the stability of gravitational-
capillary waves in the presence of a vertically non-uniform current. Izv. Atm. Ocean. Phys.
16, 220-222.
Winant, C. D. & Browand, F. K. 1974 Vortex pairing: the mechanism of turbulent mixing-layer
growth at moderate Reynolds number. J. Fluid Mech. 63, 237-255.
J. Fluid Mech. (1998), vol. 364, pp. 147-162. Printed in the United Kingdom 147
© 1998 Cambridge University Press
The simple shear-flow model of Stern & Adam (1973), in which a layer of uniform
vorticity and depth overlies an infinitely deep fluid, is here extended by the addition
of an upper fluid layer of uniform thickness and constant velocity. In this way many
experimentally observed velocity profiles can be approximated. The normal mode
instabilities of such a model can be found analytically, and their properties calculated
through the solution of a quartic polynomial equation. The dispersion relation is here
determined and illustrated in its dependence on the Froude number and on the ratio
Я 1/ Я 2, where Hi and H 2 denote the mean depths of the surface layer and the base
of the shear layer, respectively. It is found that two branches of instability which are
distinct when Я 1/ Я 2 is moderate or small can become merged when H i / H 2 ^ 0.4924.
Also calculated are the fastest-growing modes, and their wavelengths. The results are
applied to some examples of surface flows generated by towed bodies, and to steady
spilling breakers.
1. Introduction
The theory of the stability of surface shear flows has important applications to the
study of surface wakes behind ships or towed bodies (Dimas & Triantyfallou 1994);
to the stability of the crest of a spilling breaker (Coakley & Duncan 1997); and to
wind-drift currents at the sea surface (Stern & Adam 1973). Such currents are usually
unstable, and over large horizontal distances may develop into a turbulent mixing
layer. But over short distances they may usefully be treated as steady horizontal
shear layers; see for example Peregrine (1974). Theoretical interest then focuses
on the normal instabilities of such layers, especially those with the largest growth
rates, which can give an indication of the horizontal length scales of the observed
perturbations.
The presence of a free surface is an additionally interesting feature. For, whereas
mixing layers in an unbounded fluid have been well studied (see especially Winant &
Browand 1974), those near a free surface are less well understood.
The simplest theoretical model of a surface shearing current was that due to Stern
& Adam (1973), who assumed a surface layer of uniform thickness and vorticity
overlying a stationary layer of infinite depth. Other models or developments of
the same model are due to Kawai (1977), Voronovich, Lobanov & Rybak (1980),
Milinazzo & Saffman (1990), Caponi et a l (1991), Morland, Saffman & Yuen (1991)
and Shrira (1993).
Recently Dimas & Triantyfallou (1994) have computed the linear stability of a
shear flow with a continuous velocity profile of the form
u = U s e c h 2 by (1.1)
2184
148 M. S. Longuet-Higgins
(see figure 9 below) and have followed the nonlinear development of a typical
sinusoidal instability. It is the linear aspect of the problem which will be considered
here. To be precise, Dimas & Triantyfallou find, as in an earlier paper (Triantyfallou
& Dimas 1989) that the normal instabilities of the profile (1.1) appear to be of
two general kinds: Branch I, which are im portant at low wavenumbers, and Branch
II which are im portant at high wavenumbers for low Froude numbers, or at low
wavenumbers for high Froude numbers. However, their method o f calculation is
elaborate and their numerical results are consequently incomplete. There seems to be
good reason for adopting a much simpler model, representing the observed current
profile almost equally well, but amenable to an analytical treatment.
In fact such a model is readily available. It involves simply an extension of the
shear-flow model of Stern & Adam (1973) by the addition of a layer of uniform flow
immediately below the surface. The model profile depends on three parameters: the
surface current U and the depths Hi and H2 of the two upper layers. One particular
com bination of these parameters fits equation (1.1) reasonably well. Moreover, there
is no difficulty in exploring other values of the parameters since the solution to
the stability problem is now reduced to determining the roots o f a simple quartic
polynomial. Not only is this procedure much faster than a purely numerical method,
it is also more complete and accurate, and reveals the relation between different
branches of the solution in a more transparent way.
In one special case (H\ = 0) the flow reduces to that of Stem & Adam (1973), and
in another case (Hi = H2) it reduces to a thin shear layer at a depth H 2 below the
free surface.
The plan of the paper is as follows. Section 2 introduces the model and the
basic equations. Section 3 treats especially the boundary conditions, and derives the
dispersion relation. This is then discussed, with examples, in §4; see figures 2 to 5.
The boundaries of unstable regions are shown in figure 6 . In §5 we discuss some
special cases. Some comparisons with continuous velocity profiles are given in § 6 and
in §§ 7 and 8 we apply the theory to the flow observed in a steady spilling breaker. A
general discussion follows in § 9 .
2. Basic equations
Consider a volume of inviscid, incompressible fluid whose velocity и is directed
horizontally in the x-direction, and depends only on the vertical coordinate y \ see
2185
Then, if the surface tension T is neglected, the basic flow depends only on two
dimensionless numbers: H \ / H and F.
We shall consider sinusoidal perturbations of this flow travelling with phase speed
с to the right. In a frame of reference travelling with the phase speed the two
components (u, u) of the fluid velocity will be given by expressions of the following
form. In the upper layer,
и = ( и —с) + ЩВе** - В е - к”)е*х, \
v= к(Аеку + Bc~kf)e'kx. ] ‘
It is understood that on the right the real part is to be taken. In the intermediate
layer,
u = Q ( y + H 2) - c + ik(C<*> - De~k> )e^ , \
u= *(Ce*' + D e-^e**. J l '
In the lowest, semi-infinite layer
u = —c + i*Ee*'+ftjc,
(2.7)
v= kEeky+'kx.
;}
These expressions satisfy the equation of continuity d u /d x + d v / d y = 0. The vorticity
is given by
c = S - | = ° - - fl ’ ° (z8)
in the three regions respectively.
3. B o u n d a ry co n d itio n s
It is assumed that at the free surface the pressure is a constant (zero), and that at
the two interfaces both the pressure and the normal velocity are continuous.
The horizontal derivative of the pressure p is found from the momentum equation
150 M. S. Longuet-Higgins
where a prime denotes the perturbation velocity. Using the equation of continuity we
then have
(3.3)
ox dy
At the free surface у = r\, say, the vertical velocity v is equal to u d tj/d x f . In
the linearized theory the boundary condition at the surface can be replaced by the
condition that
p = ( g q - T d ^ / d x 2) (3.4)
at the mean surface level у = 0. Differentiating with respect to x then gives us
и — gv — T d 2v / d x 2 on у — 0. (3.5)
( U — c)2-^—— gv on у = 0. (3.6)
dy
At the interface between the two upper layers, continuity of the normal velocity
implies, to first order, the continuity of v . This and the continuity of d p / d x gives us
Ce~kHl + DekHli =
= E e - klH\ )
(3.9)
c2(Ce~kHl - DekHl) + c(Q /k)(C e~kH' + De*H ) = c2£ e - ‘H!.J
t This is an approximation, based on the assumption that the rate o f growth o f the instability is
sufficiently small.
2187
X\A + В X\C + D,
( Z + q ) ( X 2C - D ) p(X2C + D) = ( Z + q ) X 2E.
Eliminating E between the last two of these equations gives us
( Z + q ) D - t f ( X 2C + D ) = 0 (3.14)
and we are left with four linear equations for А, В, С and D with matrix
/ ( Z 2 - 1) ~ ( Z 2 + 1) 0 0 \
Xi 1 -Xy -1
(3.15)
X{Z -Z -x ^ z+ fi) (Z -P )
\ о 0 - \ X 2fi (Z + q -tf)J
The vanishing of the determinant of this matrix gives an equation for Z and hence
the dispersion relation.
where
_ l (4.2)
p-
On expansion of the determinant, equation (4.1) reduces to
A,Z(Z2- l ) ( Z + 4 )
+ Я,(Я, - A 2) a Z ( Z 2 + l)
152 M. S. Longuet-Higgins
Figure 2. The radian frequency <rr and rate of growth <7,- as functions of к in dimensionless units,
when h\ = 0.5, F — 1.5.
by (oT/ к and the rate of growth by co(, where cor and со,- are the real and imaginary
parts of со. It is appropriate to plot cor and ct>,- as functions of к for given values of
the Froude number F = U / ( g H ) i/2 and the depth ratio H \ / H .
For convenience in presentation we introduce the dimensionless wavenumbers and
frequency
к = kH, (a r + <7j) = (ci>r + cOi)H/U (4-5)
and also the depth ratios
h , = Я ,/Н , h2 = H2/ H (4-6)
so that h i + h 2 = 2 .
A typical example, when hi = 0.5 and F = 1.5, is shown in figure 2. For large values
of к there are four real roots Z yielding four separate branches of <rr (figure 2d).
2189
However between Q and R two of the roots become a conjugate complex pair, and
the segment QR is doubled; this is indicated by a ‘parallel’ mark. Similarly between 0
and P another branch is doubled. In each of the intervals OP and Q R , is non-zero
(figure 2b).
Figure 3 shows a second example, hy — 0.7, F = 1.5. In this case the two points P
and Q have coalesced into PQ. The double branch is now continuous through P Q \
the single branch crosses it with no exchange of stability.
A third example, F — 0.5, hi = 0.3 is shown in figure 4. This case is similar to figure
2, except that Q and R have moved very close together. A blow-up of the region QR
(figure 5) shows that there is a small ‘bubble’ of instability between Q and K.
To obtain a synoptic view, we have plotted in figure 6 (a) the wavenumbers к
corresponding to the points P , Q, R, that is to say the points of marginal stability, for
2190
M. S. Longuet-Higgins
FIgure 6. Synoptic stability diagrams: full curves, loci of marginal stability; crosses, points of
maximum growth rate at constant Froude number F. (a) fi, = 0.3, (b) hi = 0.5, (c) /ii = 0.7, (d)
h\ = 0.9.
values of the Froude number F (on the vertical axis). It will be seen that there are
two distinct areas of stability and two areas of instability. Similarly in Figure 6 (b),
when h\ = 0.5. However in figure 6 (c), when h\ = 0 .7 , the two areas of instability
have joined in the middle, creating altogether three separate areas of stability. This
situation is accentuated in figure 6 (d), when h\ = 0.9. Now two of the stable areas have
moved far to the right, and only a very thin wedge of stability at small wavenumbers
remains.
The pinch-off point between figures 6{b) and 6 (c), when the two unstable areas just
meet, can be determined numerically (see the Appendix) as hx = 0.6599 (i.e. H J H 2 =
0.4924), and F = 1.5361. The dispersion relation for this particular configuration is
shown in figure 7. Three modes coincide at the wavenumber к = 1.1091.
In each of figures 6 (a) to 6(d) the wavenumbers corresponding to maximum rates
2191
R g u r e 7. crr and <7, as functions of к for the triple-point mode at h\ = 0.6599, F = 1.5361.
1
tT
Branch I
of growth (<T/)raax have been marked by cross plots. In figures 6 (c) and 6(d) the circular
plots along the arc Л В correspond to minimum growth rates \<rt\.
Of special interest are the wavelengths (2n /k) of the fastest-growing modes. The
wavenumbers к are shown in figure 8 (a) as functions of F, for depth ratios h{ ranging
from 0.1 to 0.7. Generally, the wavenumbers increase with increasing hu but decrease
with increasing F, both for Branch I and for Branch II.
The corresponding rates of growth are shown in figure 8 (b). Again, the rates of
growth generally increase with increasing depth ratio, that is, with decreasing thickness
of the vortical layer. The growth rates on Branches I and II are of the same order of
magnitude.
2192
156 M. S. Longuet-Higgins
5. Limiting cases
As hi -*> 0 so that the upper layer becomes infinitesimally thin, it is found that the
branch point P in figure 2 moves leftwards towards the origin O. In the limit, the
left-hand curve in figure 6 (a) coincides with the vertical axis к = 0. Setting X\ — 1 we
find that equation (4.3) becomes
Z (Z 2 - 1)(Z + q) + (1 - Л2)а Z (Z 2 + 1) + 2a [q + (X2 - l)a ]Z 2 = 0 (5.1)
which has a zero root Z . Removing this root we are left with a cubic equation which
reduces to
Z 3 + (m + 2<j)Z2 - ( l + 2 a m )Z + m = 0, (5.2)
where
m = a (1 —A2) — q. (5*3)
This is the same equation as found by Stern & Adam (1973).
Suppose on the other hand that Iii tends to 1, so that (H 2 — H Y) / H -*• 0 and the
shear layer becomes very thin, while remaining at a finite distance H below the free
surface. Then we have
Ai —A2 ~ 2 k (H2f c—
-H iU
(5.4)
Q = U / ( H 2 -Я 0 J
and hence а = Q /2kco becomes large while
Thus о has two finite values, which correspond to waves travelling at a speed
±[(g/& )tanhfcHi] 1/2 relative to the surface current U. They are similar to gravity
waves in the upper layer, propagated as though this shear layer were a rigid boundary.
The remaining roots of equation (4.3 ) tend to infinity as (Я 2 — H \ ) / H —►0 for
finite k. They represent instabilities of the shear layer itself. As shown in Batchelor
(1967, Section 7.1), for example, a shear layer of thickness Ah in an unbounded fluid
is unstable to all modes with wavenumber к of order less than (Ah)~l . In the presence
of a free surface, however, the modes can be significantly modified.
Thirdly let us consider the case of low Froude number F. This is the ‘rigid-lid’
approximation, when the free surface is assumed to be planar. Then q and a are both
of order F -1/2. Equation (4.3) has two roots
Z ~±1, c~ U ± c0 (5-9)
which correspond to surface gravity waves travelling in either direction with high
speed c0 relative to the surface current U. The remaining roots are approximately
2193
F i g u r e 9. The velocity profile given by equation (1.1) (continuous curve), and its approximation by
the model profile (2.1).
e* = V 2 - l (6 . 1)
when у — —1, and hence
b = 0.8814. (6.2)
To approximate this profile we may draw a tangent to the curve at the mid-point
(u,y) = { \ U , —1) and use this to represent a layer of constant shear fl. The velocity
2194
158 M. S. Longuet-Higgins
R g u r e 10. The growth rate or, as a function o f wavenumber for the model profile in figure 9.
and hence
- h x = 1 - ( b j . 2) ' 1 = 0.1977. (6.4)
In figure 10 we show the growth rates cr, calculated as in §5, when h{ takes the value
(6.4), and for Froude numbers F = 0.5, 1.5, 2.5 and 5.5. The two families of curves
are qualitatively similar to those calculated numerically for the velocity profile ( 1. 1)
by Dimas & Triantyfallou (1994), and shown in their figure 9. Apparently they were
unable to present accurate results when crf < 0.01. As noted by Morland et al. (1991)
for the simpler case of the wind-induced drift current, the piecewise linear solution
has narrower bands of unstable modes and somewhat higher growth rates.
