Sattinger D.H. - Scaling, Mathematical Modeling & Integrable Systems (1998) .Ps
Sattinger D.H. - Scaling, Mathematical Modeling & Integrable Systems (1998) .Ps
Scaling, Mathematical
Modelling, &
Integrable Systems
Initial Profile t=.2 t=.4
0 0 0
0 2 4 6 0 2 4 6 0 2 4 6
0 0 0
0 2 4 6 0 2 4 6 0 2 4 6
D.H. Sattinger
University of Minnesota
2
Preface
This project grew out of a conversation with Willi Jager in Heidelberg on a
visit in January, 1996. I discussed with Willi my investigations with Mariana
Haragus (now Haragus-Courcelle) at Stuttgart Universitat on the validity of
the Kortweg-deVries approximation to the Euler equations of gravity waves.
Many of the completely integrable systems of partial dierential equa-
tions arise as formal, singular asymptotic approximations, in some manner
or other, to more complicated nonlinear dispersive wave equations. It is a
curious fact that the methods of asmyptotic expansion and choice of scaling
should lead so often to completely integrable models. This artifact has never
been properly explained.
These derivations are generally formal in nature, and, with a few ex-
ceptions, have never been properly justied with mathematical rigor. Willi
proposed that I give a series of Deutsche Mathematische Vereinigung lectures
on the subject of scaling and mathematical modelling.
My hope here is to present an introduction to dispersive waves and the
derivation of some of the associated completely integrable systems, raise a
number of issues that have not been fully dealt with, introduce some method-
ologies (including the use of numerical animation of weakly nonlinear disper-
sive wave models using Matlab), and provide a comprehensive bibliography.
I would like to thank my colleagues and students for many helpful com-
ments and suggestions in the preparation of these notes, especially Sam Al-
bert, Jerry Bona, Mariana Haragus-Courcelle, Yi Li, Peter Olver, and Jacek
Szmigielski.
D.H. Sattinger
University of Minnesota
December, 1997
3
4
Contents
1 Dispersion 9
1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.2 Group and Phase Velocities . . . . . . . . . . . . . . . . . . . 10
2 Nonlinear Schrodinger Equation 21
2.1 Multiple Scales Expansion . . . . . . . . . . . . . . . . . . . . 21
2.2 Pulse solutions . . . . . . . . . . . . . . . . . . . . . . . . . . 25
3 Korteweg de-Vries 29
3.1 Background and History . . . . . . . . . . . . . . . . . . . . . 29
3.2 Plasmas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
3.3 Water Waves . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
3.4 The Solitary wave of the KdV equation . . . . . . . . . . . . . 46
4 Isospectral Deformations 49
4.1 The KdV hierarchy . . . . . . . . . . . . . . . . . . . . . . . . 49
4.2 The AKNS hierarchy . . . . . . . . . . . . . . . . . . . . . . . 54
5 Inverse Scattering Theory 59
5.1 The Schrodinger equation . . . . . . . . . . . . . . . . . . . . 59
5.2 First Order Systems . . . . . . . . . . . . . . . . . . . . . . . 73
5.3 Decay of the scattering data . . . . . . . . . . . . . . . . . . . 82
6 Variational Methods 85
6.1 Water Waves . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
6.2 Method of Averaging . . . . . . . . . . . . . . . . . . . . . . . 88
5
6 CONTENTS
7 Weak and Strong Nonlinearities 95
7.1 Breaking and Peaking . . . . . . . . . . . . . . . . . . . . . . 95
7.2 Strongly nonlinear models . . . . . . . . . . . . . . . . . . . . 99
7.3 The extended AKNS hierarchy . . . . . . . . . . . . . . . . . . 106
8 Numerical Methods 111
8.1 The nite Fourier transform . . . . . . . . . . . . . . . . . . . 111
8.2 Pseudospectral Codes . . . . . . . . . . . . . . . . . . . . . . . 119
List of Figures
1.1 Decay of a centered Gaussian. . . . . . . . . . . . . . . . . . . 15
1.2 A traveling oscillatory pulse. . . . . . . . . . . . . . . . . . . . 16
1.3 The Airy equation . . . . . . . . . . . . . . . . . . . . . . . . 18
2.1 Pulses of the NLS equation . . . . . . . . . . . . . . . . . . . . 26
2.2 Dark Soliton . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
3.1 Homoclinic orbit . . . . . . . . . . . . . . . . . . . . . . . . . 47
3.2 Two soliton interaction . . . . . . . . . . . . . . . . . . . . . . 48
6.1 A modulated pulse. . . . . . . . . . . . . . . . . . . . . . . . . 92
7.1 The extreme wave as conjectured by Stokes. . . . . . . . . . . 96
7.2 Breaking Wave . . . . . . . . . . . . . . . . . . . . . . . . . . 97
7.3 Peakons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
8.1 Dispersive dierence scheme . . . . . . . . . . . . . . . . . . . 115
8.2 Implicit spectral method . . . . . . . . . . . . . . . . . . . . . 116
8.3 Dispersive Cut-O . . . . . . . . . . . . . . . . . . . . . . . . 123
7
8 LIST OF FIGURES
Chapter 1
Dispersion
1.1 Introduction
In these lectures I shall describe the modelling of the competing eects of
dispersion and nonlinearity in physical systems. In particular, I shall discuss
in detail the derivation of the Korteweg-deVries and nonlinear Schrodinger
equations. The interplay of dispersion and nonlinearity is at the heart of the
phenomena which these equations purport to model.
The nonlinear equation
ut + uux = 0 (1.1.1)
develops shock discontinuities, even if the initial data is innitely dieren-
tiable. But when dissipation is added,
ut + uux = "uxx;
the solutions are globally smooth, even for arbitrarily small positive ".
The process of dispersion, on the other hand, is a more subtle phe-
nomenon. For example, the Korteweg-deVries equation
ut + uux = "uxxx
possesses globally smooth solutions, given smooth initial data, but the third
order derivative uxxx is neutral from the energy point of view. It neither
creates nor destroys energy. Instead, it acts to disperse energy at rates de-
pendent on the wave number, and the formation of shocks by the nonlinear
9
10 CHAPTER 1. DISPERSION
term uux is counterbalanced by the dispersion due to the term "uxxx, no
matter what value of " we take, positive or negative. The behavior of the
solutions as " ! 0 is called the zero dispersion limit, and has been studied
extensively [47]. This is of course, a rough statement of a general mathe-
matical principle that dispersion and nonlinearity often act to balance each
other.
In these lectures we shall focus on dispersive equations and their mod-
elling by asymptotic and scaling arguments. The arguments currently found
in the literature are, for the most part, purely formal, and lack mathemat-
ical justication; yet have considerable physical plausibility. They are thus
interesting grounds for mathematical investigation.
We begin by explaining the notions of group velocity and dispersion re-
lation.
We take the initial data u0(x) = f (x) and its Fourier transform to be
r
f (x) = eik0 xe x2 =4 ; fb(k) = e (k k0 )2 :
0.8 0.8
0.6 0.6
0.4 0.4
0.2 0.2
0 0
−0.2 −0.2
0 2 4 6 0 2 4 6
2 2
1.5 1.5
1 1
0.5 0.5
0 0
−0.5 −0.5
−1 −1
−1.5 −1.5
−2 −2
−2.5 −2.5
0 2 4 6 0 2 4 6
= x v0 t; y = k k0:
The following heuristic argument, due to Lord Kelvin, is routinely given in
the literature [74], p. 372. Since the Gaussian exp( y2) is rapidly decaying,
only the values of the integrand near y = 0 make a contribution, and the
term
ty3R0(y)
18 CHAPTER 1. DISPERSION
can be neglected. This is the method of stationary phase, rst proposed by
Lord Kelvin. The statement is roughly correct, but it is hard to quantify
since the method of stationary phase does not give the spatial dependence
of the integral
For example, consider the case of the simple equation
ut = uxxx
with dispersion relation ! = k3 . The exact solution is
u(x; t) = ei(k0 x k03t) A(; t); = x !0(k0)t;
Z1
A(; t) = exp[i(k 3k0tk2 tk3) k2 ] dk:
1
2 2 2
1 1 1
0 0 0
−1 −1 −1
−2 −2 −2
Figure 1.3: The evolution of a Gaussian pulse under the Schrodinger equation
vs. the Airy equation.
The function A(; t)) is a modied Airy integral. If the term eitk3 is
ignored, we obtain the Gaussian integral
Z1
G(; t) = eik e (3ik0 t+)k2 dk:
1
1.2. GROUP AND PHASE VELOCITIES 19
The integral A can be evaluated approximately by the method of steepest
descent [28]. The functions G and A satisfy the following linear partial
dierential equations:
Gt = 3ik0G ; At = 3ik0A + A :
The real parts of the solutions of the two equations with a modulated
Gaussian pulse as initial data are displayed in Figure 1.3.
20 CHAPTER 1. DISPERSION
Chapter 2
The Nonlinear Schrodinger
Equation
The nonlinear Schrodinger equation arises in a multitude of contexts, but its
best known application is to the self-focussing of one-dimensional electromag-
netic waves, for example optical pulses along an optical ber, in nonlinear
media [76]. There are a number of formal derivations of this equation, one
of which is based on multiple scales expansions and is described below.
The nonlinear Schrodinger equation is the principle model equation in
nonlinear optics, that is, in the modelling of propagation of electromagnetic
pulses in nonlinear optical bers, which act as wave guides. G. B. Whitham
[74], p. 549, has derived the nonlinear Schrodinger equation by a method
of averaging the variational equations. This procedure is closely related to
the method of multiple scales; but it has several conceptual advantages.
Whitham's method will be discussed in x5.3.
The nonlinear Schrodinger equation was shown to be integrable by the
method of inverse scattering by V.E. Zakharov and A.B. Shabat [76]. We
will discuss the inverse scattering theory associated with this model in x4.4.
where @2
@ 2 + m2 :
L0 = @t2 @x 20
0
We look for a solution of the rst equation in the form
U1 = <A(x1 ; x2; t1; t2 )ei ; = kx !t; !2 = k2 + m2:
This term is to be interpreted as a carrier wave (ei ) with a modulation by a
complex envelope A(x1 ; x2 ; t1; t2 ) in the slow variables. Then U1 is a solution
of the rst equation, and we shall derive a nonlinear equation, in fact the
nonlinear Schrodinger equation, for the complex envelope A.
When we substitute U1 into the equation at second order (2.1.1) we get
@A @A
L0 U2 = < 2i k @x + ! @t ei
1 1
The presence of the term ei on the right creates secular terms in the solution,
since it is in the null space of the operator L0. Thus, one solution of
@2 @2
2 i
@t20 @x20 + m u = e
is
ei ;
2im2
which grows linearly in both x and t.
In order to eliminate such secular terms we require that
@A + ! @A = 0
k @x
1 @t1
This equation implies that A is a function of the variable 1 = !x1 kt1 ,
that is
A(x1 ; x2 ; : : : ; t1; t2; : : : ) = A(!x1 kt1 ; x2 ; t2; : : : ) = A(1; x2 ; t2; : : : ):
24
CHAPTER 2. NONLINEAR SCHRODINGER EQUATION
A is a waveform traveling to the right with velocity k=! in the slow variables
x1 ; t1. Note that k=! is the group velocity of the wave:
@! = k :
@k !
Thus, in a frame of reference moving with the group velocity, the secular
terms vanish.
The equation for U2 is then simply
@2
@ 2 + m2 U = 0;
@t20 @x20 2
=ik @A
@x
2 + i! @A2
@t
1 1
=ik! @A
@
2 i!k @A 2
@ = 0:
To avoid secular terms in the equation for U3 we again require the coe-
cients of ei on the right side of the equation to vanish. A calculation shows
the coecient of ei on the right side to be
@A @A
1 @2
@ 2 A + 3 jAj2A:
i k @x + ! @t 2 @t21 @x21 8
2 2
Since !2 k2 = m2,
@2
@ 2 A ( ; ; ) = 1 ( ! 2 k 2 ) A = 1 m2 A ;
@t21 @x21 1 2 2 2 1 1
2 1 1
2.2. PULSE SOLUTIONS 25
hence the necessary and sucient condition that the secular terms vanish is
that A satisfy the nonlinear partial dierential equation
@A @A 1
i k @x + ! @t + 2 m2 A1 1 + 38 jAj2A = 0:
2 2
Introducing the coordinates
2 2
2 = (!x2 kt2 ) 2m!k ; 2 = (!x2 + kt2 ) 2m!k ;
we obtain, after some simplications, the equation
iA2 + 21 A1 1 + 2 jAj2A = 0; = 43m2 ; (2.1.3)
as the equation to remove secular terms. This is the nonlinear Schrodinger
equation.
The case > 0 (hence > 0) is called the self-focussing case. This is
the case where there are solitons. Note that the scales are mixed, the \time
variable" 2 being on the second order scale, while the \space variable" 1
is on the rst order scale. If the wave envelope A were a function only of
1; 2, then the pulse would travel with the group velocity, without change
of shape. It is a function of 2 however, so it is modulated on a scale of order
"2 as it travels.
This formal argument extends in a more or less straightforward fashion
to a variety of nonlinear wave equations. This is done in many places in the
literature. A rigorous justication of such expansions in a variety of cases,
including nonlinear wave equations, has been given by Kirrmann, Schneider,
and Mielke [41].
0.8 b=30
0.8 b=20
0.6
0.6
0.4
0.4
0.2
0
0.2
−0.2
0
−0.4
−0.2
−0.6
−0.4 −0.8
−5 0 5 −5 0 5
Figure 2.1: Pulses of the nonlinear Schrodinger equation for the parameter
values b = 20; 30.
0.8
0.6
0.4
0.2
−0.2
−0.4
−0.6
−0.8
−1
−10 −8 −6 −4 −2 0 2 4 6 8 10
Figure 2.2: Prole of the real part of a dark soliton for the parameters
a = 8; b = 32:09,
= 2 = 0.
In the de-focussing case, <p0; > 0, the origin is a center for a family
of closed periodic orbits, and = are now unstable equilibria joined
by two heteroclinic orbits. These are the so-called dark solitons, shown in
Figure (2.2). There are two such heteroclinic orbits to (2.2.4) in this case,
r 1=2
u(x) = ! tanh 2 (x
); ! =
28
CHAPTER 2. NONLINEAR SCHRODINGER EQUATION
so the solutions to (2.1.3) are
r
! ei(a1 +b2 ) tanh 2 (1 a2
)
Chapter 3
The Korteweg deVries
Equation
3.1 Background and History
The Korteweg-deVries equation,
ut + uux + uxxx = 0 (3.1.1)
is the simplest equation that includes both the eects of nonlinearity and
dispersion. The equation appears in various forms in the literature, some-
times with a factor of 6 or -6 in front of the nonlinear term. We shall use
the present normalization, since that is the unique normalization for which
the Korteweg-deVries equation is Galilean invariant. The rst order equation
(1.1.1) is known to develop shock discontinuities in nite time, even with C 1
initial data; but when the term uxxx is added, the solutions remain smooth
for all time. Even more, the equation has been found to possess remarkable
mathematical structure. It is an innite dimensional Hamiltonian system
that, in a precise sense, is completely integrable. Moreover, the equation
models some of the phenomena found in more complicated systems of non-
linear partial dierential equations, such as the Euler equations for waves on
the surface of an inviscid, irrotational
uid.
In this chapter we discuss the derivation of the Korteweg-deVries equation
as a model for long waves in dispersive media. It also arises as an approximate
model in numerous other elds, including magnetohydrodynamics waves, ion-
acoustic plasma waves, and anharmonic lattice vibrations. Martin Kruskal,
29
30 CHAPTER 3. KORTEWEG DE-VRIES
who played a major role in the discovery of the fundamental mathematical
properties of the equation, states [42]
The essential signicance of the KdV equation in the present context is
that it is the unique leading order reduced equation approximation.
An excellent reference for the theory of water waves prior to 1957 is J.J.
Stoker's treatise [70]. An account of the history of the rigorous mathematical
theory of the solitary and periodic waves is given beginning on p. 342. The
variational principle described in Chapter 6, however, was only discovered in
1967, and so no account of it is to be found in Stoker's book, After the com-
plete integrability of the Korteweg-deVries equation was discovered in 1967,
the theory of the KdV equation and its relationship to the Euler equations
as an approximate model obtained from perturbation theory became of more
interest; and Whitham's book [74] treats the KdV approximation in some
detail.
In addition, a history of the Korteweg-deVries equation and its role as a
model equation for water waves can be found in the articles by Bona [15],
and Newell [52], and so I will only review the highlights which are relevant
to our discussion.
The original observation of the solitary wave in water was reported by
Russell [65]1
\I believe I shall best introduce this phenomenon by describing the cir-
cumstances of my own rst acquaintance with it. I was observing the motion
of a boat which was rapidly drawn along a narrow channel by a pair of horses,
when the boat suddenly stopped { not so the mass of water in the channel
which it had put in motion; it accumlated round the prow of the vessel in
a state of violent agitation, then suddenly leaving it behind, rolled forward
with great velocity, assuming the form of a large solitary elevation, a rounded,
smooth and well-dened head of water, which continued its course along the
channel apparently without change of form or diminution of speed. I followed
it on horseback, and overtook it still rolling on a a rate of some eight or nine
miles an hour, preserving its original gure some thirty feet long and a foot
to a foot and a half in height. Its height gradually diminished, and after a
chase of one or two miles I lost it in the windings of the channel. Such, in
This passage, as well as discussion of other topics related to Russell's observations,
1
can be found in the monograph by Alan Newell.
3.1. BACKGROUND AND HISTORY 31
the month of August 1834, was my rst chance interview with that singular
and beautiful phenomenon which I have called the Wave of Translation, a
name which it now very generally bears."
Russell's subsequent experiments stimulated great interest in the subject
of water waves, and his ndings were immediately taken up by Airy [2] and
Stokes [71]. Stokes took special note of Russell's assertion that \the velocity
of propagation of a series of oscillatory waves does not depend on the height of
the waves." Stokes computed the Fourier series of the formal approximations
to second order in the case of nite depth, and to third order in the case of
innite depth.
\I have proceeded to a third approximation in the particular case in which
the depth of the
uid is very great . . . This term gives an increase in the
velocity of propagation depending on the square of the ratio of the height of
the waves to their length."
Among the many conclusions of his investigations, Stokes concluded, er-
roneously, as it turns out, \that it is only an indenite series of waves which
possesses the property of being propagated with a uniform velocity."
From our present-day vantage point we can see that the controversy sur-
rounding the discovery of the solitary wave was due precisely to the problem
of correctly balancing the eects of dispersion and nonlinearity in the models.
If dispersion is ignored, one obtains the shallow water equations, of which
(1.1.1) is a simple prototype, which produces shocks, or breaking of waves.
At the other limit, if one ignores the nonlinear eects, he obtains an equation
of the type
ut + ux + uxxx = 0 (3.1.2)
which exhibits only a dispersive decay of the waves, and supports neither
periodic wave trains, nor solitary waves.
Boussinesq [18], [19] and Rayleigh [61] took up the investigations of such
solitary waves in the 1870's. Boussinesq [19] actually found the equation
later named for Korteweg and deVries, and, moreover, found the exact sech2
solution for the solitary wave. Boussinesq's treatise was, however, 630 pages
long, and much of his work was not fully appreciated. The Korteweg deVries
32 CHAPTER 3. KORTEWEG DE-VRIES
equation, as we shall see, exhibits both periodic wave trains (the cnoidal
waves), and, in the limit of innite period, the solitary wave.
There are three relevant length scales in the theory, h, the depth of the
uid; l, the length of the wave; and a, the amplitude of the wave. Accordingly,
there are two dimensionless parameters,
" = hl ; = ha :
The KdV approximation is valid in a regime where = O("2); and this is
where the solitary wave occurs.
Levi-Civita [48] gave a rigorous proof of the existence of periodic wave
trains in innitely deep water; Struik [72] modied Levi-Civita's analysis to
prove the existence of periodic wave trains in water of nite depth. Levi-
Civita used a conformal mapping technique to map the unknown domain
conformally into a half-plane. Levi-Civita's method is of considerable inter-
est, and may prove also to be useful in other problems. Friedrichs and Hyers
[33] gave a rigorous construction of the solitary wave.
Before describing the derivation of the KdV equation in ion acoustic waves
and in water waves, which is considerably more involved, I would like to
present a very simple derivation of the KdV equation due to Kruskal [42].
He begins with the weakly nonlinear dispersive wave equation2 governing the
evolution
utt = uxx(1 + "ux) + uxxxx;
where "; 1 are small parameters measuring nonlinearity and dispersion
respectively. The linear equation, with " = 0, has dispersion relation
!2 = k2 k4;
so that we must really require < 0 for well-posedness.
For = " = 0 the equation reduces to the linear wave equation, which
has as a general solution left and right progressing wave-forms. We look for a
solution of the full equation which is essentially right progressing, with only
a slow variation. To make this precise, we look for a solution of the form
u(x; t; "; ) = w(; ) = x t; = "t:
This is the continuum model for the Fermi-Pasta-Ulam system [29]. Cercignani [23]
2
and Palais [59] have given an account of the Fermi-Pasta-Ulam experiment and its rela-
tionship to the Korteweg-deVries equation.
3.1. BACKGROUND AND HISTORY 33
By the chain rule
@u = " @w @w ; @u = @w ;
@t @ @ @x @
and, in the variables ; the equation becomes
"2w 2"w + w = w (1 + "w ) + w :
We drop the quadratic term in " since it is second order. Putting U = w =2
we obtain the KdV equation for U :
U + UU + 2" U = 0:
This delightfully simple derivation indicates two of the primary ingredi-
ents in the KdV approximation: First, one must specialize to a unidirectional
frame; and Second, one must scale the time variable in an appropriate way.
These two features also appear in the more complicated derivation of the
KdV approximation to the plasma and Euler equations.
Before proceeding with these derivations, I mention the important paper
of Benjamin, Bona, and Mahoney [11], who argue that "in all examples the
assumptions leading to the KdV equation equally well justify the equation
ut + ux + uux uxxt = 0:
Their argument is that, since to lowest order the solution represents a right
moving wave, ut ux and so one may replace uxxx with uxxt. This
argument is disputed in rather strong terms by Kruskal, cf. [42], though
Bona, Pritchard and Scott [16], [17] have conducted extensive numerical and
physical experiments to ascertain the validity of both the KdV and BBM
approximations. Their conclusions are that \on a long time scale T naturally
related to the underlying physical situation, the equations predict the same
outcome to within their implied order of accuracy". [16] In other words, both
the KdV and BBM approximations are accurate for a nite time scale3, and
they are equally valid models over that time interval.
Craig [25] has shown that the KdV equation is accurate on a time scale of order " 3 ,
3
where " is a small parameter in the theory.
34 CHAPTER 3. KORTEWEG DE-VRIES
3.2 Ion Acoustic Waves in Plasmas
The Korteweg-deVries equation was obtained as a model for magetohydro-
dynmic waves in a cold collsionless plasma by Gardner and Morikawa in an
unpublished article in 1960 [36]. Since that derivation is somewhat easier
than in the case of water waves, we shall discuss that case rst.
We consider a plasma consisting of negatively charged electrons and pos-
itively charged ions. The electrons are treated as a gas and we obtain equa-
tions of motion for the ions. The ion density is denoted by n, the electron
density by ne, the electric force eld by E and the velocity of the ions by v.
The equations of the plasma may be written in the following form [44], [69]
nt + (nv)x = 0; vt + vvx = E;
E + (log ne)x = 0; Ex + ne = n;
where n is the ion density, ne is the electron density, v is the ion velocity,
and E is the electric eld.. We eliminate ne from the equation by dening
' = log ne, ' being the electric potential, and the equations reduce to three
equations in three unknowns [26]
v 2
nt + (nv)x = 0 vt + ( 2 + ')x = 0 'xx e' + n = 0: (3.2.3)
There are a number of perturbation arguments one can try on this system
of equations; and since there is a certain art to the method, I will carry out
the discussion in some detail. First let us try a straightforward perturbation
approach for small disturbances of the equilibrium states. We look for small
perturbations about the quiescent state n = 1; v = c; ' = 0:
n = 1+ "n1 + "2n2 + : : : ; ' = "'1 + "2'2 + : : : ; v = c + "v1 + "2v2 + : : : :
Substituting these expansions into (3.2.3) we obtain, at lowest order, the
linear equations for small disturbances
n1;t + (v1 + cn1 )x = 0;
v1;t + (cv1 + '1 )x = 0;
'1;xx '1 + n1 = 0:
3.2. PLASMAS 35
We determine the dispersion relation for this linear system by looking at
the Fourier modes 0n 1 0a 1
@ v11 A = @a12A ei(kx !t):
'1 a3
This leads to the linear algebraic system
0! ck k 0
1 0a 1
@ 0 ! ck k A @a12A = 0:
1 0 1 + k2 a3
The system has a nontrivial solution i the determinant of the above matrix
vanishes, and this leads to the condition (! ck)2 (1+ k2) k2 = 0, or, solving
for !,
! = ck p k 2 :
1+k
This dispersion relation is unbounded in k; but it is of quite a dierent nature
than the dispersion relation for the nonlinear Klein Gordon equation. We
will return to it in a minute.