As a second application consider the velocity profile
u = U [ I - t & n h (by2/ b 2)] (6-5)
representing the shearing current in the second wave trough behind a steady breaking
wave (Duncan & Dimas 1996). Following the same procedure as before we find that
the mid-point of the current occurs when
t byX/Jl = 7 3 , y = d (6 .6 )
and hence
b = 0.5493. (6.7)
Fitting the shear layer by a tangent at the mid-point of the velocity profile we find in
a similar way that
hi = 1 —(3b)"1 =0.3132. (6.8)
F i g u r e 11. (From Coakley & Duncan 1996). Observed fluid velocities relative to a 15 cm hydrofoil
towed at a speed of 69 cm s-1 to the left; average of ten runs.
The Froude number F — U / { g H ) l/2 may be determined from their figure 16, in which
U = 14.8 cm s-1 and H = 3.9 cm, giving F = 0.239. The dispersion relation, which is
similar to figure 4 above, shows there is one maximum growth rate at a dimensionless
wavenumber K\ — 0.338 and a very small bubble of instability at k 2 = 18.3. These
correspond to dimensional wavelengths 2n /k\ = 72.5 cm and 2п /к2 = 1.34 cm
compared with the two wavelengths of 36 cm and 1.4 cm calculated by Duncan and
Dimas (1996)|. The present calculations indicate that the rate of growth of the second
instability was exceedingly small.
f It may be noted that the authors adopt a different definition of F from that given in §4 above
or in Triantyfallou & Dimas (1989).
160 M. S. Longuet-Higgins
X H, H2 U H L\ l2
(mm) (cm) (cm) (cm s-1) (cm) hi F Ki (cm) *2 (cm)
20 0.20 0.42 66 0.31 0.64 3.8 0.13 15 1.24 1.6 (1.6)
38 0.25 0.79 58 0.52 0.48 2.6 0.29 11 1.07 3.1 (3.0)
56 0.27 0.85 49 0.56 0.48 2.1 0.27 13 1.16 3.0 (2.9)
74 0.27 1.17 45 0.72 0.37 1.7 0.24 19 1.21 3.8 (3.5)
92 0.27 1.32 45 0.79 0.34 1.6 0.23 22 1.25 4.0 (3.8)
T a b le 1. Fitted parameters for the current profiles in figure 11, and the predicted dominant
wavelengths
x
X
1.54
Ft")
1.52
X
X
j---1---1---1---1---1---1---1---1---1---1---1---1---1---1---1---1---1---L
0.6599 0.66
A.
F i g u r e 12. Plot of F(l) and F(2) as functions o f hi, close to the critical point.
the scope of this paper, though a nonlinear theory might well be based on the present
simplified model.
In this paper we have considered only two-dimensional instabilities. The three-
dimensional instabilities of a surface shear layer are likely to be of considerable
interest under typical oceanic conditions. Such instabilities could well be investigated
by means of a similar simplified approach.
162 M. S. Longuet-Higgins
For any given value of hi slightly exceeding the critical value, one can solve equations
(A 5) for F and к by Newton’s method of successive approximation. There are two
solutions (F (1),k (I)) and (F(2),k (2)), in general. Close to the critical point, (F (n),/ii) lies
closely on a parabola; see figure 12. The critical values of hi and F correspond to the
vertex of the parabola, and similarly for к(п) and h i.
REFERENCES
B a t c h e l o r , G. K. 1967 An Introduction to Fluid Dynamics. Cambridge University Press, 615 pp.
C a p o n i, E. A., Y u e n , H. С., M i l i n a z z o , F. A. & S ajffm an, P. G. 1991 Water-wave instability induced
by a drift layer. J. Fluid Mech. 222, 207-213.
C o a k le y , D . B. & D u n c a n , J. H. 1997 The flow field in steady breaking waves. Proc. 21st Symp.
on Naval Hydrodynamics, Trondheim, Norway, 24-28 June 1996, pp. 534-549. Washington, DC,
Natl Acad. Press.
D im as, A. A. & T r i a n t y f a l l o u , G. S. 1994 Nonlinear interaction of shear flow with a free surface.
J. Fluid Mech. 2 6 0 , 211-246.
D u n c a n , J. H. & D im a s, A. A. 1996 Ripples generated by steady breaking waves. J. Fluid Mech.
329, 309-339.
K a w a i, S. 1977 On the generation o f wind waves relating to the shear flow in water. A preliminary
study. Sci. Rep. Tohoku Univ. 5, Geophys. 24, 1-17.
L o n g u e t - H i g g i n s , M. S. 1994 Shear instability in spilling breakers. Proc. R. Soc. Lond. A 4 4 6 ,
399-409.
M i l i n a z z o , F. A. & S a f fm a n , P. G. 1990 Effect of a surface shear layer on gravity and gravity-
capillary waves o f permanent form. J. Fluid Mech 2 1 6 , 93-101.
M o r l a n d , L. С., S a f f m a n , P. G. & Y u e n , H. C . 1991 Waves generated by shear layer instabilities.
Proc. R. Soc. Lond. A 433, 441-450.
P e r e g r i n e , D. H. 1974 Surface shear waves. J. Hydraul. Div. ASCE 1 0 0 , 1215-1227.
S h r i r a , V. I. 1993 Surface waves on shear currents: solution o f the boundary-layer problem. J.
Fluid Mech. 252, 565-584.
S t e r n , М. E. & A d a m , Y. A . 1973 Capillary waves generated by a shear current in water. Mem.
Soc. R. Liege 6, 179-185.
T r i a n t y f a l l o u , G. S. & D im as, A. A. 1989 Interaction o f two-dimensional separated flows with a
free surface at low Froude numbers. Phys. Fluids A 1, 1813-1821.
T \ j r n b u l l , H. W. 1 9 5 7 Theory o f Equations, 5 th ed n . E d in b u r g h : O liv e r an d B o y d , 166 pp .
V o r o n o v ic h , A. G., L o b a n o v , E. D. & R y b a k , S. A. 1 9 8 0 O n th e s ta b ility o f g r a v ita tio n a l-c a p illa r y
w a v e s in th e p r e se n c e o f a v e rtic a lly n o n u n ifo r m cu rren t. Izv. Atmos. Ocean Phys. 16, 2 2 0 - 2 2 2 .
W i n a n t , C. D. & B r o w a n d , F. K. 1 97 4 V o rtex p a ir in g : th e m e c h a n is m o f tu r b u le n t m ix in g -la y e r
g r o w th a t m o d e r a te R e y n o ld s n u m b er. J. Fluid Mech. 6 3 , 2 3 7 - 2 5 5 .
2199
P a p e rs N1 to N 8
every third w ave crest is flattened, the next is sharp-pointed and the third is
interm ediate. T he rate o f grow th o f the instability is m easured, and is found to be
practically independent o f the initial w ave steepness.
N. Standing Waves
2203
J. Fluid Mech. (2000), vol. 423, pp. 275-291. Printed in the United Kingdom 275
© 2000 Cambridge University Press
1. In tro d u c tio n
The equations of motion for the irrotational motion of an ideal, incompressible fluid
with a free surface are commonly derived from a Hamiltonian expression, in which the
surface displacement and the velocity potential at the surface are taken as canonical
variables (Zakharov 1968). The evaluation of the potential function generally requires
a Hilbert transformation, and especially in high-order expansions in powers of a
parameter the coefficients in these expansions become extremely complicated; see for
example Glozman, Agnon & Stiassnie (1993). A very much simpler and more natural
system of equations has been introduced by Balk (1996), based on a Lagrangian. In
this system not only are the kinetic and potential energies polynomial expressions of
finite degree « 4 ) in the independent coordinates (as opposed to infinite order in
the case of a Hamiltonian system) but the coefficients in these expressions are mainly
low integers. This makes the equations of motion relatively simple to program for
numerical computation, and facilitates the discussion of various approximations.
The Lagrangian scheme introduced by Balk for any general, time-dependent defor
mation of the free surface (including overturning waves) is in fact a generalization
of the cubic system of equations for steady, progressive Stokes waves discovered by
the present author (Longuet-Higgins 1978). These equations also were found to be
derivable from a low-order polynomial Lagrangian (see Longuet-Higgins 1985).
However, the comparative simplicity of the general Lagrangian system has so far
remained unexploited. The purpose of the present sequence of papers is to show how
Balk’s (1996) analysis can be developed, and to examine some of its applications. We
begin in the present paper with the case of standing waves, periodic in space but not
2204
276 M. S. Longuet-Higgins
necessarily in time. General equations are derived in §§ 2 to 4. An im portant part in the
analysis is played by the determ inant Л of the equations of motion, which must not
vanish in order for time-stepping to proceed. It is shown in § 5 that this determinant
factorizes. Some insights can be gained by considering low-order approximations (i.e.
truncations o f the Fourier series) as is shown in §§ 6 to 8 . High-order approximation
will be used in later papers for a discussion of two phenomena displayed by standing
waves, namely breaking at the wave crests and the forming of strong vertical jets
(‘flip-through’) rising out o f the wave trough.
2. The Lagrangian
It is shown by Balk (1996, § 2) that in any two-dimensional irrotational wave motion,
periodic in the (horizontal) x-direction with period 2л, the coordinates (X , Y ) of a
point on the free surface may be expressed in the form
oo
X + i Y = « + ^ ( X n + i r „ ) e - in{, (2.1)
—00
where X_n = X* and Y_„ = Y*, a star denoting the complex conjugate; also
X n = i c nYn, (2-2)
where a n equals + 1 , - 1 or 0 as n is positive, negative or zero respectively. From (2.1)
it follows that
00
—CO (2.3)
У = £ у „ е - “ <,
so that the Yn(t ) may be taken as the time-dependent coordinates of the fluid. These
are subject to the constraint that the mean level Y, given by
Уо = - £ |п|У„У_. (2-5)
-0 0
i< r „ B n e
—
OO ( 2.8)
00
—00
Then the kinetic energy density T is given by
l г2к 00
2T = 2 iij0 = H N B»B— (2.9)
The crucial step taken by Balk (1996) was to show that the kinematic boundary
condition at the free surface can be expressed as
y>( = X t Y( - Y tX ( , (2.10)
which allows the coefficients Bn to be related directly to the Y„ by
3. S ta n d in g waves
So far the equations are quite general. However, in standing waves the coordinates
Y„(t) may be chosen to be all real. The above formulae are then simplified and, from
equations (2.8) to (2.13), we find
4 Г = {[{(1 4* Я2 —flifli)<*i + \i.a \ + —0102)02 + |(o 2 + 04 — 0103)03 H-----]2
+ (ЗЛ)
or more compactly
00/00 \ 2
4 T = ^ 2 ( X / ^ mn йи J *
m = 1 \n = l /
2206
278 M. S. Longuet-Higgins
where
(3-3)
provided that a„_m is replaced by 0 when n < m. Writing equation (3.2) in the form
(3.4)
where
1 00 (3.6)
Qki = 2 ^ ,nk
m=l
In other words, the m atrix of the quadratic form (3.5) is the column-by-column
product of the matrices (Р„*) and (Pm/). Note that P„* is of maximum degree 2 in
the coefficients an. Hence T is of degree 2 in and of maximum degree 4
in fli, Й2, . . . .
Consider now the degree o f the terms in the potential energy V , given by equation
(2.7). The right-hand side o f (2.7) still contains the dependent variable Yo- The first
sum m ation contains term Y02, while the second contains some terms with no more
than one o f J, m, n equal to 0 (for, if any two of /, m, n are zero, then so are all three,
making |/| vanish). Since by (2.5) Yo is quadratic in the remaining Yn, altogether V
has maximum degree 4 in Yi, Y2, —
(4.1)
(4.3)
and therefore
(4.4)
where
(4.5)
2207
^ 2 Qm - X) X T ^ (4-7)
/ / к
where
<«>
Note that when к = n or / = n, two of the terms in this expression cancel.
In order to calculate d Q u /d a n we have from (3.6)
<«>
while from (3.3)
^ = 0
dan
+1 if n = I- m
+1 if n = I+ m
—am if n= I
—ai if n = m. (4.10)
Lastly for the terms involving V in equations (4.2) and (4.7), note that
d v _ * d v d Y k
8a„ £ д у к да„ ’ (4Л1)
к ——oo
where
-i- if k = ± n
д П = < 2|n| (4.12)
да„ - Y k if * = 0
0 otherwise;
also that
1!
HI YmY„
5 (|/| Y,YmY„) = < |/|
I/I Y,Y„ if
i ( kк = m (4.13)
dYk
*-4
II
I/I Y,Ym
a
280 M. S. Longuet-Higgins
5. Solution of the equations
In order to solve equations (4.14) we may suppose an(t) = 0 for all n greater than
N, say, and proceed by successive approximations as N is increased indefinitely. Given
some starting values for an and a„ at an initial time t = 0, equations (4.13) comprise
a set of simultaneous equations for determining йь й2,...й ^ . These can be solved
provided that the determ inant
Д * -К М (5-«
does not vanish. Some interest therefore attaches to the conditions under which
AN = 0 . (5.2)
Д* = 2 n (5.3)
so that the vanishing of A N depends solely on the vanishing of ||Py| Moreover from
equation (3.3) we have
(5.4)
= (N \)V 2^ n’
where DN is the (N x N ) determ inant
(1 + a 2 — a ia i) (ax + a3 — a {a2) (a2 + (З4 — a\a$)
( a3 - a 2a {) (1 + 0 4 - 02 ^ 2) (a\ + as — a2a$)
Dn = (5.5)
( ал - а га х) ( а $ - а ъа2) (1 + a 6 - a 3a3)
On subtracting from the zth column of the determinant a{ times the first column we
reduce each element of the first row to 0, except for the first element which is 1. It
follows that
Dn = En . (5-7)
Now in E n the sum of the elements of each column is
v-^\(+) (5.8)
1 + a x + a2 H-------1- a^ = ^ ,
say. Each element in the first row may therefore be replaced by Hence 5D a
factor of the determinant E n . N o w multiplying the second row by 2, adding to this
2 x each of the other even rows and subtracting from this the first row, we see that
2209
E (~) = 1 - a x + a2 - a3 + a4 ------. (5 .9 )
Hence £ (-) is also a factor of EN. After extraction of these two factors the remaining
determinant is
1 1 1 1
-1 1 -1 1
ЯЗ a4 as ( 1 +*б)
‘■ ■ а д Е ’ Г Ч ' p")
As will be seen below, the vanishing of £ (+) or is associated with the formation
of a cusp at the free surface.