Next, let us consider the eect of rescaling the space and time variables.
Namely, we make the transformations x0 = "x, t0 = "t. In the rescaled
variables the equations take the form
nt0 + (nv)x0 = 0;
2
vt0 + ( v2 + ')x0 = 0;
"2'x0x0 + e' = n:
Letting " ! 0 we obtain formally e' = n, and the nonlinear system
2
nt0 + (nv)x0 = 0; vt0 + ( v2 + log n)x0 = 0:
This is a nonlinear hyperbolic system with characteristic speeds 1; for gen-
eral initial data such systems of equations have shock discontinuities.
The scaling we have just described eliminates the dispersion. In fact, the
linearization of the equations above is the rst order hyperbolic system
nt + vx = 0 vt + nx = 0
36 CHAPTER 3. KORTEWEG DE-VRIES
and this system propagates pulses unchanged; i.e. it is dispersionless. We
may conjecture that the dispersion of the full plasma equations prevents the
build-up of shocks.4
We now derive an approximation to the plasma equations which preserves
the dispersion and leads to the Korteweg-deVries equation. Consider the
dispersion relation for the linearized plasma equations. The so-called long
wave length approximation is to formally approximate !(k) by its Taylor
expansion about k = 0. k = 2=L, where L is the wavelength; k is called:
the wave number. The Taylor expansion of the dispersion relation is ! =
k 21 k3 + O(k5). This is a good approximation for small k, that is, for long
waves.
The leading term k in this approximation corresponds to a dispersionless
system with a wave speed 1; for, if ! = k then
Z1 Z1
b
ei(kx !t) f (k) dk = eik(x t) fb(k) dk = f (x t)
1 1
where fb is the Fourier transform of f . This suggests transforming to a frame
of reference moving with unit speed. Waves traveling to the left are obtained
by taking the negative root for ! in the dispersion relation.
Now observe that the plasma equations are Galilean invariant; that is,
they are unchanged under the one-parameter group of transformations
x 7! x ct; t 7! t v 7! v + c:
This means that the equations are the same in any reference frame, and
we can shift to a moving frame of reference simply by adding a constant
to the velocity v. This amounts to expanding about the quiescent state
n = 1; v = c; ' = 0, where c is the velocity of the reference frame. If our
dispersion relation was originally
p
! = ck p k 2 = p k 2 ( 1 + k2 1);
1+k 1+k
we choose the velocity of the reference frame to be c = 1. Then the dispersion
relation is
! = k p k 2 = k(1 [1 12 k2 + : : : ]);
1+k
This is precisely the situation for the KdV equation itself. Cf. Lax, Levermore, and
4
Venakides [47] for a discussion of the zero dispersion limit of the KdV equation.
3.2. PLASMAS 37
and we obtain, in the long wave approximation,
! =: 21 k3:
The partial dierential operator associated with this dispersion relation
is
@ 1 @3 ;
@t 2 @x3
for which the natural scaling is x0 = "x, t0 = "3t. If we introduce this scaling
into the equations (3.2.3) we obtain (after division by ")
"2nt0 + (nv)x0 = 0;
v2
"2 vt0 + 2 +' = 0;
x0
"2'x0x0 + e' = n:
This perturbation scheme is singular, since the character of the equations
is changed when " = 0.
Since only "2 appears in these equations, we expand all quantities in
powers of "2:
n = 1 + "2n1 + : : : ; v = 1 + "2v1 + : : : ; ' = "2'1 + : : : :
When we do this, substitute the expansions into the above equations, and
collect terms, we get at order "2:
(n1 + v1 )x0 = 0; (v1 + '1 )x0 = 0; '1 = n1 :
Assuming v1 and n1 tend to zero as x ! 1 we have n1 = v1 .
At next order we obtain
n1;t0 + (n1 v1)x0 + (n2 + v2 )x0 = 0;
v2
v1;t0 + 1
2 + v2 + '2 x0
= 0;
'y = 0 on y =0:
Here, g is the acceleration due to gravity, the
uid density, and the
displacement of the
uid surface from equilibrium. We have neglected surface
tension. At rest, the
uid lies in the region 0 y h.
The rst equation is Laplace's equation for the velocity potential ', while
the third is Bernouilli's equation at the free surface. The second equation is
a kinematic condition that states that the relations
(x;_ y_ ) = r' and y = h + (x; t)
are consistent. Taking the total (Lie) derivative of the second equation, we
have
y_ = t + xx_ = t + x'x = 'y :
We now turn to a derivation of approximate models to the Euler equa-
tions. There are a number of derivations of the Boussinesq and Korteweg-
deVries equations in the literature, e.g. [9], [74]. Peter Olver [56] has derived
the approximations from a Hamiltonian perspective. We begin by rewriting
the equations in non-dimensional quantities. There are three characteristic
length scales in the problem: l, the length of the wave, the depth of the layer
h, and a, the amplitude of the wave. If we take the natural length scale
3.3. WATER WAVES 39
of the problem to be the length of the wave, l, then there are a priori two
dimensionless parameters:
" = hl ; = al :
The velocity potential ' has dimensions of velocity length. We need
to determine an appropriate velocity scale. By dimensional analysis [12] one
can
pgl argue that the velocity of a wave in deep water is a constant multiple of
where l is the wave length. But p in our case there is also the possibility
that the velocity is proportional to gh. This indeterminacy can be resolved
by examining the dispersion relation at k = 0, namely, by computing d!=dk
at k = 0.
The dispersion relation is obtained from a Fourier analysis of the equa-
tions linearized about the state
= 0; ' = cx 21 c2t:
This is a special solution of the Euler equations which corresponds to a state
in which the
uid is moving with uniform horizontal velocity c, and there is
no displacement of the surface.
The linearized equations about this state are:
' = 0; 'y y=0 = 0;
'y =0 y = 0: (3.3.7)
The parameter can be scaled out of the equations by setting w = .6
When this is done the equations take the form
"2'xx + 'yy =0; 0 y 1 + w; (3.3.8)
"2(wt + wx'x) ='y ; y = 1 + w; (3.3.9)
'y =0 y= 0: (3.3.11)
Before proceeding, it will be useful to note that (3.3.9) can be written in
the simplied form
d Z 1+w
wt + dx 'x dy = 0: (3.3.12)
0
6We shall see below that this scaling of the surface height is consistent with the re-
quirement that = O("2 ) in the KdV approximation. Whitham [74] instead rescales
the velocity potential '. This has the consequence of keeping the parameter in the
equations.
42 CHAPTER 3. KORTEWEG DE-VRIES
This form is obtained as follows. From Laplace 's equation (3.3.8) and the
boundary condition 'y = 0 on y = 0 we obtain
Z 1+w
'y y=1+w = " 2 'xx dy:
0
Now note that
d Z 1+w ' dy = w ' + Z 1+w ' dy:
x x x xx
dx 0 0
This is a singular perturbation problem when " 1, that is, when the
waves are long compared to the depth of the
uid layer h. This is the shallow
water case. We look for solutions by expanding in powers of "2 :
' = '0 + "2'1 + : : : ; w = w0 + "2w1 + : : : :
Substituting these expansions in (3.3.4) we obtain, successively,
0 ='0;yy + "2('0;xx + '1;yy ) + "4('0;xx + '1;yy ) + : : : (3.3.13)
0 = '0;y + "2(w0;t + '0;xw0;x '1;y + : : : (3.3.14)
1.5
E=0
1
homoclinic orbit
0.5 E<0
−0.5
−1
−1.5
−0.5 0 0.5 1 1.5 2 2.5 3 3.5
In addition to the solitary wave and the periodic wave, the KdV equation
has other special solutions which are interesting. Foremost among these are
the multi-soliton solutions. For any integer n 1 these are given explicitly
by the formula
d2 dj e 8!j3 te !j +!k)x
Un(x; t) = 12 dx2 log jk + !j + !k ;
where !j and dj are positive constants, with !1 < !2 < < !n. This Un
constitutes a 2n parameter family of exact solutions of the KdV equation,
parametrized by !j and dj . Asymptotically, as t ! 1 these solutions tend
uniformly in x and exponentially fast in time [38] to a superposition of n
solitary waves. Roughly, the dj constitute relative phases of the waves, while
the wave speeds are given by 4!j2; j = 1; : : : n.
The interaction of two solitary waves is depicted in the sequence Figure
48 CHAPTER 3. KORTEWEG DE-VRIES
3.2.
Initial Profile t=.2 t=.4
1 1 1
0 0 0
0 5 0 5 0 5
1 1 1
0 0 0
0 5 0 5 0 5
Figure 3.2: The interaction of 2 solitary waves in the exact solution. Two
solitary waves are pictured in the rst frame. As time progresses the two
solitary waves interact and separate. Note the dip in the larger wave as they
interact, indicating clearly that the interaction is nonlinear and not a simple
superpostion. After the interaction they regain their shape, but are displaced
from where they would have been had there been no interaction.
As t ! 1, Un(x; t) looks like a superposition of n solitary waves. As
time evolves, the faster waves overtake the slower (smaller amplitude) waves
and interact nonlinearly. As t ! 1 the waves are arranged in the order of
increasing amplitude and again assume the shape of the sech2 solitary wave.
The only evidence of the nonlinear interaction is a phase shift in the waves.
That is, they are displaced from the position they would occupy if they
were simply a linear superposition of solitary waves. In the appendix I have
inserted a Matlab program which numerically integrates the KdV equation
with a superposition of two solitary waves as initial data.
Chapter 4
Isospectral Deformations
4.1 The KdV hierarchy
The Korteweg-deVries (KdV) equation,
ut + uxxx + uux = 0; (4.1.1)
has smooth solutions for all positive and negative time given initial data
which is suciently smooth, say C 3. We won't go into the existence theory
of the KdV equation here since the existence and uniqueness issues are ulti-
mately resolved very simply by the method of inverse scattering, and since
the algebraic/ geometric structure of this simple nonlinear equation is far
more interesting.
The KdV equation has as a special solution the solitary wave,
u(x; t) = 12c2sech2 [c(x 4c2t)]
These waves move to the right with speed 4c2. Note that their amplitude
depends on the wave speed, and that larger waves travel faster. One could
choose as initial data two solitons separated by a distance great enough so
that their interaction was extremely small, since they decay exponentially in
either direction. Suppose the soliton to the left is larger. As time evolves, the
larger soliton will overtake the smaller one. Since the equation is nonlinear
they will react in a nonlinear way. After a period of time the two solitons
again separate, the larger one moving ahead to the right and regaining its
original shape. For large time, the two solitons are perturbed only by a
phase shift: they are not quite where they would be had they been purely
49
50 CHAPTER 4. ISOSPECTRAL DEFORMATIONS
solitary waves. These facts were discovered by computational experiments
by Kruskal and Zabusky in the mid 60's [43].
Moreover, the same thing happens when the initial data consists of several
solitons, separated originally into distinguishable solitary waves. As time
progresses, the faster solitons overtake the slower ones, and as time goes
to innity, the solution evolves into separated solitons, each with its own
original amplitude and speed, but with slightly displaced phase. In Chapter
8 I have described a Matlab le that numerically computes and animates the
two soliton interaction.
The discovery of the elastic scattering of the solitons for the KdV equa-
tion prompted an intense, and, as it turned out, highly fruitful theoretical
investigation of the KdV equation. The original theoretical breakthrough
was made by Gardner, Greene, Kruskal, and Miura, [35] Later researchers
claried and simplied their arguments, and ultimately constructed myriad
further examples of such special systems. One of the early papers which had
an enormous in
uence on the development of the subject was the 1968 paper
by Peter Lax [45]. Gardner, Greene, Kruskal, and Miura [35] had found that
the eigenvalues of the Schrodinger operator
L = D2 + 61 u; D = dx d;
are constant in time if u evolves according to the KdV equation.
The early papers in the subject were complicated by the extensive compu-
tations that accompanied the original discoveries. Lax simplied and claried
the situation conceptually by introducing the Heisenberg picture. Suppose
that the family of operators fL(t)g is unitarily equivalent under the
ow.
Assume U is a one-parameter family of unitary operators:
UU = I; Ut = BU; U L(t)U = L(0);
where B is a skew-adjoint operator. Dierentiating the third equation with
respect to time we get
U B L(t)U + U Lt U + U LBU = 0;
hence
Lt + B L + LB = 0:
In these calculations we interpret Lt as the operation of multiplication by
the function ut. Since B = B this equation reduces to the Lax equation
Lt = [B; L] (4.1.2)
4.1. THE KDV HIERARCHY 51
where [B; L] is the commutator BL LB . The pair of operators L and
B is called a Lax pair. Equation (4.1.2) is the Heisenberg picture of the
Korteweg-deVries equation.
Since Lt is a multiplication operator, the commutator [B; L] must also be
a pure multiplication operator. For example, taking B = D we nd
[D; L] = 1 ux; L 1
t = ut
6 6
and the Lax equation is ut = ux. This equation generates the one parameter
family of translations, u(x; t) = u0(x + t); and so, of course, L is unitarily
equivalent under the
ow.
The KdV equation itself is obtained by taking a third order skew adjoint
operator
B = 4D3 + 21 (uD + Du)
The details of the calculation are left as an exercise.
In all these calculations we may replace L by L + k2, so the KdV equation
is formally obtained as a consistency condition for the overdetermined system
of partial dierential equations
(D2 + 16 u + k2 ) = 0; @ =B :
@t
The two isospectral
ows
ut = ux; ut + uxxx + uux = 0
are only two
ows in an innite hierarchy of commuting Hamiltonian
ows.
this hierarchy of
ows is generated by a recursion relation, namely
DFj+1 = ( D3 31 (uD + Du))Fj ; F1 = u: (4.1.3)
The j th
ow is then given by
ut = DFj :
The recursion relation (4.1.3) will be derived in section (7.2). This recursion
relation was rst proposed by A. Lenard; Peter Lax [46] showed that each Fj
52 CHAPTER 4. ISOSPECTRAL DEFORMATIONS
in the recursion relation is a dierential polynomial in u and, furthermore, is
the gradient of a functional, Hj , so that the
ows have the form
Hj :
ut = D u
For example, the rst two terms in this recursion relation and their corre-
sponding functionals and
ows are
Z1 1
F1 = u; H1 = 2
2 u dx; ut = ux;
1
u2 ; Z1 1 3
F2 = D 2 u H1 = u 2 u dx; ut = uxxx uux:
2 2 x 6
1
The operators D; (D3 + 31 (uD + Du) are an example of a bi-Hamiltonian
pair. Bi-Hamiltonian pairs of operators can be used to generate hierarchies
of commuting Hamiltonian
ows [51], [57].
The n soliton solutions of the Korteweg de-Vries equation also have a
variational characterization. The KdV equation has an innite number of
conserved quantities, which may be represented as integrals. These are gen-
erated, for example, by the Lenard relation discussed above. The rst few
are:
Z1 Z1 Z1 u2 1 u3 dx
I0 = u dx; I1 = u2 dx; I2 = x
2 6
1 1 1
Z1 2
I3 = au4 + buu2x + u2xx dx
1
for some constants a and b. These integrals are all in involution with respect
to the Gardner Poisson bracket, dened by
Z1 F d G
fF; Gg = dx:
u dx u
1
4.1. THE KDV HIERARCHY 53
That is,
fIj ; Ik g = 0:
Each of the functionals Ij is conserved under the KdV
ow; and each of them
generates a Hamiltonian
ow.
The multi-soliton solutions can be obtained from a simple variational
principle. For example, consider the constrained variational problem
min I :
I1 =c1 2
The Euler-Lagrange equations for this constrained minimum are
I2 + c I1 = 0:
u u
This leads immediately to (3.4.20), with the Lagrange multiplier c precisely
equal to the wave speed. More generally, the n-soliton solution is obtained
as the constrained variational problem
min I
I1 =c1 ;:::In =cn n+1
The associated Euler-Lagrange equation for the n soliton solution is then an
ordinary dierential equation of order 2n + 1:
In+1 + c In + : : : c I1 = 0:
u n
u 1 u
The orbital stability of the n-soliton solutions can be proved using this
variational characterization [50]. Such variational arguments go back to
Boussinesq and have been used more recently by Benjamin [8] to prove the
stability of the solitary wave. A simplied sketch of the argument is as fol-
lows. Since all the integrals Ij are conserved under the KdV
ow, the
ow is
restricted to a manifold Mn dened by
Mn = fu : I1(u) = c1; : : : ; In(u) = cng:
The n-soliton solution Un is obtained as the minimum of the functional In+1
on this manifold. Hence the level surfaces of In+1 on Mn in a neighborhood
of Un are closed manifolds surrounding Un. If the initial data is initially in
a neighborhood of Un , it must therefore remain in that neighborhood under
the
ow.
54 CHAPTER 4. ISOSPECTRAL DEFORMATIONS
4.2 The AKNS hierarchy
All of the known examples of innite dimensional completely integrable sys-
tems arise as compatibility conditions for an overdetermined system of equa-
tions. Two well-known examples, the modied KdV equation
ut + u2ux + uxxx = 0; (4.2.4)
and the nonlinear Schrodinger equation
ut = iuxx + 2ijuj2u; (4.2.5)
arise in such a way. They can be obtained as `zero-curvature' conditions for
the connection
Dx = @x@ zJ q; D = @ A (4.2.6)
t
@t
where J is a 2 2 diagonal matrix, q is a 2 2 o-diagonal matrix, and A is
a traceless 2 2 matrix which is a dierential polynomial in the entries of q.
The matrix A is generated by the following simple routine, and the argu-
ments go through without change even in the d d case, where
0 1 0 1
0 q12 : : : q1n
BB 1 2 CC Bq21 0 : : : q2nCC
J =B@ ... CA ; q(x) = B
B@ ... CA :
n qn1 : : : 0
We look for wave functions satisfying Dx = 0 of the form = mexzJ .
It is easily seen that m satises the dierential equations
mx = z[J; m] + qm (4.2.7)
The functions m(x; z) can be normalized so that they tend to the identity
matrix I as x ! 1; and in that case they tend also to I as z ! 1. The
functions m are called the reduced wave functions.
Just as the Schrodinger operator generates the hierarchy of isospectral
ows known as the KdV hierarchy, the operator Dx generates a hierarchy of
deformations in q, i.e.
ows, for which the connection in (4.2.6) is
at. The
argument runs as follows:
4.2. THE AKNS HIERARCHY 55
Let be a constant matrix and set F = mm 1 : Then
Fx =mxm 1 mm 1 mxm 1 = [mxm 1 ; F ]
=[zJ zmJm 1 + q; F ]
=[zJ + q; F ]:
If q 2 S (R ) then it can be shown that F has an asymptotic expansion valid
as z ! 1 along any ray in the complex plane.
X
1
F Fj z j ; F0 = :
j =0
Substituting this asymptotic expansion into the equation for F we obtain
a set of recursion relations, analogous to the Lenard relation for the KdV
hierarchy,
[J; F0 ] = 0; d q; F ] = @Fj [q; F ]:
[J; Fj+1] = [ dx (4.2.8)
j j
@x
Since F0 = and [J; ] = 0, must be a diagonal matrix. With no loss of
generality, we may work in the class of traceless matrices, hence we take
to be a diagonal matrix of trace 0.
Now dene
@ (z n F ) ; X
n
Dt = @t + (z n F ) += Fj z n j :
j =0
In particular, we see that the t component of the connection is determined
by the x component; in fact the function F which satises Fx = [zJ + q; F ]
is a generating function for the entries of Dt. It is a simple consequence of
the recursion relations (4.2.8) that:
d q; F ] = q [J; F ]
[Dx; Dt] = qt [ dx n t n+1
Choosing 0 b
F2 = a3 + c 0
we nd 0 b 0 1
[3 ; F2] = 2 c 0 = ux 1 0 ;
4.2. THE AKNS HIERARCHY 57
hence
b = c = u2x :
The coecient a is determined at the next stage.
The equation for F3 is
@F 1
0 1
[3 ; F3 ] = 2
@x [q; F2] = ax3 + 2 uxx 1 0
1 0 1
=(ax + uux)3 + 2 uxx + 2au 1 0 :
Since the diagonal entries of [3 ; F3 ] vanish, the coecient of 3 on the right
side of the above equation must vanish as well, hence we obtain
2
ax + uux = 0;a = u2 :
Note that the expression for ax is an exact derivative, so that a is local in u,
as is assured by Theorem (4.2.1); this fact holds at all orders of the recursion
procedure, even in the d d case.
We thus have
u 2 1
0 1
F2 = 2 3 + 2 ux 1 0 :
By the same argument we nd
1 1 u3
0 1
F3 = a3 3 + 4 uxx 1 0 ;2
where a3 is determined in the computation of F4. Namely, we nd
@F @
1 1
0 1
3 3
[3 ; F4] = @x [q; F3] = a3;x3 + 2a3 q + @x 4 uxx 2 u 1 0 :
j =1 ia (kj )
0 x
Z1
= f2 (x + s) f2(s + t)G(x; t)dt;
x
where
X
N
f2(s) = ia0c(jk ) eik s: j
j
j =1
Putting these together, we get (5.1.12), with f = f1 + f2:
To establish (5.1.14) we apply D2 + k2 + 61 u to (5.1.11) to get
0= xx + (k
2 + 1 u)
6
Z 1
= Gxx(x; s) + (k2 + 1 u(x))G(x; s)
6 eiksds
x
d G(x; x) + ikG(x; x) 1 u(x)):
eikx(Gx(x; x) + dx 6
68 CHAPTER 5. INVERSE SCATTERING THEORY
Now
Z1 Z 1 @2eiks
k2eiksG(x; s)ds = @s2 G(Zx; s)ds
x x
1
=ik G(x; x)eikx Gs(x; x)eikx Gss(x; s)eiksds:
x
Combining these two calculations and multiplying by e ikx we get
Z1
(Gxx Gss uG)(x; s)eik(s x)ds (2 d G(x; x) 1 u(x)) = 0:
dx 6
x
This holds for = k 0: Letting k tend to innity in the upper half plane, we
see that the integral tends to zero. Since the other term does not depend on k
it must vanish identically. But then the integral must also vanish identically;
and, by the uniqueness of the Fourier transform,
..k the integrand must also
vanish identically, thus proving the result. `
The derivation of (5.1.15) is carried out as follows. We know jaj2 from r and
the relation jtj2 + jrj2 = 1, and we know the location of the zeroes of a in the
upper half plane since we are given the kj . If we dene ea by
Y
N
k kj
a(k) = ea(k)
j =1 k + kj
5.1. THE SCHRODINGER EQUATION 69
then ea is analytic in the upper half plane, tends to 1 as k tends to innity,
and jeaj = jaj on the real axis. Therefore log ea is analytic in the upper half
k plane and tends to zero at innity.
The function A dened by
(
A(k) = log ea(k); = k > 0;
log ea(k) = k < 0:
is sectionally holomorphic in = k 6= 0 and tends to zero as k tends to innity.
Its jump across the real axis is
A(k + i0) A(k i0) = [A] = log ea(k) + log ea(k) = log jeaj2 = log jaj2:
These properties uniquely determine A. We claim that
1 Z 1 log jaj2
A(k) = 2i dt; = k 6= 0
1 t k
In fact, the expression on the right is sectionally holomorphic and tends to
zero as k tends to innity. By the Plemelj formulae,
1
Z 1 log jaj2 1 P Z 1 log jaj2
lim 2
dt = 2 log jaj + 2i dt;
z!ki0 2i 1 t z 1 t k
where P denotes the Cauchy principal value. Hence the sectionally holo-
morphic function dened by the Cauchy integral has the same jump across
the real axis as A, and both functions tend to zero as k ! 1. Hence they
coincide.