If TN and VN denote the truncated forms of the kinetic and potential energies, and
if = TN — VN we note that an exact integral of the equations
m v s t’
is
HN = TN + VN = constant. (5.13)
Equation (5.13) may be derived on multiplying each side of (5.12) by an and then
summing with respect to n from 1 to N. For, if / is any function of a i , . . . a N and
formally independent of the time t, we have identically
S - t( ‘S +8-S)
-I ( £ * ■ ! : ) ,5Л4>
When / = <?n the last group of terms vanishes by (5.12), and since TN is quadratic
in A t,... 6 jv we obtain
282 M. S. Longuet-Higgins
(x-n )
6 . T h e case N = 1
In the simplest case we set a2 = a 3 = • • • = 0, so that the free surface profile is
given by
X = t + o tsin £ , | (61)
у = Y0 4 - a cos J
where а = ai and Yo = —5 a2. This is the parametric equation of a cycloid; see figure
1. Physically admissible solutions are limited to |a| < 1. When a = ±1 the profile has
a downward-pointing cusp.
The dynamical equations take into account only the self-interaction of the funda
mental harmonic. Thus from § 2 we have
4T, = (1 —а 2)2а2, 1
( 6 .2)
Щ = a2- \<x\ J
0.8
0.6
t
0.4
0.2
0
-2 - 1 0 1 2
a
F i g u r e 2. Graphs o f a(f) covering half a wave period, for different values of ao = a(0),
assuming that dc(0) = 0.
and
( i - « 2)
da. (6.7)
- f [4Я —a2(l — | a 2) ] 1/2
Equation (6.7) can also be written as
t = ± ^ L [(1 — a 2) 2 — C 2] ‘/2 d a ’ ( 6 8 )
where
C 2 = (1 —(Xq)2 = 1 —8 Я. (6.9)
Figure 2 shows a as a function of the time t for various values of ao = a(0), assuming
that at t — 0 the fluid is at rest. When a 0 is small, a(t) is simply a sine-wave with
period 2k. Only the section of each curve for which a > 0 is shown. In the hypothetical
limiting case ao = 1 we have С — 0 and so equation (6.9) gives
*= ±V 2( l - a ) , (6 . 10 )
that is to say a pair of straight-line segments. The crest of the wave then rises or falls
uniformly when t Ф 0, with an instantaneous, large negative acceleration at t — 0.
7. The case N = 2
On writing ai = a and a2 = P, the free surface is now given by
x = £ + asin<^ + ! 0 s in 2 £ ,
(7.1)
у = Уо + acos^ + {P cos 2 £,
where
Y0 = - W - \ f . (7.2)
2212
284 М . S. Longuet-Higgins
Outer limits for a and p are given by the vanishing of A# in equation (5.11). Since in
(5.10)
1 1 1
-1 1 -1 = (!-/* ), (7.3)
0 1
the outer limits are given by
( l + a + /? )(l-« + /? )(l-0 ) = O ( 7- 4 )
where
L = 2 ( 1 + 0 —a2)2 + a 20 2,
N = № ( 1 - P ) 2 + \ { 1 - P 2)2,
2214
286 M. S. Longuet-Higgins
and
Ci = 2a(l + /? —a2) —a/J2, 1 (?9)
C2 = a2(l —Д) — |/?(1 —/52). J
Some examples of the trajectories of (a,/?) are shown in figures 5 to 7. In each case
the motion starts from rest: a(0) = /?(0) = 0, so that the total energy H is just equal
to the initial potential energy K(0 ).
For disturbances of very small initial amplitude one would expect the coefficients a
and /? to oscillate independently, each with its own frequency. Thus a would oscillate
harmonically with frequency 1 and /? with frequency J 2 = 1.414... . The ratio of
the frequencies being irrational, the combined motion would be non-periodic, in the
small-amplitude limit. However, very slight nonlinearity may cause the frequencies to
become compatible, resulting in a periodic orbit. One such example is shown in figure
5(a) where the initial amplitudes are
a = -0.4, p = 0. (7.10)
2215
Figure 6. (a) Trajectory of (a,fi) when N = 2 and with starting values a = 0.4, fi = —0.4,
a = P = 0. t = 0.0(0.1)12.7. (b) Graph of a(t) for (a). (с ) Graph of 0(t) for (a).
The crosses mark points on the trajectory separated by time-intervals 0.1 (one hundred
time steps Ar). After time t = 22.3 the point (a,/?) is close to the opposite point (0.4,
0.0) where, because of conservation of energy, the velocities a are very small. At time
t = 44.6, (a, p) has returned practically to its initial position. In the intervening time,
a has executed seven complete oscillations and P has executed ten (see figures 5b and
5c). So the ratio o f ‘frequencies’ is 10 ч -7, i.e. 1.429 approximately.
A simpler example is shown in figure 6 (a), where the starting point is
a = 0.4, jJ = -0.4. (7.11)
At t = 12.8, (a ,P) returns to its starting point, a having executed two oscillations
and p three, the ratio of frequencies being 1.5. Compared to a(f), the form of p{t) is
remarkably sinusoidal.
A third example is shown in figure 7(a). Here the initial amplitudes are
a = 0.5, P =0. (7.12)
The initial energy is sufficiently large that after just over half an oscillation of the
fundamental a(r) the trajectory of (a,/?) runs off the scale. The point (a,/?) accelerates
toward the boundary, as is shown by the curve of a(t) in figure 7(b). The critical
2216
288 M . S. Longuet-Higgins
time t at which (a,/?) crosses the boundary is tc = 4.211 approximately. The final
acceleration |a| as t approaches tc is shown in figure 7(c) on a logarithmic scale. The
acceleration varies as \t —tc\x, where Я is a constant. This behaviour is typical o f any
function f ( t ) satisfying an equation of the form
/ / = C f2 (7.13)
when f passes through a zero.
In all of the above numerical integrations the total energy H remained a constant
to within one part in 10 5.
8 . T h e case N = 3
This last special case to be discussed is the simplest for which not all the factors
of A n are linear. Thus from equation (5.11) on writing N = 3 and a i,a 2>fl3 —
respectively we have
where by (5.10)
0 1 0 1
-1 1 -1 1
F< = (8 .2 )
P У 1 a
У 0 0 1
which on expansion gives
( 1 - /? ) + ( « - -y)y (8.3)
Whereas in (8.1)
l+ a + (8.4)
+
0
II
and
l- a + 0 -y = O (8.5)
represent bounding planes in the space of (a, f$, y), the last factor
(1 - Д ) + ( а - у ) у = 0 (8 .6 )
represents a quadric surface (hyperboloid). Nevertheless the intersection of this surface
with any plane у = constant is a straight line in that plane. When у — 0, of course,
equations (8.4), (8.5) and (8 .6 ) reduce to the lines aa, bb and cc of figure 3. In the
more typical case when у = 0.5, these lines appear shifted as in figure 8 . Thus aa
is lowered by an amount y, bb is raised by у and cc is tilted with inclination у and
passes through the point { a j ) = (y,l). We may call the triangle formed by aa, bb
and cc the basic triangle.
The origin (0,0), which corresponds to infinitesimal orbits, always lies in the
interior of the basic triangle so long as |y| < 1, and it can be verified that points (a,/?)
lying on the sides of the basic triangle always correspond to surface profiles having
downwards-pointing cusps.
We show one example of a time-history a (t)yfi{t)>y{t) when the initial conditions
2218
290 M . S. Longuet-Higgins
T ’ V -------------------- i ' i — 1-------- ----------1-------- ---------Г------- ---------1 . J. ’ T — •— 1—
♦♦ ♦
0 .6 ♦ ♦ -
♦ ♦
4 ± ♦ ♦
0 .4 ♦
tl ♦ ♦ ♦
4 ♦ ♦
♦ * j\
V + + + + + + H rt^ 0 .2
♦ - + ii ♦ /= 1 0 ♦ % j
f = 10 + + +
I * + + + + , = 0 Л
♦ J;.........................
\ . r - o
.
'
- ...... * ................................................. 1 r 0
+ ♦ Г **
w !! ♦ / - 0 .2
\ ♦♦ *
♦ I**♦
- 0 .4
*
■ (a) - 0 .6 . 0ь ) V
i . t i i . i 1 1 . I i - » - i - i
N>
04
о1
- 0 .4 - 0 .2 i0 0 .2 0 .4 0 0 .2 0 .4 0 .6
0I
0
a a
t
R g u r e 9 . (a ) Trajectory o f (a , P) when N = 3 and with starting values a = 0 .4 , p = у = 0,
a = p = у = 0 (compare figure 5a). (b ) Trajectory of (a .y ) corresponding to (a), (c) Root-mean-square
values o f a, P and у as functions of the time t.
are th at at time t = 0
a = 0.4, P=y = 0 (8-7)
and а, 'р,у all vanish as before. A plot of p vs. a alone is shown in figure 9(a), and
the corresponding plot of у vs. a in figure 9(b). Comparing figure 9 (a) with figure
6 (a), where the initial conditions were similar except that у was restricted to be 0 , we
see that trajectories of (a,/?) appear very different and the motion is less periodic. An
approximate complete cycle 0 < t < 44.0, is shown in figure 9(c).
It will be noted that the r.m.s. values of a, p and у are all of comparable magnitude.
This work has been supported by the US Office of Naval Research under Contract
N00014-94-1-0008.
R EFEREN CES
A g n o n , Y. & G lo z m a n , M. 1996 Periodic solutions for a complex Hamiltonian system: New
standing water waves. Wave motion 24, 139-150.
B a lk , A. M. 1996 A Lagrangian for water waves. Phys. Fluids 8, 416-420.
G lo z m a n , М ., A g n o n , Y. & S tia s s n ie , M . 1993 High-order formulation of the water-wave problem.
Physica D 66, 347-367.
L o n g u e t - H i g g i n s , M . S. 1978 Some new relations between Stokes’s coefficients in the theory of
gravity waves. J. Inst. Maths Applies. 22, 261-273.
L o n g u e t - H i g g i n s , M. S. 1985 Bifurcation in gravity waves. J. Fluid Mech. 151, 457-475.
L o n g u e t - H i g g i n s , M. S. 2000 Vertical jets from standing waves, (submitted).
Z a k h a r o v , V. E. 1968 Stability o f periodic waves on the surface of a deep fluid. J. Appl. Mech.
Tech. Phys. 2, 190-194.
2220
ТНЕ ROYAL
Ш S O C IE T Y
W hen a shallow-water wave meets a steep cliff or harbour wall, a jet is sometimes
throw n upwards w ith very high initial acceleration. In the present work it is shown
th a t a similar phenomenon can occur in a standing wave in deep water, provided th at
the initial shape of the trough is narrow and an almost circular arc. During collapse of
the trough, vertical accelerations exceeding 10 g are found, with subsequent vertical
velocities exceeding 1.7c, where с is the phase-speed of a small-amplitude progressive
wave. The vertical acceleration, plotted as a function of the time, is shown to be very
sensitive to the initial shape of the wave trough. The presence of a much shorter wave
superposed on the initial wave profile can further increase the maximum acceleration
and velocity.
K eyw o rd s: w a ter w aves; su rface w aves; je t form ation ; sta n d in g w aves o n w ater;
w ave breaking; a cceleration s
1. Introduction
Among the more spectacular phenomena displayed by ocean waves is the vertical
jet sometimes thrown upwards when a steep wave encounters a steep cliff face or
harbour wall. For laboratory experiments (see Chan & Melville 1988). Although
commonly observed around coastlines, the first theoretical representation of such
jets was not attem pted until Cooker & Peregrine (1990, 1991) used a boundary-
integral technique to follow the development of a shallow-water wave impacting on
a vertical wall. Under certain conditions, they found vertical acceleration of the
free surface at the wall exceeding 250# (where g is the acceleration due to gravity),
and correspondingly high pressures just below the water surface. The phenomenon,
term ed a ‘flip-through 5 by Cooker & Peregrine, was analysed by Longuet-Higgins &
Oguz (1997), who showed th at it was typical of a class of jets occurring in near-
critical situations where two surfaces come together and nearly, but not quite, trap a
pocket of ‘air’. Other examples of such behaviour are known to occur in a x i s y m m e t r i c
standing waves when a jet arises out of a wave trough (Longuet-Higgins 1983) and in
collapsing axisymmetric bubbles and cavities (Longuet-Higgins &; Oguz 1995, 1997).
It is natural then to ask what initial conditions, in the case of two-dimensional sur
face waves, will lead to a high vertical acceleration or ‘flip-through’. In the absence
of any complete analytical solution we must resort to numerical time-stepping or for
ward integration. The first problem is how best to characterize the initial conditions.
For the sake of simplicity, it is obviously preferable to find a description involving as
few free parameters as possible, and this will be our aim.
496 м . S. Longuet-Higgins
join smoothly at the points A and A ' at a distance x \ from the line of symmetry
(x = 0). At the joins, the surface slope is continuous and the slope angle is ± a . Thus
the equation of the inner curve can be w ritten parametrically,
x — pi sin в ,
when —a < в < a , (2.1)
У = 2/1 - p ic o s f l
relative to the level of the points A and A! . Here, y \ is the у-coordinate of the
centre C \ of the inner arc. The equations of the two outer arcs are similar. W ith the
wavelength being taken as 27t, it is easily shown that, in this coordinate system, the
mean surface level is
Since later we shall assume the mean water level to be zero, the constant у has to
be subtracted from the expression for у in (2 . 1 ). Thus, altogether, we can write for
the surface profile ys {x),
Vs(x)
_ f(v i-y )~ (Pi - X2 ) 1/ 2 , |i| <X i, (2.3)
\ ( У 2 - y ) + [p\ - (5Г - Z l ) 2] 1 / 2 , X l < | x | < 5Г,
where у i is the vertical coordinate of the centre of curvature C 2 relative to the level
AA '.
3. Equations of motion
The equations th at we shall use are essentially the ‘natural’ system derived by Balk
(1996) and developed by Longuet-Higgins (2000). It will be convenient to restate
them in a shortened form.
498 M. S. Longuet-Higgins
x = f + ^ 2 r n sm n 4 ,
n=l
oo (3.1)
V = Y0 + £ > n cosn*.
71=1
where У0, *i, • • • are functions of the time t (see Balk 1996). The coefficients
an = 2nYn(t), n = l,2 ,..., (3.2)
and their rates of change
dn = ^ 5 ti = 1 , 2 , . . . , (3.3)
are then a natural system for defining the motion at any given instant. If the mean
level is assumed to be constant at у = 0, the conservation of mass requires th at Yq
be a function of the time t given by
It is shown in Longuet-Higgins (2000) that Balk’s equations of motion for the Yn are
equivalent to
4. Initial conditions
We first seek a simple approximation to the initial profile (2.3), th at is, one given
by a series of the form (3.1) with as few non-zero coefficients Yn as possible. It turns
out th at this may be done remarkably well by the use of only two such coefficients,
Yi and Y2.
Figure 2. (a) A piecewise-circular profile (dotted curve) corresponding to X i/n = 0.20, a = 52.9°
in (2.3), compared with a truncated series expression (3.1) with a i = 2Y\ — 0.7814 and
a2 = 4У2 = 0.2764, shown by the solid curve. (b) Central section enlarged.
Let us initially fix X\, the half-width of the wave trough. Then by choosing any
particular value of a , the form of the profile is determined; in particular, the coordi
nates ут and y c of the wave trough and the wave crest. Then in (3.1) we can chose
the coefficients Vi and Y2 so as to make у — ут when £ = 0 and у — y c when £ = тт.
Then, to determine a closest-fitting profile, we may adjust the angle a , keeping x i
constant, so as to minimize the rectified difference
#= [ \ y - y B\d x . C4-1)
J-x i
For example, when x i / n = 0.25 (so the wave trough is one-fifth of the total wave
length), we obtain the profile shown by the full line in figure 2a. This is compared
with the piecewise-circular profile shown by the dotted line.
Figure 26 is an enlargement of the central segment of figure 2a, showing th a t the
agreement between the original piecewise-circular profile and the truncated-series
expression is indeed close. This may be attributed in part to the fact th a t the first
term in the series (3 . 1 ) by itself represents a profile that, like the solid curve, has a
sharper curvature in the trough than at the crest of the wave.
500 м . S. Longuet-Higgins
Xi / tt a a i = 2 Yi a 2 = 4V2
we can use the time-stepping equations (3.5) to follow the ensuing motion, and so
obtain the sequence of profiles shown in parts (a) and (6) of figure 4. The vertical
acceleration of the surface particle on the central plane x = 0 is plotted as a function
of the time in figure 4c. As will be seen, this rises from a value just less than g
at t = 0 to a maximum of nearly 2.8g near t = 8 .8 . It then falls away steeply
to negative values. The computation is carried only as far as t = 0.96, when the
free surface begins to form a sharp corner and the present method loses accuracy.
However, in the region of maximum acceleration, the surface is locally almost flat.
No loss of accuracy is evident until after the acceleration becomes negative.
In this integration, the time-step A t was taken as 0.001 and the number N of
harmonics was 100. The total energy E remained a constant to one part in 106 until
t = 0.95 and to one part in 10s thereafter.
Similar calculations were carried out for other profiles in figure 3, and the results
are shown in figures 5-7. The accelerations are compared in figure 8 . As the half-width
X i / n of the trough is decreased and the maximum surface slope a simultaneously
increases, so the maximum acceleration j/(0 ) increases and the time to maximum
acceleration falls. In figure 7, when х\/тг = 0.10 and a = 63.4°, the maximum
vertical acceleration is seen to exceed Юр.
This is to be compared with the acceleration of a particle in the trough of an
ordinary time-periodic standing wave, which does not exceed about ^ g .
The vertical velocities of particles at the central point x = 0, which may be found
by integrating the acceleration in figure 8 , are shown in figure 9. The maximum
velocities attained are of order unity, in non-dimensional units, and are somewhat
less variable than the maximum accelerations. It will be recalled th at (g / k ) 1^2 is the
phase speed of a progressive wave of the same wavelength 2тг/к as the standing wave.
(Here, к = 1.) The maximum vertical velocities in figure 11 are between 1.4 and 1. 8
times this phase speed.
Figure 3. A sequence of piecewise-circular profiles (dotted curves) similar to figure 2, with the
corresponding truncated-series curves (solid): (a) x \ / i r = 0.25; (b) x i / i t = 0.20; (c) xx/'K = 0.15;
( d ) x \ / t z = 0 .1 0 .
The initial accelerations in figure 8 are rather low, showing th at the pressure field
initially changes only slowly. However, if the normal velocities in the wave trough
were to be assumed non-zero and directed inwards, we should expect the resulting
accelerations to be consistently higher.
Over the ranges of time indicated in figure 8 , no instabilities in the surface profile,
numerical or otherwise, were observed. This may be due to the fact th at the vertical
acceleration always exceeded —g, not only at the central point x = 0 , but also at
other points along the free surface.
502 M. S. Longuet-Higgins
W mt
Figure 4. (a) Development of the initial wave profile corresponding to a\ = 0.6879, <22 = 0.2379
and di = a.2 = 0; A t = 0.001, N = 100. Time-interval between successive profiles is 0.02.
(6) Central section of (a) enlarged, (c) Vertical acceleration of the surface at x = 0 as a function
of time.
In figure 8 , the increase in acceleration from case (a) to case (d) might be thought
to be due to the increase in depth of the wave trough, which tends to increase
the pressure locally, hence the inward acceleration of the particles. However, this
kx
kx
Figure 5. (a) Development of the initial wave profile corresponding to a\ = 0.7815, a-i — 0.2764
and ai = a.2 —0; Д£ = 0.001, N — 100. Time-interval between successive profiles is 0.02.
(6) Central section of (a) enlarged, (c) Vertical acceleration of the surface at x = 0 as a function
of time.
would not be strictly true, for, if the scale of any wave is enlarged while leaving
the shape of the surface profile unaltered, the resulting particle acceleration would
remain unchanged, despite the changes in the corresponding particle velocities. The
M. S. Longuet-Higgins
kx
W mt
Figure 6. (a) Development of the initial wave profile corresponding to oi = 0.8779, a2 = 0.3112
and di = <Z2 = 0; A t — 0.001, N = 150. Time-interval between successive profiles is 0.02.
(6) Central section of (a) enlarged, (c) Vertical acceleration of the surface at x = 0 as a function
of time.
real cause of the increased acceleration must be sought in the narrowing of the circular
trough relative to the overall wavelength of the wave, together with an increase in the
maximum slope angle a. The latter is a purely geometrical effect enabling relatively
more fluid to be directed inwards towards the centre of the circular arc.
kx
Figure 7. (a) Development of the initial wave profile corresponding to ai = 0.9701, аг = 0.3270
and di = fa = 0; A t = 0.001, N = 200. Time-interval between successive profiles is 0.02.
(b) Central section of (a) enlarged, (c) Vertical acceleration of the surface at x = 0 as a function
of time.
506 M. S. Longuet-Higgins
W mt
Figure 8. Comparison of the accelerations shown in figures 4-7: (a) x i/ж = 0.25;
(6) x i /я- = 0.20; (с) xi/7r = 0.15; (d) x \/ n = 0.10.
In figure 10a axe shown successive profiles corresponding to the initial conditions
ai = 0.50, di = 0,1
02 = —0.25, a.2 = 0, > (6 . 1 )
03 = —0.18, аз = 0 , J
and N = 80. A close-up of the central part of the profile is seen in figure 106. The
parameters in (6 . 1) were chosen so that the initial profile had a rounded trough,
somewhat similar to that in Cooker & Peregrine (1990) (see also fig. 11 of Longuet-
Higgins & Oguz (1997)).
The acceleration у of a particle at the free surface on the axis of symmetry (я = n)
is shown as a function of time in figure 106. Fairly high accelerations у ~ 3 are
attained at around t = 0.47, well before the occurrence of the steepest surface slopes.
The next example shows the sensitivity of the maximum acceleration to small
changes in the initial conditions. If, instead of (6.1), we initially take
a i = 0.50, <21 = 0,
a 2 = —0.25, <22 = 0, ► (6-2)
аз = —0.25, аз = 0,
with N = 80 as before, we obtain the sequence of profiles shown in figure 11a.
The corresponding curve of acceleration versus time is shown in figure 116. This is
Figure 9. The vertical velocities y /(g k ) 1/,a corresponding to the accelerations shown in figure 8.
generally higher than in figure 106 and, instead of falling off after t = 0 .4 7 5 , it has a
second spurt, leading to a maximum of over 5g. This may be attributed to the fact
th a t the wave trough at around t = 0 .6 , instead of developing monotonically, itself
develops a smaller secondary trough superposed on the main trough. This acts as
a secondary standing wave, in the gravitational field defined by the first, and hence
produces a secondary vertical acceleration.
Indeed, it is possible to imagine a succession of smaller and smaller standing waves
superposed on each other, as in a fractal (see Longuet-Higgins 1994), leading to
indefinitely large vertical accelerations.
M. S. Longuet-Higgins
(a)
(gk)m t
Figure 10. Development of a wave with initial conditions a\ = 0.5, 0.2 = —0.25, аз = —0.18 and
dx = a,2 = аз = 0. (a) Central section enlarged. (6) Acceleration of the surface at x = 0, as a
function of the time. Time-interval between successive profiles is 0.02.
The process has much in common with that of the shaped charge, where it has
been called the ‘Munro effect’ or the ‘Neumann effect’ (see Birkhoff et al. 1948). Note
that the mathematical model offered by Birkhoff et al. is a steady two-dimensional
flow resulting from the convergence of two sheets of fluid onto a central plane. The
situation considered in this paper is different, in that the whole body of fluid below
the surface participates in the motion. Moreover, here we give an account of the
complete history of the time-dependent flow, starting from a state of rest.
The calculations have been carried as far as the times when the wave profile
approaches a sharp angle; then the present method, using Fourier analysis, loses
accuracy. The further development of the jet, which displays several interesting types
R e fe ren c es
Balk, A. M. 1996 A Lagrangian for water waves. Phys. Fluids 8, 416-420.
Birkhoff, G., MacDougall, D. P., Pugh, E. M. & Taylor, G. 1948 Explosives with lined cavities.
J. Appl. Phys. 19, 563-582.
Chan, E. S. & Melville, W. K. 1988 Deep-water plunging wave pressures on a vertical wall. Proc.
R. Soc. Lond. A 417, 95-131.
Cooker, M. J. & Peregrine, D. H. 1990 Water wave impact: computations, theory and a compar
ison with measurements. In Proc. 6th Int. Workshop on Water Waves and Floating Bodies,
Woods Hole, MA, USA, pp. 49-53.
Cooker, M. J. & Peregrine, D. H. 1991 Violent surface motion as near-breaking waves meet a
wall. In Breaking waves (ed. M. L. Banner & R. H. J. Grimshaw), Proc. IUTAM. Symp.,
Sydney, Australia, pp. 291-287. Springer.
Longuet-Higgins, M. S. 1983 Bubbles, breaking waves and hyperbolic jets at a free surface. J.
Fluid Mech. 127, 103-121.
Longuet-Higgins, M. S. 1994 A fractal approach to breaking waves. J. Phys. Oceanogr. 24,
1834-1838.
Longuet-Higgins, M. S. 2000 Theory of water waves derived from a Lagrangian. I. Standing
waves. J. Fluid Mech. 423, 275-291.
Longuet-Higgins, M. S. & Oguz, H. N. 1995 Critical microjets in collapsing cavities. J. Fluid
Mech. 290, 183-201.
Longuet-Higgins, M. S. & Oguz, H. N. 1997 Critical jets in surface waves and collapsing cavities.
Phil. Trans. R Soc. Lond. A 355, 625-639.
Tuck, E. O. 1998 Solution of nonlinear free-surface problems by boundary and desingularised
integral equation techniques. In Computational Techniques and Applications Conf. (ed.
J. Noye, M. D. Teubner & A. W. Gill), pp. 11-26. World Scientific.
t In recent work, vertical accelerations as high as 100$ in deep w ater have been obtained.
2236
Phys. Fluids. Vol. 13, No. 6, June 2001 On tfie breaking of standing waves 1653
FIG. 1. (a) Profiles of the free surface when C=0_5 and /=0(0.05) 1.65. (b) Enlargement of the central segment of (a), (c) Profiles of the free surface when
С = 0.5 and i - 1.65(0.05)3.25. (d) Enlargement of the ccntral segment of (c).
developed by Longuet-Higgins.15 This is useful and accurate where Y\ ,Y 2,...,Y n are functions of the time t only. Y0,
as long as the initial velocity С does not exceed about 0.5. which determines the mean level, is given in terms of
The results obtained by this method are described in Sec. Ш. .
у ...... У* by
The second method of calculation is the boundary- N
integral technique described in Longuet-Higgins and К0= - Е 2лК*. (22)
Dommcrmuth,16 which is a refinement of the technique in
troduced by Longuet-Higgins and CokeleL5 The results are
described in Sec. V. In Sec. VI we make comparisons with It is convenient to write
periodic waves. Conclusions and a discussion follow in an = 2nYn, (2.3)
Sec. VO.
and then the equations for time-stepping the coefficients
a„(r) have the form
II. METHOD A
1654 Phys. Fluids, Vol. 13. No. 6, June 2001 M. S. Longuet-Higgins and D. G. Dommermuth
FIG. 2. Plot of the maximum surface slope angle о against the curvature к FIG. 3. Plot of the vertical acceleration у of a panicle at the crest, as a
at the wave crest jc = 0, when С =0.5. function of i, in the case С =0.5.
Phys. Fluids. Vol. 13, No. 6, June 2001 On the breaking of standing waves 1655
Oc
FIG. 4. Graphs of the maximum surfacc slope angle a vs the curvaiure к at FIG. 6. Profiles of the free surfacc when С =0.7 and f “ 0(0.5)4.0.
the wave crest, when С =0.3, 0.4, and 0.5.
FIG. 5. Profiles of the free surface when C = 0.6 and r=0(0.5)4.0. FIG. 7. Profiles of the free surface when 0 0 . 8 and l = 0(0.5)4.0.