The next step is to solve the GLM equation using the Fredholm theory
of integral equations. We x x and consider (5.1.12) as a Fredholm integral
equation on the half line (x; 1). It involves the integral operator dened by
Z1
(FxK )(s) = f (s + t)K (t)dt; s x:
x
We rst observe that if f (s) decays suciently rapidly as s tends to +1,
then Fx is a Hilbert-Schmidt operator, i.e.
Z 1Z 1
jf (s + t)j2dsdt < +1:
x x
70 CHAPTER 5. INVERSE SCATTERING THEORY
In fact, the simple change of variables u = s + t; v = s shows that this
integral is equal to
Z1
(u 2x)jf (u)j2du (5.1.16)
2x
so Fx is a Hilbert-Schmidt operator if juj1=2f is in L2 (2x; 1).
Now look at (5.1.13). The discrete sum decays exponentially as s ! +1.
The integral term decays if r is dierentiable. (This is a standard fact about
Fourier transforms). For example, if the second derivative of r is in L1 then
f decays like s 2 . In fact, if u 2 S then so does r and hence so is f [6].
Therefore, for each x the GLM equation is a Fredholm integral equation.
The Fredholm alternative states that the integral operator I + Fx has a
bounded inverse if and only if the homogeneous equation (I + Fx)G = 0 has
only the trivial solution. Therefore, to prove the existence of a solution of
(5.1.12) it is enough to prove that ker (I + Fx) is trivial.
Suppose that
Z1
G(s) + f (s + t)G(t)dt = 0; s x:
x
We may extend G to be zero for s < x: Then, since G is real,
Z1 Z 1Z 1
jG(s)j2ds + f (s + t)G(t)G(s)dsdt = 0:
x x x
But f = f1 + f2 , where
X
N
!j s ;
f2 = dj e dj > 0; kj = i!j
j =1
Therefore
Z 1Z 1
f2(s + t)G(t)G(s)dsdt
x x
X Z 1Z 1
N
!j t e !j s G(t)G(s)dsdt
= dj e
j =1 x x
X Z 1 ! t
N 2
= dj e G(t) dt 0 j
j =1 x
5.1. THE SCHRODINGER EQUATION 71
Similarly,
Z 1Z 1
f1 (s + t)G(t)G(s)dsdt
x Zx Z
1 1 1 Z1
=
2 1
r(k)eik(s+t) G(t)G(s)dkdsdt
Zx 1 x
= r(k)(Ge(k))2 dk
1
where 1 Z
Ge(k) = 21 G(t)eikt dt:
x
This latter quantity is real, since both f and G are real; hence
Z1 Z1
0 = ((I + Fx )G; G) jGej2 + r(k)Gf2 dk jGej2(1 jr(k)j)dk;
1
1
where we have applied the Plancherel theorem to the Fourier transform for
G: Z1 Z1
jGj2 dk = jGej2 dk:
1 1
Since jtj2 + jrj2 = 1, jr(k)j = 1 implies t = 0: In that case a = t 1 is innite;
and this can happen only at k = 0. Therefore Ge and hence G must vanish
identically.
We have proved:
Theorem 5.1.6 The Gel'fand-Levitan-Marchenko equation is uniquely solv-
able whenever the integral in (5.1.16) is nite.
By the Fredholm alternative, the Gel'fand-Levitan-Marchenko equation
is solvable whenever r 2 L2
The GLM equation can be solved explicitly in the so-called re
ectionless
potential case where r(k) 0; the choice 2
X
N
!j s ;
f (s) = dj e dj > 0; kj = i!j
j =1
2 Note that dj > 0 by Lemma 5.1.2 and (5.1.13).
72 CHAPTER 5. INVERSE SCATTERING THEORY
leads to the multisoliton solutions. We look for a solution of the GLM equa-
tion of the form
XN
G(x; s) = gj (x)e !j s:
j =1
This ansatz leads to an algebraic system of equations for the gj :
X
N (!j +!k )x
dj e !j x + Djk gk (x) = 0; Djk = jk + dj d!k e + ! :
k=1 j k
We may solve this equation in closed form as follows. Let E denote
the column vector with entries dj e !j x and let Ck denote the kth column
vector of jjDjkjj. By Cramer's rule, gk (x) = Dk (x)D 1 (x), where D(x) =
det jjC1; : : : ; Cnjj and Dk = det jjC1; : : : ; E; : : : ; Cnjj. Hence
X
N
Dj (x)e !j s
G(x; s) = D :
j =1
Now note that Ck0 = Ee !k x, so that
d D = det jjC 0 ; : : : ; C jj + + det jjC ; : : : ; C 0 jj
dx 1 n 1 n
X
N
!j x
= Dj (x)e = G(x; x)D;
j =1
or
d log D:
G(x; x) = dx (5.1.17)
From the second equation in (5.1.14),
d2 d d e (!j +!k)x
u(x) = 12 dx2 log det jk + j !k + ! (5.1.18)
j k
To get the multi-soliton solution of the KdV equation, replace dj in (2.2.5)
by
dj e8i(i!j )3 t = dj e8!j3 t:
This formula for the multisoliton potentials can be extended to the general
solution of the GLM equation by the method of Fredholm determinants [54],
[60].
5.2. FIRST ORDER SYSTEMS 73
5.2 First Order Systems
We now turn to the study of the forward and inverse scattering theory asso-
ciated with the rst order the dierential operator
d
i 0 0 p
Dx = dx zJ q; J = 0 i q= r 0 (5.2.19)
The scattering problem for Dx was rst proposed and solved by Zakharov and
Shabat [76] in their study of the nonlinear Schrodinger equation. Ablowitz
et. al. [1] showed that this operator plays the role of an isospectral operator
for the hierarchy of integrable systems discussed in x (4.2), now known as
AKNS systems. Under the symmetry reduction r = p it is the isospectral
operator for the nonlinear Schrodinger equation; with r = p one gets the
hierarchy containing the sine-Gordon equation as well as the modied KdV
equation; and with r = p one gets the sinh-Gordon equation.
The inverse scattering problem for 2 2 systems can be formulated as
an integral equation of Gel'fand-Levitan-Marchenko type, just as for the
Schrodinger operator. This method does not extend to the case n > 2; rather,
higher order inverse scattering problems must be formulated as Riemann-
Hilbert problems [5], which are considerably more complicated technically.
We shall develop the scattering theory of (5.2.19) and obtain the Gel'fand-
Levitan-Marchenko integral equation. We begin by constructing the wave
functions of Dx.
Theorem 5.2.1 There exist fundamental matrix valued solutions of Dx =
0 of the form = mexzJ where:
1. m0 = z [J; m] + q;
2. m is meromorphic in Im z > 0;
3. For regular values of z , m are bounded for all 1 < x < 1;
4. m ! I as x ! 1 in Im z > 0; and m ! (z ) as x ! +1 in
= z > 0, where are diagonal matrices.
5. m ! I as z ! 1 in Imz > 0. If q 2 C n then m each have an
asymptotic expansion
X n
m mzjj ; m0 = I; [J; mj+1] = dm dx
j
qmj ;
j =0
74 CHAPTER 5. INVERSE SCATTERING THEORY
the mj are uniquely determined. If q 2 S then the asymptotic expansion
is valid to all orders.
6. m are uniquely determined by properties 1,3,4.
7. det m = 1.
Proof. Let 1 0
e+ = 0 e = 1 ):
The wave functions are obtained by converting the dierential equation to a
Volterra integral equation. For example, by writing = `e ixz we obtain the
dierential equation
d` = zE` + q`; E = 2i 0 :
dx 0 0 (5.2.20)
This may be converted to a Volterra integral equation, for example
Zx
`=e + ezE(x y)q`dy: (5.2.21)
1
We leave it to the reader to verify that if q 2 L1(R ), then the successive
approximations to (5.2.21) converge uniformly in = z 0. The solution ` is
analytic in the upper half plane, is bounded for all real x, and tends to e
as x tends to 1.
If q 2 C n then ` has an asymptotic expansion
X
n
` `j z j (5.2.22)
j =0
where
`0 = e ; E`1 = qe ; E`j+1 = `0j q`j :
To validate (5.2.22), let
X
n
j
Rn = ` `j z
j =0
and substitute this into (5.2.20). We nd
dRn = (zE + q)R E`n+1 :
n
dx zn
5.2. FIRST ORDER SYSTEMS 75
Since Rn tends to zero as x ! 1, Rn satises the integral equation
Zx 1
Zx
Rn = e zE (x y ) qRndy zn ezE(x y) E`n+1dy
1 1
The solution of this integral equation obeys the estimate
jRn(x)j C (x)jzj n
for some constant C which depends on x. This proves the validity of (5.2.22).
If q 2 C 1 then we obtain the innite asymptotic expansion
X1
` z`jj :
j =0
The existence of wave functions and , analytic in = z > 0 with
the asymptotic behavior
eixz e; x ! 1; e e ; x ! +1
ixz
The limits of the other reduced wave functions are evaluated in the same
manner. The reduced wave functions m are dened by
jj d+ ; +jj = m+ exzJ ; jj; a jj = m exzJ :
It is a straightforward exercise to check that properties (1)-(5) of the reduced
wave functions given in Theorem (5.2.1) are satised with this choice of m.
The poles of m+ ; m are located at the zeroes of d and a respectively.
Finally, we prove that items 1,3,4 uniquely determine the reduced wave
functions. Suppose m1 and m2 satisfy the dierential equation
dm = z[J; m] + qm:
dx
for z, say, in the upper half plane, are bounded in x, and tend to the identity
as x tends to 1. Then w = m1 1 m2 satises
dw = z[J; w]; w ! I; x ! 1:
dx
5.2. FIRST ORDER SYSTEMS 77
The general solution of the dierential equation is w = exzJ w0(z)e xzJ ; for
some matrix valued function w0(z) indepdendent of x. Fix z. Since m1
and m2 are bounded for all x 2 R , and det m1 = 1, so is m1 1 m2. This
implies that w0 is a diagonal matrix, hence commutes with exzJ . Therefore
w(x; z) = w0(z). Since w tends to the identity as x ! 1, we must have
w = I . `..k
0
We may combine all these equations and write them in a simple matrix
form. Let
f (s) 0 0 1
+
G = jjG+; G jj; f (s) = 0 f (s ; = 1 0 ):
)
Then
Z1
(x; k) = exkJ + G(x; y)eykJ dy; (5.2.25)
x
5.2. FIRST ORDER SYSTEMS 79
and the integral equations can be written as
Z1
0 = f (s + x) + G(x; s) + G(x; y)f (y + s)dy; s > x: (5.2.26)
x
Substituting (5.2.25) into the equation Dx (x; k) = 0 and proceeding as
in the proof of Theorem (5.1.5) we get
0 = (q(x) + G(x; x) JG(x; x)J 1 )
Z1
+ q(x)G(x; y) @G @x J @G J 1 e(y x)kJ dy:
@y
x
Since the rst term is independent of k, we may let k tend to innity. Then
the integral term goes to zero by the Riemann-Lebesgue lemma, assuming
the integrand is in L1 (x; 1). Hence the rst term must vanish, and
q(x)J = [J; G(x; x)]: (5.2.27)
Equation (5.2.27) is the analog of (5.1.14) in the theory of the Schrodinger
equation. From this equation we nd
p(x) = 2B (x; x); r(x) = 2C (x; x): (5.2.28)
These formulae allow us to recover the components of the potential from the
kernel of the GLM equation.
The integral term must also vanish, and it may be shown that the kernel
G satises a hyperbolic system of partial dierential equations.
Writing A B
G= C D ;
we may break (5.2.26) down into four equations:
Z1
f+(x + s) + B (x; s) + A(x; y)f+(y + s)dy =0;
Z1
x
A(x; s) + B (x; y)f (y + s)dy =0;
Zx1
f (x + s) + C (x; s) + D(x; y)f (y + s)dy =0;
Z1
x
D(x; s) + C (x; y)f+(y + s)dy =0:
x
80 CHAPTER 5. INVERSE SCATTERING THEORY
These give
Z1 Z1
f+(x + s) + B (x; s) B (x; y) f (y + z)f+(z + s)dzdy = 0
x x
(5.2.29)
and
Z1
A(x; s) f+(x + y)f+(y + s)dy (5.2.30)
x Z1 Z1
A(x; y) f+(y + z)f (z + s)dzdy = 0: (5.2.31)
x x
Under the assumption that
Z1
(u x)jf(u)j2du < +1
x
both the integral operators dened on L2(x; 1) by
Z1
(F
x )(s) = f(s + y) (y)dy
x
are Hilbert Schmidt operators; therefore so are the compositions Fx+Fx and
Fx Fx+: Hence the integral equations (5.2.26) are Fredholm integral equations.
Their solvability is therefore a consequence of uniqueness; but uniqueness can
only be shown to hold in cases where there is a symmetry of the potential
matrix q, hence of the scattering data.
Lemma 5.2.2 Let q = q and let
0 1
= 1 0 :
Then
+(x; z) = (x; z); +(x; z) = (x; z);
S is unitary: SS = I :
a b
S= b a :
5.2. FIRST ORDER SYSTEMS 81
G(x; s) 1 = G(x; s); hence
A B
G = B A ):
kj+ = kj ; c+j = cj ;
f+ (s) = f (s).
Lemma 5.2.3 Let q be real and skew symmetric. Then in addition a(k) =
a( k); b(k) = b( k) for real k;
A B
G(x; s) = B A );
where A and B are real; and f (s) are real.
Proof. We know that + (x; z ) e+ eixz as x ! +1 or as z ! 1 in the
upper half plane. Hence, for = z < 0,
+ (x; z) e+ e ixz = e e ixz :
On the other hand, a short calculation shows that Dx(z; q) 1 = Dx(z; q),
so +(x; z) is also a solution of the dierential equation. Since the the wave
functions are uniquely determined by their asymptotic behavior, we obtain
the rst statement of Lemma (5.2.2). The others are derived in exactly the
same way.
From the rst statement we nd that
(x; k) 1 = (x; k); (x; k) 1 = (x; k);
hence S (k) 1 = S (k): Since det S = 1 this implies that S 2 SU (2), hence
a = d and c = b.
We leave the proof of the remaining statements of Lemma (5.2.2) to the
reader. In the case of Lemma (5.2.3) we have the additional symmetry that
q is real. Thus (x; k) 1 = (x; k), etc., and
S (k) 1 = S ( k); a(k) = a( k); b(k) = b( k):
Moreover, we now nd that the entries A and B in G are now real. The
symmetry of a can now be written a(k) = a( k); and from this we see that
82 CHAPTER 5. INVERSE SCATTERING THEORY
if kj+ is a zero of a then so is kj+ . Call this k+j . From the symmetry in
Lemma (5.2.2) we also know that if k is a zero of a then k is a zero of a.
Thus the bound states appear in foursomes: k; k, or in pairs if k is purely
imaginary. From the symmetries of the wave functions we nd that c+j = cj .
From this one may show that f+(s) = f (s) is real. ` ..k
Theorem 5.2.4 In the case of the symmetries discussed above we have [53]
d2 log det jjI + F F jj;
ju(x)j2 =2 q = q;
x x
dx2
d arg det jjI + iF jj;
u(x) = 2 dx qy = q; q real:
x
where r is the re
ection coecient. If we group + and +=a on the left and
and = a on the right side, we can write these equations in the form
+
1 r
1 0
+ a
0 1 =
a
r 1
5.3. DECAY OF THE SCATTERING DATA 83
Since a and r are independent of k we may also dierentiate these equa-
tions with respect to x. Combining them with the relations above, and noting
that 1 01 r 1 1 r 1 r
r 1 0 1 = r 1 jrj2 = r jtj2
we obtain
0 1
+=a
=
=a
@ 1 r
A: (5.3.32)
+
r jtj2
These conditions form multiplicative jump conditions across the real axis and
consitute in part a Riemann Hilbert problem for the wave functions.
The reduced wave functions m are dened by
0 1 0 1
+=a
@ +
A = m+eixz3 ; @ =a A = m eixz3 :
0 0+=a 0 =a 0
+
where 1
3 = 0 01 :
The sectionally meromorphic function dened by m = m in = z > 0
satises the jump conditions
0 1
1 r
m+ = m eixk3 @ Ae ixk3
2r jtj
across the real axis. At x = 0 we get
0 1
1 r
@ A = m (0; k) 1m+(0; k): (5.3.33)
2
r jtj
The column vectors of m are meromorphic in the upper and lower half
z planes respectively and may be obtained as solutions of Volterra integral
equations by the method of successive approximations. On the real axis, they
are C 1 functions of k, provided the potential u 2 S . This may be seen by
84 CHAPTER 5. INVERSE SCATTERING THEORY
dierentiating the Volterra integral equation with respect to k and solving
the respective equations for the derivatives.
Furthermore, the asymptotic expansions of m in k are valid down to the
real axis, and may be dierentiated termwise with respect to k. Moreover,
the asymptotic expansions for m are uniquely determined by u; and in
particular, m+ and m have the same asymptotic expansions on the real k
axis. Therefore
m (0; k) 1m+(0; k) I
and 0 1
1 r
@ A I 2 S:
r jtj2
This shows not only that jrj ! 0 as jkj ! 1, but also that jaj2 ! 1 as
k ! 1 on the real axis. `..k
Chapter 6
Variational Methods
6.1 A variational principle for water waves
Variational methods for nonlinear waves occur in a number of contexts, and
I will describe some of these methods in this lecture. There is also an exten-
sive theory for free boundary value problems in potential
uid
ow [13]{that
is, the
ow of an irrotational, inviscid
uid. These problems may be formu-
lated as variational problems [32] for the stream function and have received
considerable attention in recent years [3, 31].
I will begin by discussing the variational principle for the time dependent
Euler equations of water wave theory. This principle was rst formulated
in 1967 by J. Luke [49]; in 1968 a Hamiltonian formulation of the Euler
equations was given by V. E. Zakharov [77]. Benjamin and Olver [10, 56] have
determined all local conservation laws of the Euler equations. This principle
applies to the potential function and so can be used in three dimensions,
whereas variational formulations for the stream function are restricted to
two dimensional
ows, or axisymmetric three dimensional
ows.
The possibility of proving proving the existence of solutions of the Euler
equations by the direct method of the calculus of variations should be con-
sidered; to my knowledge this has yet been done. The variational principle
might also allow for a weak formulation of the free boundary value problem,
which might be suitable in discussing weak solutions of the Euler equations.
This would be particularly convenient in the discussion of breaking waves,
etc.
The variational principle proposed by Luke is the following. We introduce
85
86 CHAPTER 6. VARIATIONAL METHODS
the local Lagrangian
Z
L(x; t) = 't + 12 (r')2 + gy dy
1
Z 2
= 't + 21 (r')2 dy + g2 + const:
1
We drop the constant and form the action
Z t2 Z1
S= L dx dt; t1 < t2 :
t1
1
The Euler equations of x3.3 are the Euler-Lagrange equations for this
integral; that is, that they are a consequence of the variational principle
S_ = 0 where
S_ = S
[
_ ] + S ['_ ]:
'
We have
S [_ ] = Z t2 Z L [_ ] dxdt;
1
t1
1
where
L = @ L = ' + 1 (r')2
@ t
2 y= + g:
Thus
Z t2 Z1
't + 21 (r')2 + g y= _ dxdt = 0
t1
1
for all variations _ implies the equation
't + 12 (r')2 + g = 0 on y = :
Now let us calculate the variation of S with respect to variations in '.
S ['_ ] = Z t2 ZZ '_ + r' r'_ dxdydt;
We have
t
' t1 D
6.1. WATER WAVES 87
where D is the region occupied by the
uid.
Since div('_ r') = r'_ r' + '_ ', we have, by Green's theorem,
ZZ ZZ I
r' r'_ dxdy = '_ ' dxdy + '_ ('xdy 'y dx)
@D
D D
ZZ Z1 Z1
= '_ ' dxdy '_ y dx '_ ('xx 'y )y= dx:
D 1 1
On the other hand,
2 3
Z t2 ZZ Z t2 6 @ ZZ Z1 7
'_ tdxdydt = 4 @t '_ dxdy t'_ y= dx5 dt
t1 t 1
D D 1
ZZ t2 Z1
= '_ dxdyt t'_ y= dxdt:
1
D 1
Hence
S ['_ ] = Z t2 ZZ Z t2 Z1
''_ dxdydt (t + x'x 'y ) '_ dxdt
' t1 t1 y=
D 1
ZZ t2 Z t2 Z1
+ '_ dxdyt 'y y= h'_ dxdt:
1 t 1
D 1
We now assume that the variation S=' vanishes for all variations '_ and
prove that the remainder of the Euler equations follow. First, we restrict the
variations to those for which '_ (x; y; t1) = '_ (x; y; t2) = 0. Now assume the
variation '_ vanishes on @D . Then we see that ' = 0 in D . Next assume
the variation '_ vanishes on the free surface. We then obtain the boundary
condition 'y = 0: Finally, let '_ be an arbitrary continuous function;
y= h y=
then we nd that
(t + x'x 'y ) = 0:
y=
88 CHAPTER 6. VARIATIONAL METHODS
A Hamiltonian formulation of the Euler equations was given by Zakharov
[77] in 1968. The Hamiltonian is
Z 1 Z
1
H=2 g2 + (r')2 dy dx
h
1
0.8
0.6
0.4
0.2
−0.2
−0.4
−0.6
−0.8
−1
−10 −8 −6 −4 −2 0 2 4 6 8 10
2
3
Thus, the full Euler equations exhibit singularities, and, moreover, there is a
limiting amplitude to the periodic wave trains, or to the solitary wave. This
shows that the KdV approximation can only be valid for suciently small
amplitudes. The KdV equation, on the other hand, has smooth cnoidal and
solitary waves of arbitrary amplitude.
G.B. Whitham [74], argued
It was remarked earlier that the nonlinear shallow water equations which
neglect dispersion altogether lead to breaking of the typical hyperbolic kind,
with the development of a vertical slope and a multivalued prole. It seems
clear that the third derivative term in the Korteweg-deVries equation will
prevent this ever happening in its solutions. But in both cases, the long
wave assumption under which the equations were derived is no longer valid.
Since some waves appear to break in this way, if the depth is small enough,
one concludes that some dispersion is necessary but the Korteweg-deVries
term is too strong for short wavelengths. This is not surprising in view of the
fact that the k3 term [is] a bad approximation to the full dispersion relation
when (kh0 )2 becomes large...
On the other hand the dispersion included in the Korteweg-deVries equa-
tion does allow the solitary and periodic waves which are not found in the
7.1. BREAKING AND PEAKING 97
shallow water theory. For these solutions, however, the Koretweg-deVries
equation cannot describe the observed symmetrical "peaking" of the crests
with a nite angle there.
0 0 0
0 2 4 6 0 2 4 6 0 2 4 6
0 0 0
0 2 4 6 0 2 4 6 0 2 4 6
This is the Harry Dym equation with an additional transport term; we get
the Harry Dym equation itself if we take q = 0.
2. The general
ow for n = 2: After some computation we nd
d2 log pm :
a2 = p1m ; a1 = 2m13=2 q + 12 dx 2
This simplies to
(1 14 D2)ut 23 (u2)x + 38 (ux)2 + 41 (uuxx)x = 0:
Finally, we note that, as in the case of the KdV theory, the squared
eigenfunctions satisfy the third order dierential equation
( R3 k 2 R1 ) = 0;
where = 1(x; k) 2 (x; k), and j satisfy
( D 2 + k 2 m q ) = 0:
Dt
@
= @t
Xn
Fn jz
j
:
j =0
7.3. THE EXTENDED AKNS HIERARCHY 107
The zero curvature conditions [Dx; Dt ] = 0 give
R0 Fn =0 (7.3.19)
R0 Fj 1 =R1 Fj ; nj1 (7.3.20)
mt =R1 F1 R0 F0 (7.3.21)
qt =R1 F0 (7.3.22)
where
R0 = ad m; R1 = @x @ ad q:
As in the extended KdV hierarchy, F0 is in general undetermined.
The AKNS hierarchy itself is obtained by taking m to be a constant
diagonal matrix J of trace 0. Then (7.3.19) implies that [J; Fn] = 0; and we
may take Fn = , where is any diagonal matrix of trace 0. The recursion
relation (7.3.20) then reads
[J; Fj 1] = @Fj [q; Fj ]
@x
Since mt = 0, F0 is determined by (7.3.21). The nonlinear
ow is then
qt = [J; F0 ]:
The F0 are dierential polynomials in the entries of q [67], hence the
ows
are local, nonlinear partial dierential equations.