1656 Phys. Fluids. Vol. 13, No. 6. June 2001 M. S. Longuet-Higgins and D. G. Dommermuth
FIG. 8. Graphs of the maximum slope angle or vs the curvature к at the FIG. 10. Graph of the maximum surface slope or vs the crest curvature *
wave crest * = 0 when С =0.55. 0.6, 0.7, and 0.8.
when C - 1.0 and 2.0.
ture continues to increase from ( = 0 onward, well beyond smoothly and without any accompanying singularity in the
the times of maximum crest elevation or maximum surface time when a = 4 5 °.
slope. The situation is clearer still from Fig. 8, where the Profiles for some higher values of С are shown in Figs. 9
maximum slope a has been plotted against the crest curva and 11. From Fig. 10, it will be seen that when С = 1.0 the
ture к for similar values of C. The times of maximum cur maximum slope a reaches 83° and then the maximum cur
vature are, respectively, t = 1.85, 1.95, 2.4, and 3.2, well be vature к is nearly 400, so that the crest is essentially a cusp.
yond the times of maximum elevation, which never exceed The subsequent fate is also interesting, for from Fig. 9(b) it
2 .0. can be seen that the falling jet produces a cavity in the wave
From Figs. 7 and 8 it is also clear that whenever С trough. The effect is a time reversal o f the “ cavity effect”
> 0 .7 the maximum surface slope increases beyond 45° described by Longuet-Higgins.13
(a) oc (b) X
FIG. 9. (a) Profiles of the free surface when C = 1.0 and i = I ). (b) Profiles of the free surface when C * 1.0 and f = 2.0(0.1)5.0.
2241
Phys. Fluids, Vol. 13, No. 6. June 2001 On the breaking of standing waves 1657
2 0 2 4
(a)
FIG. 11. (a) Profiles of ibe free surfacc when C = 2.0 and f = 0(1.0)5.0. (b)-(e) Profile of the free *uifa< : when C=2 end 1= 5.5, 6.0,6.5. and 6.65. Number
of grid points: 751.
Figure 12, in which the crest elevation y, velocity y, and The case С - 2 . 0 is somewhat similar. From Fig. 10 the
acceleration у are plotted against time, shows that when maximum surface slope a is above 88°, and the maximum
1 .6 < f< 2 .6 the ratio jl/g was nearly - 1 , so that the particles curvature к is over 900. From Fig. 13, the fluid near the crest
near the crest were effectively in freefall. was effectively in free fall when 1 < /< 6 .0 . This time, how
2242
1658 Phys. Fluids, Vol. 13, No. 6, June 2001 M. S. Longuet-Higgins and D. G- Dommenmuth
VII. CONCLUSIONS
We have investigated the response of a body o f deep
water bounded by an initially flat horizonal surface when that
surface is given an initial normal velocity v of the form
v = Ccoskx (g = k = \ ) . (7.1)
When C <0.55 the surface oscillates approximately as a
periodic standing wave, but as С is increased there is an
increasingly pronounced asymmetry between the upward and
t subsequent downward motions. In the upward motion the
FIG. 13. Graphs of the crest elevation y, vertical velocity >, and accelera profile o f the crest is more rounded, and on the downward
tion у as functions of the time t, when С =2.0. motion the crest is narrower and more sharply curved. When
2243
Phys. Fluids, Vol. 13. No. 6. June 2001 On the breaking of standing waves 1659
C > 0 .8 the falling crest develops into a sharp, upward point ety in November 1999, and at the IUTAM Symposium on
ing spike that falls back into the wave trough. When С Free-Surface Flows, Birmingham, UK, in July 2000.
> 2 .0 , the falling crest may trap a bubble of air in the trough. M.S.L.H. is supported by the Office of Naval Research under
Since internal friction is ignored, we can, of course, Contract No. N00014-94-1-0008. D.G.D. is supported partly
imagine time to be reversed. Starting with a circular cavity in under Contract No. N00014-97-C-0345.
the trough of a standing wave, a jet will rise up vertically. Its
*M. S. Longuet-Higgins. "A theory of the origin of microseisms.” Philos.
tip will become rounded before falling back again.
Trans. R. Soc. London, Ser. Л 243, I (1950).
A comparison with the forms of time-periodic standing 3W G. Penny and A. T. Pricc. “ Finite periodic stationary gravity waves in
waves of energy slightly below the maximum has high a perfect fluid. Pan 2," Philos. Trans. R. Soc. London, Ser. A 244, 254
lighted the existence, at the same total energy £, of two (1952).
JG. L Taylor, “ An experimental study of standing wives," Proc. R. Soc.
distinct kinds of standing-wave profiles for periodic waves London. Ser. A 218. 44 (1953).
also: profiles with rounded crests and those with relatively 4P. G. Saffman and H. C. Yuen, "A note on numerical computations of
sharp crests. Urge amplitude standing waves." ). Fluid Mech. 95. 707 (1979).
In the above investigation, both the viscosity and the 5M. S. Longuet-Higgins and E. D. Cokelet, “ The deformation of steep
surfacc waves on water. L A numerical method of computation." Proc R.
surface tension T o f the water have been neglected, so that Soc. London. Ser. A 350. I (1976).
the conclusions are really limited to standing waves on a eH. Aold, "Higher order calculation of finite periodic standing gravity
scale much larger than ( T /p g ) in. For shorter waves, the waves by means of the computer,” J. Phys. Soc. Jpn. 49, 1598 (1980).
occurrence of local regions o f sharp curvature would un 7L. W. Schwartz and A. K. Witney. “ A semi-analytic solution for nonlinear
standing waves in deep water." J. Fluid Mech. 107, 147 (1981).
doubtedly produce a wealth o f parasitic capillaries. At the *G. N. Mercer and A. J. Roberts, "Standing waves in deep water Their
tips of the jets, we expect to find the formation of droplets. stability and extreme form," Phys. Fluids A 4. 259 (1992).
Note added in proof. Since this paper was submitted, the *Y. Agnon and M. Glozman, “ Periodic solutions for a complex Hamil
tonian system: New standing water waves." Wave Motion 24. 139 (1996).
authors’ attention has been drawn to two different laboratory
10M. J. Cooker and D. H. Peregrine, "Water wave impact: Compulations,
experiments in which the trapping o f air by a falling jet has theory and a comparison with measurements." in Proceedings o f the 6th
apparently been seen. Chan and Melville18 generated a fo International Workshop on Water Waves and Floating Bodies. Woods
cused train of progressive waves which then impinged on a Hole, MA. 1990. pp. 49-53.
n M. J. Cooker and D. H. Peregrine, "Violent surface motion as near-
vertical wall. They observed both the formation of a jet brcalring waves meet a wall." in Breaking Waves. Proceedings of the
against the wall and the entrainment of air when the jet fell IUTAM Symposium, Sydney. Australia, edited by M. L Banner and R. H.
back into the wave trough. The jet may have been influenced J. Grimshaw (Springer-Vcrlag, Berlin, 1991), pp. 287-291.
by frictional forces at the wall. Second, Jiang et a i 19 gener HM. S. Longuet-Higgins and H. N. Oguz, "Critical jets in surfacc waves
and collapsing cavities," Philos. Trans. R. Soc. London, Ser. A 355, 625
ated standing waves subharmonically in a tank 60 cm long (1997).
(one wavelength). The waves sometimes produced a vertical 1JM. S. Longuet-Higgins, "Vertical jets from standing waves in deep wa
jet that then fell back into the wave trough. By this procedure ter," Proc. R. Soc. London, Ser. A 457, 495 (2001).
MA. M. Balk. “ A Lagrangian for water waves," Phys. Fluids 8,416 (1996).
they avoided frictional effects at the endwalls o f the tank. Air
15M. S. Longuet-Higgins, “ Theory of water wnves derived from a Lagrang
was certainly entrained, as was evidenced by the emission of ian. L Standing waves," J. Fluid Mech. 423. 275 (2000).
sound from the entrained bubbles. But a lack of stabib'ty in I6M. S. Longuet-Higgins and D. G. Dommermuth, "Crest instabilities of
the falling jet, leading (on such a small scale) to its breakup gravity waves. Pan 3. Nonlinear development and breaking," J. Fluid
Mech. 336. 33 (1997).
into droplets, will have affected the phenomenon. 17A. Prospcreui and H. N. Oguz, "Air entrainment upon liquid impact,”
Philos. Trans. R. Soc. London. Ser. A 355, 491 (1997).
ACKNOWLEDGMENTS UE. S. rhan and W. K. Melville, “ Deep-water plunging wave pressures oo
a vertical plane wall," Proc. R Soc. London, Ser. A 417, 95 (1988).
A part of the calculations described here was presented »L. Jiang, M. Pcrlin, and W. W. Schultz, “ Period tripling and energy dis
at the New Orleans meeting of the American Physical Soci- sipation of breaking standing waves,” J. Fluid Mech. 369, 273 (1998).
2244
Michael S. Longuet-Higgins
Institute for Nonlinear Science, University of California, San Diego,
9500 Gilman Drive, La Jolla, California 92093-0402
[email protected]
A b stra c t It is shown that a sufficient condition for the occurrence of a high vertical
acceleration in a two-dimensional standing wave is for the initial profile
of to have a narrow trough that is approximately circular. In such cases
■we find vertical accelerations of the free surface exceeding lOg. Much
higher accelerations are obtained when the initial wave trough is some
what flattened so that jets emerging from the two corners of the trough
meet on the axis of symmetry. In such a case we find accelerations
exceeding lOOg.
Lastly we consider a different, but complementary, class of initial
conditions when the free surface is initially flat but is given an upwards
velocity dependent sinusoidally on the horizontal coordinate. When the
kinetic energy injected into the fluid exceeds the maximum possible for
a time-periodic standing wave it is found that a new type of breaking
can occur in which air is trapped by the collapse of the wave trough.
Throughout this paper surface tension is ignored and the fluid flow
is treated as inviscid and irrotational.
1. INTRODUCTION
Vertical jets thrown upwards when a steep progressive wave meets a
cliff or harbor wall are a common phenomenon around coastlines and in
shallow water. A numerical model for shallow water, using a boundary-
integral technique, was constructed by Cooker and Peregrine (1991), who
found that in some cases the vertical acceleration of a particle at the wall
could exceed 250g. Longuet-Higgins and Oguz (1997) showed that such
high accelerations were typical of a class of jets occurring in near-critical
situations where two surfaces came together and nearly, but not quite,
trapped a pocket of “air.” Other examples of high vertical accelerations
195
A.C. King and YD. Shikhmurzaev (eds.), UJTAM Symposium on Free Surface Flows, 195-203.
© 2001 Kluwer Academic Publishers. Printed in the Netherlands.
2245
196
197
Figure 2. A sequence of analytic profiles of the form (2.1) (solid curves) approx
imating piecewise circular profiles of the type shown in Fig. 1 (dotted curves), (a)
p - 0.25, (b) p = 0.20, (c) p = 0.15, (d) p = 0.10.
198
'h
Figure 3. (a) Successive wave profiles corresponding to the initial profile (d) of Fig.
2. Number of harmonics N = 100. Time-stepping interval dt — 0.001. Interval At
between successive profiles: Дt = 0.02 (b) central section of (a) enlarged.
a)
l
We can then use the system of O.D.E.’s for the Yn first given by Balk
(1996) and developed by Longuet-Higgins (2000). These are much sim
pler than the corresponding Hamiltonian equations.
As a result we obtain, for example in the case 2x \/L = 0.25, the
sequence of profiles shown in Figures 3a and 3b. The vertical acceleration
у of a point particle on the mid-line cr = 0 is plotted as a function of the
time in Figure 4, curve d. Evidently у rises to a maximum of about lOg
at about t = 0.41, before falling abruptly to below zero as the curvature
of the wave crest becomes increasingly sharp.
Figure 4 shows the sequence of such acceleration curves, corresponding
to the four initial profiles in Figure 2. It is notable that as the width of
the wave trough 2xi diminishes in relation to the wavelength L, so the
maximum acceleration increases; see Longuet-Higgins (2001a).
When, however, the crest curvature becomes large, the number of sig
nificant harmonics Yn in the series (2.1) increases rapidly, and the above
method is no longer useful. From this time on, or alternatively from the
initial instant t = 0, we can use a standard time-stepping program using
2248
199
200
1
(fr>
|
/
I
!
1
$
1. . Л-1
201
Кос
202
Figure 8. As in Fig. 7, but with С = 2.0. (a) t = 0 to 5, (b) t = 5.5, (c) t = 6.0,
(d) t = 6.65.
2252
203
Acknowledgement
The above work has been supported by the U.S. Office of Naval Re
search under Contract N00014-94-1-0008.
References
Balk, A.M. 1996 A Lagrangian for water waves. Phys. Fluids 8 , 416-420.
Birkhoff, G., MacDougall, D.P., Pugh, E.M. and Taylor, Sir G. 1948 Explosives with
lined cavities. J. Appl. Phys. 19, 563-582.
Chan, E.S. and Melville, W.K. 1988 Deep-water plunging wave pressures on a vertical
plane wall. Proc. R. Soc. Lond. A 417, 95-131.
Cooker, M.J., and Peregrine, D.H. 1991 Violent surface motion as near-breaking
waves meet a wall. pp. 291-297 in Breaking Waves eds., M.L. Banner and R.H.J.
Grimshaw, Berlin, Springer 387 pp.
Jiang, L., Perlin, М., and Schultz, W.W. 1998 Period tripling and energy dissipation
of breaking standing waves. J. Fluid Mech. 369, 273-299.
Longuet-Higgins, M.S., and Oguz, H.N. 1997 Critical jets in surface waves and col
lapsing cavities. Phil. Trans. R. Soc. Lond. A 355, 625-639.
Longuet-Higgins, M.S. and Dommermuth, D.G. 1996 Crest instabilities of gravity
waves. Part 3. Nonlinear development and breaking. J. Fluid Mech. 336, 33-50.
Longuet-Higgins, M.S. 2000 Theory of water waves derived from a Lagrangian. I.
Standing waves. J. Fluid Mech. 423, 275-291.
Longuet-Higgins, M.S. 2001 Vertical jets from standing waves in deep water. Proc. R-
Soc. A 457, 495-510.
Longuet-Higgins, M.S. and Dommermuth, D.G. 2001 a Vertical jets from standing
waves. II (submitted).
Longuet-Higgins, M.S. and Dommermuth, D.G. 2001b On the breaking of standing
waves (submitted).
Mercer, G.N. and Roberts, A.J. 1992 Standing waves in deep water: Their stability
and extreme form. Phys. Fluids A 4 , 259-269.
Prosperetti, A., and Oguz, H.N. 1997 Air entrainment upon liquid impact, Phil. TYans.
R. Soc. Lond. A 355, 491-506.
Zeff, B.W., Kleber, B., Fineberg, J., and Lathrop, D.P. 2000 Singularity dynamics in
curvature collapse and jet eruption on a fluid surface. Nature 403, 401-404.
2253
ШЩ THE ROYAL
10.1098/rspa.2001.0832 Ш 1 SOCIETY
In a previous paper it was shown that in steep standing waves on water the collapse
of a rounded cavity in the wave trough can produce quite high local vertical acceler
ations, initiating the growth of a strong vertical jet. In one example given here, the
acceleration exceeds lOOp. The main purpose of the present paper is to follow the
subsequent development of the jet by means of a boundary-integral time-stepping
technique. It is found that the jets tend to bifurcate into two halves, each forming
a plunging or spilling breaker. The transition of the rising wave trough into a thin
jet is here compared with an asymptotic flow in which the free surface is given by a
quartic equation.