Let us consider the case qt = 0 for n = 2; d = 2; this corresponds to the
Harry Dym equation in the extended KdV hierarchy. We take m = 3 , that
is, a scalar multiple of 1 0
3 = 0 1 ;
and q to be a constant vector, to be determined below. Then qt = 0 implies
that
R1F0 = @F 0
@x [q; F0] = 0:
We take F0 = q. The equation [m; F2 ] = 0 implies F2 is diagonal, and since
we are working in the class of traceless matrices, we take F2 = f2 3 .
108 CHAPTER 7. WEAK AND STRONG NONLINEARITIES
The recursion relation is then
[3 ; F1] = @f 2
@x 3 f2[q; 3 ]:
The diagonal part of both commutators vanish, hence we must have
@f2 = 0;
@x
and
[3 ; F1 f2q] = 0:
This last equation implies that
F1 f2q = n3
for some function n. Solving for F1 we obtain
F1 = f2 q + n 3 :
The
ow is then
f n
t 3 D m2 q 3 + [3 ; q] + n [q; 3 ] = 0
From the diagonal part of this equation we obtain
t = D n
and the o-diagonal part gives
n
[3 ; q] = f2 D 1 q:
Hence q must be an eigenvector of ad 3 , i.e. q = . We then have
n = D1;
and
1
t = x : (7.3.23)
xx
7.3. THE EXTENDED AKNS HIERARCHY 109
We now turn to the negative
ows. For the
ow at n = 1 we get the
equations
R1F1 =0 (7.3.24)
qt + R0F1 =R1F0 ; (7.3.25)
mt + R0F0 =0 (7.3.26)
We rst consider the special case qt = 0. In that case the rst equation
is solved by taking F1 = q, and the second equation reduces to
[q; m F0 ] = @F @x
0: (7.3.27)
The third equation cannot be solved if we take m = f3, so we take
m = a0 3 + a++ + a ; F0 = f03 + f++ + f :
We do look for a symmetry reduction and choose all matrices to lie in
su(2), the real subalgebra of s`(2; C ) spanned by the skew adjoint matrices
i 0 0 i 0 1
0 = 0 i ; 1 = i 0 ; 2 = 1 0 :
Specically, we take
1
i a f
q = 2 0 ; m = a i ; F0 = f 0
0
so that the form of the isospectral operator is
@
i a 1 1 0
Dx = @x z a i i 2 0 1
Then (7.3.27) leads to the equations
f0;x = 0 a = i(D i)
We take f0 = 0. After some computations the
ow equations (7.3.26) lead to
t + Dj j2 = 0; (D i) t = 2 : (7.3.28)
110 CHAPTER 7. WEAK AND STRONG NONLINEARITIES
Chapter 8
Numerical Methods
In order to test the validity of various model equations for nonlinear disper-
sive systems, it is extremely handy to have some numerical codes to solve
dispersive systems of partial dierential equations and compare them with
the various approximations proposed. To this end, Yi Li1 and I have writ-
ten some simple codes to be run in Matlab. These solve nonlinear evolution
equations in one space variable and then animate the solutions.
The codes are based on implicit spectral schemes. The use of spectral
methods for the numerical analysis of dispersive equations has been studied
extensively, [30], [75]. Spectral, or pseudo-spectral, methods, are based on
the use of the fast Fourier transform. In this chapter I will describe the fast
Fourier transform, and explain, by way of a simple numerical experiment,
why the use of spectral methods is preferable to the use of nite dierence
schemes for this class of problems. The implicit schemes to be used are
unconditionally stable and employ the Crank-Nicolson, or trapezoid, method
for the time-step integration. Since the nonlinear terms play a predominant
role in the calculations, we shall describe nonlinear time step solvers.
111
112 CHAPTER 8. NUMERICAL METHODS
trigonometric polynomial
X
N 1
u(x) qk eikx (8.1.1)
k=0
and interpolate at the points jh. Thus, we require that
X
N 1
uj = qk !jk; j = 0; : : : ; N 1 ! = e2i=N ;
k=0
where uj = u(jh):
The mapping q 7! u is described by a simple matrix multiplication
u =
q;
=
!jk
; j; k = 0; : : : ; N 1:
A direct computation shows that
= NI ;
hence
1 = 1
;
N
and NX1
1
qj = N uk ! kj :
k=0
The mapping from the vector u = (u1; 0tsuN ) is called the nite Fourier
transform.
A second way to interpret the nite Fourier transform is the following.
Let u be a 2-periodic function on the line, with uj = u(jh) for all j 2 Z.
Dene the centered rst order dierence operator
Duj = uj+1 2h uj 1 :
The operator D is skew symmetric. Its eigenvalues and eigenvectors are
easily found. For xed k, let k;j = !jk. Then
! (j +1)k ! (j 1)k !k ! k
Dk;j = =! jk = k k;j
2h 2h
8.1. THE FINITE FOURIER TRANSFORM 113
where
sin 2Nk
k = i 2 :
N
Thus, the eigenvectors of D are precisely the column vectors of the -
nite Fourier transform. Since D is skew symmetric, the eigenvectors of D
are orthogonal, and the nite Fourier transform diagonalizes D. This is the
discrete analog of the fact that the Fourier series diagonalizes the skew sym-
metric operator d=dx.
The implementation of the nite Fourier transform requires matrix mul-
tiplication by
. This computation is of order N 2 . On the interval [0; 2] a
choice of N = 256 gives a spatial resolution h = :0245, while 2562 = 65536.
In a numerical integration of a system of partial dierential equations, N 2
oating point operations, or '
ops', must be performed at each time step.
Spectral methods in numerical analysis are based on the discovery, in
1965, by Cooley and Tukey [24], that when N is a power of 2 the nite
Fourier transform can be implemented in N log N steps. For N = 256 = 28
this is roughly 8 256 = 2048. This is a very decided advantage. Most nu-
merical packages, including Matlab, have a nite Fourier transform package.
In Matlab, the fast Fourier transform and its inverse are implemented by the
commands 't' and 'it'.
To compare spectral methods with nite dierence schemes, consider the
numerical integration of the partial dierential equation
ut + ux = 0; u(x; 0) = f (x): (8.1.2)
We partition both space and time into nite intervals, with the spatial step
denoted by h = x and the time step denoted by t.
First we construct a nite dierence scheme. For reasons we shall explain
later, we are going to use an implicit method, which dates back to Euler. We
rst replace dierentiation in x by the centered dierence operator D dened
above. Then write the equation in the form
ut = Du
Now let's integrate this equation from t to t + t. We obtain
Z t+t
u(x; t + t) u(x; t) = Du dt:
t
114 CHAPTER 8. NUMERICAL METHODS
Now we approximate the integral on the right by the trapezoid rule to obtain
u(x; t + t) u(x; t) = Du(x; t + 2t) + Du(x; t) t:
This equation may be written in the form
(I + 21 tD)u(x; t + t) = (I 21 tD)u(x; t);
or symbolically, by
I 12 tD
u(x; t + t) = Uu(x; t); U =
I + 21 tD
There is no problem in inverting I + 21 tD since D is skew adjoint and its
eigenvalues are imaginary. The operator U is unitary2, and its L2 norm is 1,
regardless of the size of t.
Now let us implement this scheme, rst as a nite dierence scheme, and
then as a spectral scheme using the fast Fourier transform. The matrix for
D on the space of periodic functions is the tridiagonal matrix
00 1 0 0 ::: 1
1
BB 1 0 1 0 ::: 0CC
B0 0C
D = 21h B C
1 0 1 :::
BB ... CC
B@ 0 0 1A
C
1 0 1 0
In order to speed the computations and use less memory, we represent this
matrix as a sparse matrix.
Here is the Matlab code:
ops(0) % set '
oating point operations' counter to 0;
N=input('N=');
T=1; L=2*pi;
h=L/N
x=(h:h:L)';
l=1; % input('l=');
2U is the Cayley transform of 1
2D ; cf. Riesz and Nagy [63]
8.1. THE FINITE FOURIER TRANSFORM 115
dt=l*h;
J=x(T/dt);
u=zeros(N,J);
u(:,1)= exp(-20*(x-.5*pi).b2);
e=ones(N,1);
D=spdiags([-e,e], -1:2:1,N,N); % sparse tridiagonal matrix.
D(1,N)=-1; D(N,1)=1; % modify for periodic functions.
id=sparse(eye(N));
A=id+.5*l*D;
B=2*id-A;
for j=2:J;
u(: , j)=An B*u(:,j-1);
end
ops
plot(x,u(:,J))
0.8
0.6 T=3
66,754,090 flops
0.2
dispersive tail
−0.2
0 1 2 3 4 5 6 7
0.9
0.8
0.7
0.6 T=3
0.3
0.2
0.1
0
0 1 2 3 4 5 6 7
uk = N1 qj !jk;
j= N 2
where it is understood that only half the contributions at N=2 are included
in the sum.
The ordering of the basis functions in the Matlab fast Fourier tranform
algorithm unfortunately corresponds to that given by the matrix
above,
and is a little harder to work with. From the centered form of the tranform it
follows easily that when u is real the Fourier coecients satisfy the familiar
symmetry
q j = qj : (8.1.3)
This implies that both q0 and qN=2 are real when N is even. This is a re
ection
of the fact that the cyclic group C N has two real representations when N is
even. This leads to some problems in using the Fourier transform for even
N.
If u is approximated by the trigonometric polynomial (8.1.1) then its
derivative is given by
X
N=2
ux ijqj eijx;
k= N=2
hence dierentiation is obtained on the transform side by multiplication by
the diagonal matrix D with entries ij , j = N=2; : : : N=2.
If u satises the partial dierential equation (8.1.2) then its Fourier trans-
form coecients qj (t) satisfy the ordinary dierential equations
q_j + ijqj (t) = 0:
118 CHAPTER 8. NUMERICAL METHODS
This
ow preserves the symmetry (8.1.3) of the Fourier coecients. In par-
ticular, q0 remains constant, corresponding to the conservation law
Z 2
u(x; t) dx = const:
0
On the other hand,
q N2 (t) = q N2 (0)ei N2 t
and cannot remain real under the
ow. Thus use of the nite Fourier trans-
form for even N leads to an error in the computation.
This error is unavoidable, but when the solutions remain smooth, it is
usually not a serious error. Just as in the case of the Fourier series for
the group S 1 , the Fourier coecients of a smooth function decrease rapidly
with j . For example, if u(x) is s times continuously dierentiable then its
Fourier coecients decrease like j s, while if u is analytic, then its Fourier
coecients decrease exponentially. The same holds for the nite Fourier
transform, so if u(x; 0) is analytic in x the quantities qj decrease exponentially
with j . In particular, qN=2 is extremely small. These facts can be conrmed
experimentally by doing a few sample calculations in Matlab.
The solution of (8.1.2) is given to a very good approximation by
0 2 1
X N
These two codes show clearly the advantage of spectral methods over
nite dierence methods. The solution of (8.1.2) using the spectral method
for T = 3 is shown in Figure 8.2. When the nite Fourier transform is used,
the choice of dierence operators is immaterial; moeover, when analyzing
higher order dierential equations, such as the KdV equation, the dierence
operators involved are no longer tri-diagonal, and the spectral method gains
more and more of an advantage as the power of the operator D goes up.
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
0 50 100 150 200 250 300
Figure 8.3: High frequency cut-o due to the dispersion acting on the non-
linear terms. t = :05; disp = :05, L = 40; N = 256.
124 CHAPTER 8. NUMERICAL METHODS
Bibliography
[1] Ablowitz, M.J., D.J. Kaup, A.C. Newell, & H. Segur. The
inverse scattering transform {Fourier analysis for nonlinear problems.
Studies in Applied Mathematics, 53:249{315, 1974.
[2] Airy, G.B. Tides and Waves. In Encyclopaedia Metropolitana, London,
1845. article 192.
[3] Alt, H.W. & L.A. Caffarelli. Existence and regularity for a mini-
mum problem with free boundary. J. Reine Angew. Math., 325:105{144,
1981.
[4] Amick, C.J., L.E. Fraenkel, and J.F. Toland. On the Stokes
conjecture for the wave of extreme form. Acta Mathematica, 148:193{
214, 1982.