K e y w o rd s: w ater w aves; su rfa ce w aves; je t form ation ; wave break in g;
sta n d in g w aves; v e rtical acceleration
1. Introduction
In a recent paper (Longuet-Higgins 2001a, hereafter referred to as paper I) it was
shown th at steep standing waves on the surface of deep water can, in some circum
stances, produce remarkably large vertical accelerations, giving rise to vertical jets
of water arising out of the wave troughs. A similar phenomenon could occur through
the impact of a water wave against a vertical wall or cliff face or against the side of
a ship in deep water. A sufficient condition for moderately large accelerations is that
the wave trough have the form of an almost circular arc, as in a shaped charge.
The previous calculations were carried out using a new method due essentially to
Balk (1996), which is particularly convenient for parametrizing the initial conditions
for the flow. The examples given in paper I showed th at the resulting vertical accel
erations of a particle in the wave trough could exceed 10 <7, before falling off rapidly
to smaller values.
After the acceleration had fallen to near zero, the tip of the jet began to develop
a sharp curvature. For the mathematical description of such a sharp corner, Balk’s
analysis, which is in terms of a Fourier series, is not so well adapted. In fact the
calculations tended to lose accuracy slightly before the sharp corner was formed.
The purpose of the present paper is to continue and extend the previous results by
making use of another but complementary method of numerical time-stepping, which
was devised for the study of breaking waves by Longuet-Higgins &; Cokelet (1976).
This boundary-integral technique has been further developed by many authors: see,
Proc. R. Soc. Lond. A (2001) 457, 2137-2149 © 2001 The Royal Society
2137
2254
for example, Vinje & Brevig (1981), Dold & Peregrine (1986), Roberts (1983) and
Mercer & Roberts (1992). In the present paper we shall make use of an accurate
version of Vinje & Brevig’s formulation, which has been used previously to follow
the nonlinear development of steep gravity waves when subjected to a normal-mode
perturbation (Longuet-Higgins & Dommermuth 1997). It can equally well be used
to study the time-history of any space-periodic surface waves, and in particular to
follow the standing-wave jets th a t concern us here.
2. Initial conditions
In paper I the horizontal and vertical coordinates (ж, у) of a point on the free surface
were given in general by equations of the form
N
x = £+ n - 1a„ sin n f ,
n=1
(2 .1)
N
У = Уо + n~l an cos n f ,
n= 1
where £ is a param eter running from 0 to 27r over one wavelength L = 27Г, and the
coefficients an(t) are functions of the time t only. Moreover, to conserve the total
mass of fluid,
» - - £ s “2 <22)
n=l
In the present system, the initial conditions are specified by the values of the
coordinates (x ,y ) of each point on the surface at time t = 0 , together with the
velocity potential at the free surface. In all of the examples considered here, the
initial velocity is zero. Hence we may simply take
d „(0 )= 0 , n = l, 2 (2.3)
The particular examples chosen in paper I had very simple initial conditions. In the
first four examples А, В, С and D we took N = 2 . The initial values of the coefficients
a\ and a 2 were given in table 1 of paper I, and will be restated in the figure captions
below. They correspond to cases where the total width 2x\ of the wave trough is
given by
x i /тг = 0.25, 0.20, 0.15 and 0.10, (2.4)
respectively. The time t is measured in units in which the wavenumber к and the
acceleration due to gravity g are both equal to 1 .
3. Results
Figure 1 shows the development of the surface profile in case A , x i/n = 0.25, from
t = 0-2.4 at intervals of 0.2. A little more than half a wavelength is shown. W hen
t = 0, the wave trough between x /L = 0.375 and 0.625 is very nearly circular. The
‘crest’ arising out of the wave trough first develops a sharp corner at around t = 1.0
(the limit of the previous calculations). At around t — 1.2 a small protrusion arises
0.3 r
—0 з Г-i. i , l i I i i i i I i i i i I i i i i I i i » i I i i i i
-0.5 -0 .4 -0.3 -0.2 -0.1 0 0.1
xlL
Figure 1. Profiles of the free surface at times t — 0 (0.2) 2.4 in case A,
x \ / L = 0.25. Initial values: ai = 0.6879, аг = 0.2379 and di = аг = 0.
out of the crest and then becomes sharper (t = 1.4 and 1.6), while the main body
of the crest thickens and continues to rise. By the time t = 2.4 it appears that the
sharp tip of the jet is being reabsorbed by the main body.
In figure 2 it can be seen that after t = 2.4 the tip is completely reabsorbed, while
on each side the surface overturns as in a plunging breaker. The calculation could
0.3
0.2
0.1
- 0.
-о.:
-о.:
-0 .5 -0 .4 -0 .3 -0 .2 -0.1 0 0.1
xJL
Figure 3. Profiles of the free surface at times t = 0 (0 .2) 2.8 in case B,
X i/L = 0.20. Initial values a i = 0.7815, a2 = 0.2764 and di = d2 = 0.
not be continued much beyond t = 4.0, owing to the sharp curvature at the tips of
the two breakers.
The corresponding acceleration у of the particle at the central point (y /L = 0.5)
is shown in figure 12, curve A. The acceleration у reaches a peak of about 2 .8*7
at around t = 0.75, when the free surface is still nearly flat locally, and then falls
abruptly to —g, when the sharp-pointed tip is formed. In other words, th a t part
of the fluid is in free-fall. The acceleration remains at —g until t = 2.5, when the
sharp tip has been reabsorbed into the main jet. The acceleration remains negative,
however, until after t = 4 .0 .
Next consider case B, when x \fL — 2.0; the initial width of the trough is only
one-fifth of the wavelength. Figure 3 shows the jet developing more rapidly than in
case A, but otherwise in a similar way. The sharp tip that is formed between t = 0.8
and 1.0 rises higher, and when it falls back into the main body it produces a small
circular cavity, shown enlarged in figure 4. This reabsorption of the mini-jet can be
viewed as the same process as the ‘shaped charge’ effect, but in reverse.
The vertical acceleration for case В is shown in figure 12, curve B. This tim e у
rises to almost bg at around t = 0.71, before falling to —g and remaining there till
after t = 2 .8 ; throughout this time the tip of the jet is in free-fall.
The very small scale of the features near the cavity in figure 4 prevents the com
putation in this case being carried beyond about t = 3.1. However, on the reasonable
assumption th at this small part of the flow does not significantly affect the devel
opment of the rest of the profile we may cut out a small central section of total
width 0 .01 , say, and splice both у and ф across the gap, using a simple interpola
tion formula, before continuing the time-stepping. This procedure yields figure 5.
An enlarged view (figure 6 ) shows the detail of one of the overturning jets for two
different grid solutions.
xIL
Figure 4. Enlargement of the top of the jet in figure 3 at time t = 2.8 (0.1) 3.1.
In case C, when x i/L = 0.15, the corresponding profiles are shown in figures 7-10.
From figure 7 we see th at the tip of the jet is reabsorbed at around t = 2.0. The
maximum vertical acceleration, from figure 12 , curve C, is now greater than 8 g.
Again, to carry the computation beyond t = 2 it was necessary to cut and splice a
narrow section of the profile including the small cavity, as in figure 8 . The resulting
xJL
Figure 6. Enlargement of the overturning jet in figure 5 (case C) at t — 3.8
(-------, 256 p oints;------ , 512 points).
profiles are shown in figures 9 (centre rising) and figure 10 (centre falling). The side
jets clearly become very sharp pointed and thin.
Case D, when X \ / L = 0.10, shows a somewhat different behaviour. Figure 11
suggests th at the sharp tip does not have time to develop before it is overtaken by
the main body of the jet. The absence of this feature enables us to continue the
xIL
Figure 8. Splicing the central section of figure 7 at t = 2.0 (------, original;-------, splice).
calculation until the plunging breakers on each side develop very sharp corners. This
time the vertical acceleration у rises to over Юр.
Figure 12 combines and compares the acceleration curves for all the cases A-D
considered so far. It is essentially an extension of fig. 8 in paper I to negative values
of у (and larger values of t) and confirms very satisfactorily the agreement between
the two quite different methods of numerical calculation.
xIL
Figure 11. Profiles of the free surface at times t = 0 (0.5) 3.0 in case D, x i / L = 0.10.
Initial values: ai = 0.9701, a -2 = 0.3270 and di = d2 = 0.
4. Transition to a jet
One of the most interesting features of the profiles in figures 1 , 3, 7 and 11 is the
smooth transition of the rising trough from a sharp corner, with slope angle less than
45°, to a more narrow, sharp-tipped jet. In figure 7, for example, this occurs between
Figure 12. The vertical acceleration as a function of the time in cases A-D.
x
Figure 13. Intermediate profiles in case С at intervals of A t = 0.02,
showing the transition from a sharp corner to a jet.
x
Figure 14. Profiles given by the asymptotic expression (4.1) when 7 = —6 ,
at intervals of A t = 1.0.
x/L
Figure 15. Profiles of the free surface at times t = 0 (0.1) 0.5 with initial
conditions ai = | , 02 = — аз = —| and di = <22 = аз = 0.
plot is interesting. In figure 18 we have plotted \n(y/g) against In |£c —1\. It can be
seen that the plot is linear over l | decades, showing that
y/ g oc |<c - t f (5.2)
over this range of t. This type of power-law behaviour has been shown to be typical
of free-surface flows that are close to being critical, in the sense that they almost
pinch off a ‘bubble of air’ (see Longuet-Higgins & Oguz 1997).
t
Figure 17. The vertical acceleration of a particle on the central line in
figures 15 and 16, as a function of t.
t
Figure 18. Logarithmic plot of y / g against 11 — t c \, where t c = 0 .54.
6. Conclusion
When a standing gravity wave is allowed to develop from a rounded cavity, the
vertical acceleration may attain quite high values compared with g. However, in the
su sequent development of the jet the accelerations are relatively small and the tip
о the jet is practically in free-fall. In the examples given here the jet bifurcates into
two plunging or spilling breakers falling away to either side. As shown elsewhere
(Longuet-Higgins & Dommermuth 2001), other initial conditions can give rise to
single, sharp-pointed jets, which relapse into the surface and form a temporary cavity:
a time-reversal of the cases discussed here. Both types of behaviour seem to have
been observed by Jiang et al. (1998) in their experiments on subharmonically forced
standing waves. However, a full dynamical explanation of these authors’ observations
has yet to be given.
The calculations in §§3-5 were presented at the IUTAM Symposium on Free-Surface Flows,
Birmingham, UK, in July 2000. M.S.L-H. is supported by the Office of Naval Research under
contract N00014-00-1-0248. D.G.D. is supported under contract N00014-97-C-0345.
References
Balk, A. M. 1996 A Lagrangian for water waves. Phys. Fluids 8 , 416-420.
Dold, J. W. & Peregrine, D. H. 1986 An efficient boundary-integral method for steep, unsteady
water waves. In Numerical methods for fluid dynamics (ed. K. W. Morton & M. J. Baines),
vol. II, pp. 671-679. Oxford University Press.
Jiang, L., Perlin, M. & Schultz, W. W. 1998 Period tripling and energy dissipation of breaking
standing waves. J. Fluid Mech. 369, 273-299.
Longuet-Higgins, M. S. 1972 A class of exact, time-dependent free-surface flows. J. Fluid Mech.
55, 529-543.
Longuet-Higgins, M. S. 2001a Vertical jets from standing waves. Proc. R. Soc. Lond. A 457,
495-510.
Longuet-Higgins, M. S. 20016 Asymptotic forms for jets from standing waves. J. Fluid Mech.
(In the press.)
Longuet-Higgins, M. S. & Cokelet, E. D. 1976 The deformation of steep surface waves on water.
Proc. R. Soc. Lond. A 350, 1-26.
Longuet-Higgins, M. S. & Dommermuth, D. G. 1997 Crest instabilities of gravity waves. Part 3.
Nonlinear development and breaking. J. Fluid Mech. 336, 33-50.
Longuet-Higgins, M. S. & Dommermuth, D. G. 2001 On the breaking of standing waves by
falling jets. Phys. Fluids 13, 1652-1659.
Longuet-Higgins, M. S. & Oguz, H. N. 1997 Critical jets in surface waves and collapsing cavities.
Phil. Trans. R. Soc. Lond. A 355, 625-639.
Mercer, G. N. & Roberts, A. J. 1992 Standing waves in deep water: their stability and extreme
form, Phys. Fluids A 4, 259-269.
Roberts, A. J. 1983 A stable and accurate numerical method to calculate the motion of a sharp
interface between fluids. IMA J. Appl. Math. 31, 13-35.
Vinje, T. & Brevig, P. 1981 Breaking waves on finite water depths: a numerical study. Report
R-111.81, Ship Research Institute of Norway.
J. Fluid Mech. (2001), vol. 447, pp. 287-297. © 2001 Cambridge University Press 287
DOI: 10.1017/S0022112001006061 Printed in the United Kingdom
Standing gravity waves forced beyond the maximum height for perfect periodicity
can produce vertical jets with sharp-pointed tips. In this paper, canonical forms for
the wave crests are derived which display sharp cusps in the limit as the time t tends
to infinity. The theoretical profile is in general quartic in the space coordinates, and
can describe the smooth transition of a fairly low wave crest to a cusped form. There
is no singularity as the surface slope passes through 45°.
1. Introduction
The forming of vertical jets in standing waves and in steep progressive waves
meeting a vertical wall has been studied in several recent laboratory experiments (Chan
& Melville 1988; Jiang, Perlin & Schultz 1998; Bredmose et al. 2000) and numerical
com putations (Cooker & Peregrine 1991; Longuet-Higgins 2001a,b; Longuet-Higgins
& D om m erm uth 2001). While the early, highly accelerated phase of the m otion has
received some attention and a certain degree of explanation (Longuet-Higgins &
Oguz 1997; Cooker 2000; Longuet-Higgins 2001a,b), later stages of the flow, when
the vertical acceleration at the crest has fallen almost to —g and the tip of the jet is in
free-fall, have yet to be represented analytically, even when the influences o f surface
tension and viscosity are theoretically absent.
Both the experiments and the more recent numerical calculations, especially those
by Longuet-Higgins & Dommermuth (2001), strongly suggest that the tip o f the
jet can tend towards a sharp cusp with increasing time t. This implies the presence
of cubic or higher-order terms in the expression for the velocity potential. Such an
analytic form has been suggested previously for a plunging breaker in a progressive
gravity wave (Longuet-Higgins 1980, §10). Here we shall apply the same type of
analysis to the simpler case of a symmetrical jet produced by a standing wave.
In all of the present calculation it is assumed that the flow is inviscid and irrotational
and that surface tension can be ignored. Moreover the motion will be viewed in a
free-fall frame of reference, so that there is no local effect of gravity. The solutions
are to be valid asymptotically near the tip of the jet and as the time t tends to infinity.
As far as fourth order in the radius r, the procedure yields a family o f canoni
cal solutions with one free parameter; see §4. These are illustrated by numerical
computations in § 5, and a discussion follows in § 6 .
2. General solution
Take rectangular coordinates as in figure 1, with the x-axis horizontal and the y-axis
vertically upwards; r and 9 are polar coordinates such that (x,y) = r(sin 0, —c o s 6).
2267
288 M. S. Longuet-Higgins
- P= 0 , + ^ 2- m (2.2)
where q is the particle velocity and F an arbitrary function of t. On substitution from
(2.1)
terms of higher degree in r being neglected. Similarly for the rate of change of p
following a particle we find
On the free surface, both p and Dp/T)t are to vanish simultaneously. This can be
ensured by specifying that
A = ±, F= J , (2.8)
where F0 is an arbitrary constant. Note that the first term in equation (2.1) represents
a stagnation point flow:
Ф= \ ( у * ~ Л (2.9)
which from (2.3) has a real free surface
x2 = tF = Fo/t3 (2.10)
consisting of two parallel vertical planes, but only if F0/ ( 3 > 0.
Next, on equating coefficients of r 3 cos 9 and r 3 cos 39 in equation (2.5) we obtain
2(А2В + A B + AB) + FRi 2A2В + \B + FR3 _ F „
AB \B F'
These constitute two equations for Ry and R3 in terms of A, В and F, namely
FRi = AB(F/F) - 2(A2B + A B + AB), 1
(2. 12)
FRi = \B (F /F )- (2 A 2B + \B ). J
For example, if we choose В = B0/ t 2 where BQ is a constant we find
(2.13)
and hence
«1=0, R3 = ^ . (2-14)
3Fo
Smiilarly by equating coefficients of r 4,r 4 cos20 and r 4 cos 40 in equation (2.5) we
can obtain three equations for S0, S2 and S4 which, if С = Q / t 3, yield
с _ s _ B l-C % Bq + 5C0 p t <5)
S° - 2 To’ S2~ — F T ' S* - — * F T - {
Note that as t increases the term in r 3 in equation (2.5) will come to dom inate the
series. The term in r 4 is relatively small, since by hypothesis r /t <C 1, and similarly
for terms of higher order in r.
2269
290 M. S. Longuet-Higgins
3. The free surface
The free surface p = 0 can be found by equating to zero the right-hand side of
(2.3). If we now substitute
A -- B -M
A~ t< B~ fi’ cr ~-fi
- 4y НП
(3-1)
in this expression we obtain
^ ( 5 - 5 co s 20 )
j.3
+ (3/? cos 9 — cos 30)
F0
t*K2r ' ' 7 t4
On replacing 1 by (cos2 0 + sin 2 0)” where appropriate we get
r2
sin 2 0
t2
r2
4- -у (P cos 3 0 + 9p cos 0 sin2 0)
Fo
+ 7 ЩР2 + У)cos4 0 + 9(p2 + 2y) cos2 9 sin2 9 + ( \p 2 - ly) sin4 9] = (3.3)
Now writing
(Z, П) = W t , y /t) = (r/0(sin 9, - cos 9) (3.4)
we have
In the limit as e —> 0 this curve (or surface) has an upward- or downward-pointing
cusp according as p is negative or positive 0. If e < 0 the curves p = 0 lie inside the
cusped curve and if € > 0 they lie outside. Since e — Fo/14 —►0 as t —►oo we see that
the limiting curves are in general cusped.
Note that if all the fourth-order terms in equation (3.5) are neglected we obtain the
lower-order approximation
Z2 - p ( r i 3 + 9 e r i ) = e (3.8)
2270
with solution
к2 _ С + Й 3 (3.9)
i - w
The corresponding profile will have a horizontal asymptote given by rj = (9f i r 1-
4. Canonical forms
W hen fi < 0, giving an upward-pointing cusp when e —►0, we may by a suitable
choice of velocity scale choose = —1. (Setting fi = +1 would simply reflect the
profile in the horizontal axis, i.e. turn it upside down.) Then by a further choice of
time scale we may set e = +t~4. Thus equation (3.8) is reduced to the canonical form
x2 = St~2 —(y3 -f 9x2y )/t, (4-1)
where <5 = +1. Similarly equation (3.5) is reduced to
** = s r 2 - (y3 + 9x2y ) / t + [(7y - \ )x4 - 9(2y + l ) x V - (7 + | ) / ] / Л (4.2)
Exceptionally we may have /? = 0, and then by choice of scales
x 2 = д Г 2 + <5;(7x4 - 1 8 х У - у4)/? . (4.3)
As always, these expressions are valid only when x /t and y / t are sufficiently small.
5. Numerical calculations
Consider first the approximation (4.1). The case S = —1 is shown in figure 2. For
clarity, the profiles have been separated by adding successive vertical displacements
At/9, where At is the time interval between profiles. Physically this is equivalent to
2271
292 M. S. Longuet-Higgins
Figure 3. Surface profiles corresponding to equation (4.1) when <5 = +1 and V = 1/9.
Figure 4. The inclination a to the horizontal of the asymptotes in the profile (4.2),
as a function o f у (or y/fi2).
imposing a uniform vertical velocity V = 1/9. The effect is to maintain the horizontal
asymptote at the same level у — 0 .
As t increases, the central part of the profile approaches the form of a semi-cubical
parabola, with increasing accuracy.
The case S = +1 is shown in figure 3. The most noticeable difference between
these profiles and those of figure 2 is that they can become very steep, indeed almost
vertical, at points other than the central cusp. If a vertical velocity were not imposed,
these curves would all be nested inside each other.
In order to classify the profiles corresponding to equation (4.2) it is convenient to
2272
F igure 5. (a) Surface profiles corresponding to the canonical form (4.2) when 8 = —1 and у = —5.
The time interval At between successive profiles is 2. To separate the curves a vertical velocity
V = 0.06 has been imposed, (b) Extension o f (a) but with At = 10 and V = —0.015.
consider the asymptotes for large values of x and у (but small values of x / t and
y /t). The directions of the asymptotes are found by equating to zero the fourth-order
terms, hence
294 M. S. Longuet-Higgins
F ig u r e 6. Surface profiles corresponding to the canonical form (4.2) when <5 = -f-1 and у = —5.
Time interval At = 10. Imposed velocity V = 1/9.
R gure 7. Surface profiles corresponding to the exceptional case 0 = 0, equation (4.3), when
<5 = —1,5' = —1. Time interval between profiles: At = 10. Imposed velocity V = 0.
—1.227 < у < 0.643, there are no real asymptotes. On the other hand as у —►±oo, the
angle у tends to a finite value: a = 31.68° (see below).
A typical example, when у = —5, is shown in figures 5(a) and 5(b). In figure 5(a),
in order to separate the curves, a vertical velocity V = 0.05 has been imposed. The
2274
F igure 8. Numerically calculated profiles of wave crests generated at time t = 0 by imposing the
vertical velocity (6.1) on a flat water surface. Here С — 1. Enlargement o f figure 9 o f Longuet-Higgins
& Dommermuth (2001), seen in a free-fall reference frame.
successive profiles seem to represent the transition of a rounded wave crest into an
upwards vertical jet. It is noticeable that the maximum slope of each curve passes
through 45° without any singular behaviour with respect to the time t.
The sequence is extended to larger times t in figure 5(b), and here, for the sake of
clarity we have imposed a uniform downwards velocity V = —0.01. As t —►oo we see
the development of a sharp cusp.
Similarly when 5 = 1 and у = —5 we obtain the sequence of surface profiles shown
in figure 6 . The curves are steeper, as in figure 4. The profile corresponding to t = 10
is not shown, since that joins with another branch coming from above the crest to
yield an unphysical solution.
When <5 = 1 and у = 5 the curves are again uninteresting.
Lastly we consider the exceptional case /? = 0, given by (4.3). The only physically
interesting case is when 5 = —1. The profiles are then as in figure 7. No vertical
velocity V has been imposed. For large values of x and у the curves approach the
asymptotes
7x4 - 1 8 x V - / = 0, (5.3)
that is to say
Х2/У г = (9 + V§8)/7 = 2.6258 (5.4)
or
y /x = ±0.6171 = ± arctan 31.68°. (5.5)
In each of figures 2, 3, 5, 6 and 7 the necessary condition that x /t and y / t shall be
small can be seen to be satisfied.
2275
296 M. S. Longuet-Higgins
Figure 9. Enlargement o f the crests in figures 5(a) and 5(b), but with superposed velocity
У = —0.020. These correspond to equation (4.2) with <5 = —l,y = —5. Time interval between
profiles, At = 4.
6. Discussion
Like the limiting Stokes comer flow in progressive surface waves, and its extension
to progressive waves of near-limiting height (Longuet-Higgins & Fox 1977, 1978) the
expressions given above are valid only asymptotically, in this case in the limit as
t —►oo and r /t —►0. The functions B(t) and C(r) in §§2 and 3 are to some extent
arbitrary, and must be determined, in any particular case, by fitting the inner flow
which they represent to the outer flow in the rest of the fluid. The fitting must of
course be carried out in both the space and time domains.
At first sight it appears that by adding suitable terms of order r 5 or higher powers
on the right of equation (2.5) one could proceed indefinitely to higher approximations.
However, there is no guarantee that the process will converge. It may well be more
convenient to adopt a low-order approximation which is certainly valid asymptotically
in the prescribed limits.
There is strong evidence that the asymptotic expressions for the surface profile
given in §§ 3 and 4 above do indeed describe the development of jets in some recent
numerical calculations. For example Longuet-Higgins & Dommermuth (2001) have
shown by precise calculation that if an upwards vertical velocity v of the form
v = С(gk)~l/2 cos kx (6 . 1)
is imparted at time £ = 0 to an infinitely deep body of fluid, where к is a horizontal
wavenumber, g denotes gravity and С is a constant of order 1, then the crest of the
wave so generated tends to develop a sharp upwards jet. Figure 8 is an enlargement
of the wave crest, seen in a free-fall frame of reference; a displacement \g{t —t0)2 has
been added to the computed wave profile, where t0 is taken equal to 2.5. This may be
compared with figure 9, which is an enlargement of the crests in figures 5(a) and 5(b),
but with a superposed uniform velocity V = —0.015. The profiles in figure 9 represent
2276
This research has been supported by the Office of Naval Research under C ontract
N00014-00-1-0248.
REFERENCES
М., P e r e g r i n e , D. H . & T k a is , L. Experimental investigation and
B r e d m o s e , H ., B r o c c h i n i ,
numerical modelling of steep forced water waves. Euromech Colloquium 416, Genova, Italy,
17-20 Sept. 2000.
C h a n , E . S. & M e l v i l l e , W . K . 1988 Deep-water plunging wave pressures on a vertical plane wall.
Proc. R. Soc. Lond. A 417, 95-131.
C ooker, M. J. 2001 Violently erupting free-surface jets. Proc. IUTAM Symp. on Free Surface Flows,
Birmingham, UK, 10-14 July 2000. Kluwer.
C ooker, M. J. & P e r e g r i n e , D. H. 1991 Violent surface motion as near-breaking waves meet a
wall. In Breaking Waves (ed. M. L. Banner & R. H. J. Grimshaw), pp. 291-297. Springer.
J ia n g , L., P e r l i n , M. & S c h u l t z , W. W. 1998 Period tripling and energy dissipation o f breaking
standing waves. J. Fluid Mech. 369, 273-299.
L o n g u e t - H i g g i n s , M. S. 1972 A class of exact, time-dependent, free-surface flows. J. Fluid Mech.
55, 529-543.
L onguet-H iggins, M. S. 1980 On the forming of sharp corners at a free surface. Proc. R. Soc.
Lond. A 371, 453-478.
L o n g u e t - H i g g i n s , M. S. 2001a Vertical jets from standing waves: the bazooka effect. In IUTAM
Symp. on Free Surface Flows, Birmingham, UK, 10-14 July 2000, pp. 195-204. Kluwer.
L o n g u e t - H i g g i n s , M. S. 2001b Vertical jets from standing waves in deep water. Proc. R. Soc. A
457, 495-510.
L o n g u e t - H i g g i n s , M. S. & D o m m e r m u th , D . G. 2001 On the breaking o f standing waves by
falling jets. Phys. Fluids 13, 1652-1659.
L o n g u e t - H i g g i n s , M. S. & Fox, M. J. Н. 1977 Theory o f the almost-highest wave: The inner
solution. J. Fluid Mech. 80, 721-741.
L o n g u e t - H i g g i n s , M. S. & Fox, M. J. Н. 1978 Theory of the almost-highest wave. II. Matching
and analytic extension. J. Fluid Mech. 85, 769-786.
L o n g u e t - H i g g i n s , M. S. & O g u z , Н. N. 1997 Critical jets in surface waves and collapsing cavities.
Phil. Trans. R. Soc. Lond. A 355, 625-639.
2277
J. Fluid Mech. (2002), vol. 466, pp. 305-318. © 2002 Cambridge University Press 305
DOI: 10.1017/S0022112002001246 Printed in the United Kingdom
1. Introduction
Surface gravity waves of low amplitude, when reflected from a vertical wall, will
produce standing waves of double the amplitude of the incident waves. If on the other
hand the incident waves are steep enough, they are found to throw up vertical jets
of water against the reflecting barrier; see for example Chan & Melville (1988). The
same phenomenon is often observed when incoming waves meet a cliff or harbour
wall. Moreover, this is not merely a shallow-water phenomenon but occurs also in
deep water. As will be seen below, it follows from a simple energy argument that
progressive waves of more than a certain steepness cannot produce periodic standing
waves; they must be aperiodic.