[5] Beals, R. & Coifman, R. Scattering and inverse scattering for rst
order systems. Comm. Pure. Appl. Math., 37:39{90, 1984.
[6] Beals, R., P. Deift, C. Tomei. Direct and Inverse Scattering on the
Line, volume 28 of Mathematical Surveys and Monographs. American
Math. Soc., 1988.
[7] Benjamin, T.B. Internal waves of permanent form in
uids of great
depth. J. Fluid Mech., 29:559{572, 1967.
[8] Benjamin, T.B. The stability of solitary waves. Proc. Roy. Soc. Lon-
don A, 328:153{183, 1972.
[9] Benjamin, T.B. Impulse,
ow-force, and variational principles. IMA
Journal of Applied Mathematics, 32:3{68, 1984.
125
126 BIBLIOGRAPHY
[10] Benjamin, T.B., and P.J. Olver. Hamiltonian structure, symme-
tries and conservation laws for water waves. J. Fluid Mech., 125:137{185,
1982.
[11] Benjamin, T.B., Bona, J. & Mahoney, J.J. Model equations for
long waves in nonlinear dispersive systems. Phil. Trans. Royal Soc.
London A., 272:47{78, 1972.
[12] Birkhoff, G. Hydrodynamics, a Study in Logic, Fact, and Similitude.
Dover, New York, 1950.
[13] Birkhoff, G. and E.H. Zarantonello. Jets, Wakes, and Cavities.
Academic Press, New York, 1957.
[14] Bogoliubov, N.N. and Y.A. Mitropolsky. Asymptotic Methods
in the Theory of Nonlinear Oscillations. Gordon and Breach, New York,
1961.
[15] Bona, J. Solitary waves and other phenomena associated with model
equations for long waves. Fluid Dynamics Transactions, 10:77{111,
1980.
[16] Bona, J.L., W.G. Pritchard, & L.R. Scott. An evaluation of
a model equation for water waves. Phil. Trans. Royal Soc. London A.,
302:457{510, 1981.
[17] Bona, J.L., W.G. Pritchard, & L.R. Scott. A comparison of
solutions of two model equations for long waves. In Norman R. Lebovitz,
editor, Fluid Dynamics in Astrophysics and Geophysics, volume 20 of
Lectures in Applied Mathematics, pages 235{267, Providence, R.I., 1983.
Amer. Math. Soc.
[18] Boussinesq, M.J. Theorie des ondes et des remous qui se propagent
le long d'un canal rectangulaire horizontal, en communiquant au liquide
contenu dans ce canal des vitesses sensiblement pareilles de la surface
au fond. J. Math. Pure. Appl., 17:55{108, 1872.
[19] Boussinesq, M.J. Essai sur la theorie des eaux courantes. Memoires
presentes par divers savants a l'Academie des Sciences Inst. France
(series 2), 23:1{630, (1877).
BIBLIOGRAPHY 127
[20] Brillouin, L. Wave Progagation and Group Velocity. Academic Press,
New York, 1960.
[21] Camassa, R. and D.D. Holm. An integrable shallow water equation
with peaked solitons. Phys. Rev. Lett., 71:1661{1664, 1993.
[22] Camassa, R., D.D. Holm, & J.M. Hyman. A new integrable shallow
water equation. Advances in Applied Mechanics, 31:1{33, 1994.
[23] Cercignani, C. Solitons. theory and application. Revista del Nuovo
Cimento, 7:429{469, 1977.
[24] Cooley, J. W. & J.W. Tukey. An algorithm for the machine cal-
culation of complex fourier series. Math. Comp., 19:297{301, 1965.
[25] Craig, W. An existence theory for water waves and the Boussinesq and
Korteweg-deVries scaling limits. Commun. Partial Di. Eq., 10:787{
1003, 1985.
[26] Dodd, R. K., J. C. Eilbeck, J. C. Gibbon & H. C. Morris.
Solitons and Nonlinear Wave Equations. Academic Press, New York,
1982.
[27] Dorfmeister, J., & Szmigielski, J. get title. Mathematica Appli-
canda, 123:456{457, 1998.
[28] Erdelyi, A. Asymptotic Expansions. Dover, New York, 1956.
[29] Fermi, E., Pasta, J.R. and Ulam, S.M. Studies of nonlinear prob-
lems. In Collected Works of E. Fermi, vol. II, pages 978{988. University
of Chicago Press, 1965. cf. also Los Alamos Sci. Lab. Tech. Report
LA-1940.
[30] Fornberg, B. & G.B. Whitham. A numerical and theoretical study
of certain nonlinear wave phenomena. Phil. Trans. Royal Soc. London,
289:373{404, 1978.
[31] Friedman, A. Variational Principles and Free Boundary Problems.
John Wiley & Sons, New York, 1982.
[32] Friedrichs, K.O. U ber ein Minimumproblem fur Potentialstromung
mit freiem Rand. Math. Ann., 109:60{82, 1934.
128 BIBLIOGRAPHY
[33] Friedrichs, K.O. and D.H. Hyers. The existence of solitary waves.
Comm. Pure Appl. Math., 7:517{550, 1954.
[34] Fuchsteiner, B. The Lie algebra structure of nonlinear evolution
equations and innite dimensional abelian symmetry groups. Progr.
Theor. Phys., 65:861{, 1981.
[35] Gardner,C.S., J.M. Greene, M.D. Kruskal, and R.M. Miura.
Method for solving the Korteweg-deVries equation. Phys. Rev. Lett.,
19:1095{1097, 1967.
[36] Gardner, C.S. & Morikawa, G.K. Similarity in the asymptotic
behavior of collision-free plasma and water waves. Technical Report
NYO-9082, Courant Inst. Math. Sci., 1960.
[37] Goodman, J. & Lax, P. D. On dispersive dierence schemes. I.
Comm. Pure Appl. Math., 41:591{613, 1988.
[38] Haragus, M. and D.H. Sattinger. On the linearized Korteweg-
deVries equation. to appear, Zeit. Ang. Math. u. Phys.
[39] Jeffrey, A. and T. Kawahara. Asymptotic Methods in Nonlinear
Wave Theory. Pittman, London, 1982.
[40] Jeffrey, A. & T. Kawahara. Asymptotic Methods in Nonlinear
Wave Theory. Applicable Mathematics Series. Pitman, Boston, 1982.
[41] Kirrmann, P. G. Schneider, & A. Mielke. The validity of mod-
ulation equations for extended systems with cubic nonlinearities. Proc.
Royal Soc. Edinburgh, 122A:85{91, 1992.
[42] Kruskal, M. Nonlinear Wave Equations. In J. Moser, editor, Dynam-
ical Systems, Theory and Applications, volume 38 of Lecture Notes in
Physics, pages 310{354, Heidelberg, 1975. Springer.
[43] Kruskal, M.D., & Z.J. Zabusky. Interaction of \solitons" in a
collisionless plasma and the recurrence of initial states. Phys. Rev. Lett.,
15:240{243, 1965.
[44] Lamb, G.L. Jr. & D. W. McLaughlin. Aspects of soliton physics.
In R. K. Bullough and P. J. Caudrey, editors, Solitons, volume 17 of
Topics in Current Physics, pages 65{106. Springer-Verlag, 1980.
BIBLIOGRAPHY 129
[45] Lax, P.D. Integrals of nonlinear equations of evolution and solitary
waves. Comm. Pure and Applied Math., 21:467{490, 1968.
[46] Lax, P.D. Almost periodic solutions of the KdV equation. SIAM
Review, 18:351{375, 1976.
[47] Lax, P. D., Levermore, C. D., & Venakides, S. The generation
and propagation of oscillations in dispersive initial value problems and
their limiting behavior. In Important Developments in Soliton Theory,
Nonlinear Dynamics, pages 205{241, Berlin, 1993. Springer.
[48] Levi-Civita, T. Determination rigoureuse des ondes permanentes
d'ampleur nie. Math. Ann., 93:264{314, 1925.
[49] Luke, J.C. A variational principle for a
uid with a free surface. J.
Fluid Mech., 27:419{448, 1967.
[50] Maddocks, J. and R.L. Sachs. On the stability of KdV multi-
solitons. Comm. Pure Appl. Math., 46:867{901, 1993.
[51] Magri, F. A simple model of integrable Hamiltonian equations. Jour.
Math. Phys., 19:1156{1162, 1978.
[52] Newell, A. C. Solitons in Mathematics and Physics. CMBS-NSF
Monographs. Society for Industrial and Appied Mathematics, Philadel-
phia, 1985.
[53] Novikov, S., Manakov, S.V., Pitaevskii, L.P. and Zakharov,
V.E. Theory of Solitons. Consultants Bureau, New York and London,
1984.
[54] Oishi, S. Relationship between Hirota's method and the inverse spectral
method- the Korteweg-deVries equation's case. Jour. Phys. Soc. Japan,
47:1037{1038, 1979.
[55] Olver, P.J. Hamiltonian and non-Hamiltonian models for water waves.
In P.G. Ciarlet and M. Roseau, editor, Trends and Applications of Pure
Mathematics to Mechanics, volume 195 of Lecture Notes in Physics,
pages 273{290, Providence, R.I., 1984. Springer-Verlag.
[56] Olver, P.J. Hamiltonian perturbation theory and water waves. Con-
temporary Mathematics, 28:231{249, 1984.
130 BIBLIOGRAPHY
[57] Olver, P.J., & P. Rosenau. Tri-Hamiltonian duality between soli-
tons and solitary-wave solutions having compact support. Physical Re-
view E, 53:1900{1906, 1996.
[58] Ono, H. Algebraic solitary waves in stratied
uids. Jour. Phys. Soc.
Japan, 39:1082{1091, 1975.
[59] Palais, Richard S. The symmetries of solitons. Bull. Amer. Math.
Soc., 34:339{403, 1997.
[60] Poppe, C. The Fredholm determinant method for the KdV equations.
Physica D, 13:137{160, 1984.
[61] Rayleigh, J.W.S. On Waves. Phil. Mag., 1, 1876.
[62] Rayleigh, J.W.S. Theory of Sound, I. II. Dover, New York, 1945.
[63] Riesz, F. & Sz. Nagy. Functional Analysis. Frederick Ungar Pub-
lishing Co., New York, 1955.
[64] Rosenau, P. Nonlinear dispersion and compact structures. Phys. Rev.
Lett., 73:1737{1741, 1994.
[65] Russell, J. Scott. Report on waves. In John Murray, editor, Report
on the Fourteenth Meeting, pages 311{390+57 plates, London, 1844.
British Association for the Advancement of Science.
[66] Sattinger, D.H. Topics in Bifurcation Theory, volume 309 of Lecture
Notes in Mathematics. Springer Verlag, 1973.
[67] Sattinger, D.H. Hamiltonian hierarchies on semi-simple Lie algebras.
Studies in Applied Mathematics, 72:65{86, 1985.
[68] Sattinger, D.H. & V.D. Zurkowski. Gauge Theory of Backlund
Transformations, II. Physica 26D, pages 225{250, 1987.
[69] Spitzer, L. Physics of Fully Ionized Gases. Interscience, New York,
1956.
[70] Stoker, J.J. Water Waves. Wiley Interscience, New York, 1957.
[71] Stokes, G.G. On the theory of oscillatory waves. Trans. Camb. Phil.
Soc., 8:197{229, 1847.
BIBLIOGRAPHY 131
[72] Struik, D.J. Determination rigoueuse des ondes irrotationnelles
periodiques dans un canal a profondeur nie. Math. Ann., 95:595{634,
1926.
[73] Toland, J.F. On the existence of a wave of greatest height and Stokes'
conjecture. Proc. Roy. Soc. London, A, 363:469{485, 1978.
[74] Whitham, G.B. Linear and Nonlinear Waves. John Wiley & Sons,
New York, 1973.
[75] Wineberg, S.B., J. McGrath, E. Gabl, L.R. Scott, & C.
Southwell. Implicit spectral methods for wave propagation problems.
Comp. Phys., 97:311{336, 1991.
[76] Zakharaov, V.E., and A.B. Shabat. Exact theory of two dimen-
sional self-focusing and one dimensional self-modulation of waves in non-
linear media. Soviet Phys. JETP, 34:62{69, 1972.
[77] Zakharov, V.E. Stability of periodic waves of nite amplitude on the
surface of a deep
uid. J. Appl. Mech. Tech. Phys., 2:190{194, 1968.
Index
132