In his well-known experiments on steep standing waves Sir Geoffrey Tayor (1954)
considered only periodic waves, whose maximum slope he found to be about 45°. In
accurate computations Mercer & Roberts (1992) showed that periodic standing waves
of given wavelength cannot have more than a certain energy, although at energies
slightly below that maximum two different periodic waves having the same energy
can exist. Jiang, Perlin & Schultz (1998) have carried out experiments on deep-water
standing waves forced subharmonically by a vertical oscillation of the wave tank. It
was shown that the waves could be made to break periodically once every three wave
‘periods’. Numerical studies of ‘super-energetic’ standing waves, with various initial
conditions, have been carried out by Longuet-Higgins & Dommermuth (2001a, b)
and it was found that either single jets could be produced, which fell back vertically
into the trough of the wave, creating a semi-circular cavity, or in other cases, starting
2278
with a circular cavity, the wave crests could become flat-topped and then break on
either side of the wave crest, like a pair of spilling or plunging breakers.
The purpose of the present paper is to describe experiments in which free progressive
waves are allowed to impinge on a vertical wall, and the complete history of the motion
is then followed. There is no forcing of the wave motion by a vertical or other kind
of oscillation of the boundary, apart from the remote wavemaker. It is found that
when the steepness ak of the incident wave train exceeds about 0.236 the waves in
the neighbourhood of the boundary develop a growing instability in which every
third wave (in time) is the steepest, the two intermediate waves having flat-topped or
rounded crests, or of a mixed type. The steepest waves become sharp-crested; see § 5.
A discussion and conclusions follow in § 6 .
where p is the density, g the acceleration due to gravity and a the wave amplitude.
The reflected wave is similar, and the two waves combine to form a standing wave of
maximum amplitude 2 a (crest-to-trough height 4 a) and of time-averaged energy
Es = 2EP. (2.2)
Hence
EP = \E S. (2.3)
According to the linearized theory there is no limit to either Ep or Esy assuming the
wavelength L to be fixed. However from the fully nonlinear theory of gravity waves
the total energy Es of a standing wave has a maximum given by
( £ s )ma* = 0.07774 (2.4)
m units where p,g and the wavenumber к are all unity; see Mercer & Roberts (1992).
Hence, if we assume that all the energy of the incident and the reflected wave goes
2279
ak
Figure 2. Graph of the energy density Ep of a progressive gravity wave of finite steepness ak.
to form a periodic standing wave, and if we ignore the contribution of the higher
harmonics to the total energy (but not to the surface profile), then we must have
Ep ^ 5 С^5)тах — Есгц (2.5)
say, where from (2.4)
Ecru = 0.03887. (2.6)
If on the other hand Ep > Ecrit we see that the resulting motion cannot be periodic.
Now a plot o f the energy density Ep of a progressive wave against its steepness ak
(figure 2 ) shows that a progressive wave may have an energy as great as
(EP) max 0.0745 (2.7)
which is achieved when
ak = 0.429. (2.8)
The maximum wave steepness ak for a progressive wave is 0.4432. So from figure 2
and equation (2 .6 ) there is a certain range of progressive wave steepnesses, namely
0.285 < ak < 0.443 (2.9)
for an incident wave, such that the incident-plus-reflected wave system cannot be a
perfectly periodic standing wave. The resulting motion must be irregular or chaotic
in some way, leading possibly to breaking.
Incident progressive waves whose steepness lies in the range 0 < ak < 0.285 we
shall call subcritical, while those that lie in the range (2 .8 ) we shall call supercritical.
It is not of course implied that all subcritical incident waves will necessarily produce
motions that are perfectly periodic. This is a matter for experiment.
3. Experimental apparatus
The experiments were carried out in a glass-sided wave tank in the Hydraulics
Laboratory at the Scripps Institution of Oceanography. Its dimensions are shown in
figure 3. A computer-controlled, horizontal-movement wavemaker was at A, and an
impervious, sloping beach with gradient 1:10 was at C. A removeable plane wall,
or barrier, was inserted at B. The distances AB and AC were 15.65 m and 26.96 m
2280
respectively. The width W of the tank and the Stillwater depth h were both equal to
0.50 m.
The vertical displacement of the water surface was measured with resistence-wire
gauges, each consisting of two parallel vertical wires separated by 0.3 cm and oriented
across the tank at B, near the mid-plane of the wave tank. With the barrier in place,
the wires were 2 mm from the surface of the wall. Calibration before and after each
run indicated that the gauges were linear to within about 1 % over the range of
measurement. The output from the gauges was digitized and recorded at a rate of
100 Hz.
A video recording of the surface at a rate of 30 frames/s was made from a position
slightly above the mean water level and to the left (wave side) of the barrier. Blue
vegetable dye was added to the water to increase the contrast between air and water,
the background on the far wall being white.
4. P ro c e d u re
The experiments were conducted at a wave frequency / = 1 . 0 or 1.1s-1, which
is somewhat less than the cut-off frequency (1.25 s"1) for the lowest-order three-
dimensional waves in a channel of width 0.5 m. Nevertheless, after a certain duration
(about 45 s) some higher-order three-dimensional instabilities invariably made their
appearance. The most prominent of these was a cross-wave with wavelength 25 cm
(half the width of the channel) which appeared when / = 1.0 s” 1. Therefore most
experiments were done at / = 1.1 s-1.
2281
a0 a <*s
G (cm) (cm) aa2/g ak (cm)
1.6 4.53 4.22 0.206 0.200 7.50
1.8 5.19 4.63 0.226 0.216 6.95
2.0 5.52 5.66 0.251 0.236 7.87
2.2 5.82 5.51 0.268 0.252 8.49
2.4 6.61 5.85 0.285 0.266 9.28
2.6 7.52 6.36 0.310 0.285 10.89
Table 1. Range o f experiments with X = 0.1, N = 40; ak is determined by the procedure described
in the Appendix.
С
(cm)
-2
~0 10 20 30 40 50
/(s)
R gure 5. Record o f the surface elevation at x = 2.30 m when X = 0.1 and N = 40 (progressive
wave: G = 1.6).
As is well known, if a wavemaker is started from rest, the wave front advances down
the channel with the group velocity cg. If the wavemaker is switched on suddenly at
time t = 0 , then according to linear theory the surface elevation £(x,t) at a horizontal
distance x from the wavemaker is given by
С = iBF{x)eio(t- ° x/g) (4.1)
approximately, where В is a constant depending on the type of wavemaker, a is the
radian frequency of the waves, т is a dimensionless time:
-(± Г ( - т )
which vanishes at the wave front t = 2gx/ g, and F(t) describes the complex wave
ю-i+ArГ*****
envelope:
I 1 -г l Jo (43)
see for example Miles (1962). The function F ( t ) is related to the Fresnel integral
(Abramowitz & Stegun 1964). A sketch of the envelope | F ( t )| as a function of time
is shown in figure 4. The wave amplitude at first increases exponentially, reaches the
value | В at time т = 0 and then oscillates about its final value В as т —►oo. The
2282
s)
F ig u r e 6. Surface elevation at x = 15.65 m when G = 1.6 (a) with no barrier and (b) with the
barrier in place.
first maximum in the amplitude is about 19% greater than the final amplitude. It has
been shown experimentally (see Longuet-Higgins 1976) that the effect of finite wave
steepness is to increase the effective group velocity so that the wave front arrives
slightly sooner than predicted by the linear theory, and to increase the maximum
wave amplitude considerably.
In the present experiments, in order to suppress the oscillations of the envelope, the
wavemaker was started gradually from rest with a horizontal displacement given by
C= { °’ t< 0 (4.4)
[ С tanh At sin 2n f t, t > 0,
where С and Я are constants. It was found convenient to take Я = 0.1. When Я
was much smaller, the wave train often did not approach its final amplitude (e-2^1
negligible) before the onset of the three-dimensional instabilities mentioned above.
On the other hand, it was found useful to terminate the wave train after a certain
number N of wave cycles, by switching off the wavemaker suddenly. This produced
a corresponding Fresnel pattern of the wave envelope at the rear of the wave train,
including some waves which were steeper than the steady waves.
The input voltage to the wavemaker was governed by a certain gain factor, which
we have denoted by G. Experiments were carried out over the range 1.6 < G ^ 2.6;
see table 1 for the corresponding wave parameters.
2283
10 20 30 40 50 60 70
/(s)
At each value o f the gain G, four types of measurements were made. First, the
surface elevation £ was recorded at a point close to the wavemaker (x = 2.30 m) but
still far enough away that local effects were negligible, in general. The barrier at В was
not in place. Second, similar measurements were made at the point B, still without
the barrier, so that the waves passed by as progressive waves. Thirdly the barrier was
inserted, and the surface elevation was recorded at the same point B. In all three
cases the wave gauge was situated on the centreline of the channel. Simultaneously
with the third recording, a video sequence of the waves near the barrier was taken as
described above.
In order to prevent any horizontal displacement of the barrier it was constructed of
1.0 in. laminated plywood, strengthened by angle brackets, and was secured in place
by a steel bar at the top, clamps above the water level and a firm tubular rubber seal
at each sidewall. During the experiment no surface waves on the down-wave side of
the barrier were detected.
A step-calibration of the wire wave gauges was earned out at the beginning and
end of each series of experiments.
5. Results
Figures 5 and 6 show the case G = 1.6. In figure 5 we see the wave amplitude at
x = 2.30 m starting almost immediately to increase monotonically towards the value
312 M. S. Longuet-Higgins and D. A. Drazen
/(s)
R g u r e 8. Surface elevation at x = 15.65 m when G = 2.2 (a) with no barrier
(b) with the barrier in place.
4.5 cm, which is attained after about 25 s. At t = 38 s one can see the maximum T of
the Fresnel envelope created by the abrupt shut-off of the wavemaker after 40 cycles
(36 s).
Figure 6 (a) shows the same progressive wave train on arriving at В (x — 15.65 m).
The final amplitude a is now only 4.2 cm, attained at about t = 35 s. It remains
constant until about t = 42 s, after which it is affected by the Fresnel pattern from
the wave cut-off. From equation (4.2) the width of the Fresnel pattern is proportional
to (2 x /g )1/2 and so is increased over the width at x = 2.30 m by a factor 2.61. When
t > 30 s there are slight indications of a Benjamin-Feir instability, but these are small
compared to the oscillations of the Fresnel envelope. The period of the envelope
oscillations diminishes with distance from T. When t = 43 s their period is less than
two wave periods.
Figure 6(b) shows the same situation as 6 (a) but with the barrier in place. Note the
difference in vertical scales. The wave amplitude a5 is roughly equal to 2a (see table
1) as one would expect from linear theory.
The corresponding three records when G = 1.8 were quite similar to those in
figures 5 and 6 , but with larger amplitudes; see table 1. However when G = 2.0
some qualitatively new features appeared. Figures 7(a) and 7(b) show the records
taken at the point В (x = 15.65 m) without the barrier and with the barrier in place,
respectively. Figure 7(a) shows the usual Fresnel envelope for a progressive wave, with
2285
*(«)
R g u r e 9. Surface elevation at x = 15.65 m when G = 2.4 (a ) with no barrier
(b) with the barrier in place.
FIg ur e 11. A sequence of frames from the video corresponding to figure 9(b) (G — 2.4), showing
consecutive wave crests between t = 28 s and 42 s. The video was taken at 30 frames/s, and the
timing of each frame is as close as possible to a maximum of the surface elevation shown in
figure 9(b), that is within 1 /6 0 s. Each frame in the left-hand column above corresponds to one of
the maxima marked S, in figure 9(6).
2288
i
F i g u r e 12. Growth o f the triple-period instability, as measured by the difference AC in
crest-elevation between the highest and lowest waves o f a triplet.
ak
R gure 13. Graph o f ao2/g against ak for nonlinear progressive gravity waves in deep water.
CD ak (c2 — 1) aa2/g
0.00 0.00000 0.00000 0.00000
0.10 0.14222 0.02042 0.14512
0.20 0.20216 0.04173 0.21060
0.30 0.24877 0.06385 0.26465
0.40 0.28843 0.08674 0.31349
0.50 0.32346 0.11020 0.35911
0.55 0.33958 0.12203 0.38102
0.60 0.35488 0.13384 0.40238
0.65 0.36936 0.14552 0.42311
0.70 0.38303 0.15687 0.44312
0.75 0.39582 0.16767 0.46219
0.80 0.40765 0.17757 0.48000
0.85 0.41839 0.18601 0.49621
0.90 0.42782 0.19211 0.51001
0.95 0.43578 0.19454 0.52056
10.00 0.4432 0.1931 0.5288
T a b le 2. Corresponding values o f ak and aa2/g for deep-water gravity waves of finite amplitude.
REFERENCES
A b r a m o w itz , M. & S t e g u n , I. A . 1964 Handbook o f Mathematical Functions. Washington, D.C.,
U S Govt. Printing Office, 1046 pp.
C h a n , E. S. & M e l v i l l e , W. K. 1988 Deep-water plunging wave pressures on a vertical wall. Proc.
R. Soc. Lond. A 417, 95-131.
J ia n g , L., P e r l i n , M. & S c h u l t z , W. W. 1998 Period tripling and energy dissipation o f breaking
standing waves. J . Fluid Mech. 369, 273-299.
L o n g u e t - H i g g i n s , M. S. 1975 Integral properties o f periodic waves o f finite amplitude. Proc. R.
Soc. Lond. A 342, 157-174.
L o n g u e t - H i g g i n s , M. S. 1976 Breaking w aves-in deep or shallow water. In Proc. 10th Symp. on
Naval Hydrodynamics, Cambridge, Mass., June 1974 (ed. R. D. Cooper & S. W. Doroff), pp.
597-605. Arlington, VA, Office o f Naval Research, 792 pp.
L o n g u e t - H i g g i n s , M. S. & D o m m e r m u t h , D . G. 2001a On the breaking o f standing waves by
falling jets. Phys. Fluids 13, 1652-1659.
L o n g u e t - H i g g i n s , M. S. & D o m m e r m u t h , D . G. 20016 Vertical jets from standing waves. II. Proc.
R. Soc. Lond. A 457, 2137-2149.
M e r c e r , G. N. & R o b e r t s , A. J. 1992 Standing waves in deep water: their stability and extreme
form. Phys. Fluids A 4, 259-269.
M i l e s , J. W. 1962 Transient gravity wave response to an oscillating pressure. J. Fluid Mech. 13,
145-150.
Advanced Series on Ocean Engineering — Volume 35
The papers are quite varied in their approach. They range from basic theory and
new computational methods to laboratory experiments and field observations.
An overall feature is the frequent comparison between theory and experiment
and the constant attention to practical applications.
Among the many advances and achievements to be found in these three volumes
are: the now generally accepted solution to the longstanding problem of how
oceanic microseisms can be generated in deep water or near steep coastlines;
a theoretical explanation of the strong drifting near the bottom in shallow water;
the first introduction of a boundary-integral technique for calculating free surface
flows; simple analytic expressions for the form and time-development of plunging
breakers; and so on.