Sei sulla pagina 1di 18

Physica D 152153 (2001) 416433

Long-time dynamics of the modulational


instability of deep water waves
M.J. Ablowitz
a
, J. Hammack
b
, D. Henderson
b
, C.M. Schober
c,
a
Department of Applied Mathematics, University of Colorado, Campus Box 526, Boulder, CO 80309-0526, USA
b
Department of Mathematics, Pennsylvania State University, University Park, PA 16802, USA
c
Department of Mathematics and Statistics, Old Dominion University, Norfolk, VA 23529, USA
Abstract
In this paper, we experimentally and theoretically examine the long-time evolution of modulated periodic 1D Stokes
waves which are described, to leading-order, by the nonlinear Schrdinger (NLS) equation. The laboratory and numerical
experiments indicate that under suitable conditions modulated periodic wave trains evolve chaotically. A Floquet spectral
decomposition of the laboratory data at sampled times shows that the waveform exhibits bifurcations across standing wave
states to left- and right-going modulated traveling waves. Numerical experiments using a higher-order nonlinear Schrdinger
equation (HONLS) are consistent with the laboratory experiments and support the conjecture that for periodic boundary
conditions the long-time evolution of modulated wave trains is chaotic. Further, the numerical experiments indicate that
the macroscopic features of the evolution can be described by the HONLS equation. Ultimately, these laboratory experi-
ments provide a physical realization of the chaotic behavior previously established analytically for perturbed NLS systems.
2001 Elsevier Science B.V. All rights reserved.
Keywords: Nonlinear Schrdinger equation; Modulated periodic waves; Long-time evolution
1. Introduction
Historically, the study of water waves has provided
researchers with a wide variety of interesting non-
linear phenomena. One of the classical examples of
nonlinear waves was the discovery by Stokes [1] in
1847 of traveling nonlinear periodic wave trains in
deep water. More precisely, Stokes found the lead-
ing terms of a series expansion for the surface wave
displacement, and how the frequency of the wave de-
pended on the amplitude for small but nite values of
the wave amplitude. In 1925, Levi-Civita [2] proved

Corresponding author.
E-mail addresses: markjab@newton.colorado.edu (M.J. Ablowitz),
cschober@lions.odu.edu (C.M. Schober).
rigorously that an innite series describing periodic
waves could be obtained and that it converged. How-
ever, the question of stability of these periodic water
waves remained open until 1967 when Benjamin and
Feir [3] established, experimentally and analytically,
that in sufcient deep water the Stokes water wave was
unstable.
The instability result can be obtained by consider-
ing slowly modulated wave trains. It can be shown that
for small amplitude, the leading-order complex ampli-
tude of the surface displacement satises the nonlin-
ear Schrdinger (NLS) equation [4,5,9]. The relevant
periodic solution of the NLS equation can be readily
shown by Fourier methods to be unstable in deep
water. Even though the modulational instability of
periodic waves has been established, the long-time
0167-2789/01/$ see front matter 2001 Elsevier Science B.V. All rights reserved.
PII: S0167- 2789( 01) 00183- X
M.J. Ablowitz et al. / Physica D 152153 (2001) 416433 417
behavior of the instability has not been completely
investigated. A natural question is whether the NLS
equation provides a satisfactory description of the
dynamics. Our results indicate that the answer to this
question is, in general, negative for periodic wave
trains.
The NLS equation is completely integrable, has
an innite number of conserved quantities and does
not possess temporally chaotic orbits [6,7]. How-
ever, water wave dynamics are described only to
leading-order by the NLS equation. The results in
this paper indicate that the perturbations to the NLS
equation associated with periodic wave trains gener-
ate chaotic evolutions. Since there is no mathematical
proof of dynamical chaos available, we use this term
only in a broad sense. The higher-order nonlinear
Schrdinger (HONLS) equation, which is obtained by
retaining terms in the asymptotic expansion through
fourth-order (see Eq. (3)), destroys the symmetry of
the NLS equation with respect to space translations
[18,20]. We show that the higher-order corrections
have a signicant effect on the evolution of the insta-
bility and of the wave train.
Earlier studies on near-integrable nonlinear wave
equations demonstrated that the modulational instabil-
ity can give rise to quite complicated dynamics. Using
initial data for quasiperiodic solutions near to homo-
clinic orbits (which we will denote as semi-stable
initial data), certain damped-driven and Hamiltonian
perturbations of the sine-Gordon and NLS equations
have been shown to trigger temporally chaotic evolu-
tions. In the seminal papers [11,12], the role of linear
instabilities and homoclinic orbits in the generation
of chaotic dynamics for damped-driven perturbations
of the sine-Gordon equation was established along
with the spectral criterion for these instabilities. Sim-
ilarly, the role of these structures in Hamiltonian
perturbations of the NLS and sine-Gordon equations
have been examined in the context of computational
chaos [14,15,17,2729] and chaotic energy transport
[25,26].
For spatially symmetric data, the mechanism for
chaotic behavior involves random crossings of the
critical level sets of the constants of motion or homo-
clinic crossings (see, e.g. [13,15,28]). Signicantly,
for a damped-driven perturbation of the NLS equa-
tion, analytical arguments have been found when even
symmetry is imposed to explain the onset of chaotic
dynamics [24]. The persistence of hyperbolic solutions
and the transversal intersection of their homoclinic
manifolds has been rigorously proven using singular
perturbation theory and Melnikov analysis adapted
to the innite-dimensional setting [24]. It should be
noted that establishing chaos in the PDE framework
is technically quite difcult and for the purely disper-
sive perturbation of the NLS considered in [15,17,28]
(which is relevant to the water wave problem dis-
cussed here) the persistence of homoclinic orbits and
chaotic dynamics has not yet been rigorously proven.
When evenness is removed, little is known the-
oretically. In this case, numerical studies on com-
putational chaos in nonlinear wave equations have
provided considerable insight. For example, using
semi-stable initial data for the NLS it was shown that
if even symmetry was not preserved in the numerical
codes then roundoff errors can excite an odd com-
ponent which subsequently evolves chaotically [16].
More recently, the evolution of asymmetric initial
data has been studied numerically for a Hamiltonian
perturbation of the NLS equation as well as for a
simple symmetry-breaking perturbation of the NLS
[17,23]. The work in [17] identied for the rst time a
distinctive new mechanism whereby homoclinic tran-
sition states develop near the homoclinic manifolds.
The solutions are characterized by random switching
across standing wave states into left- and right-going
modulated traveling waves. The occurrence of chaos
without homoclinic crossings in the noneven regime
is a novel mechanism by which nonlinear disper-
sive wave systems can be chaotically excited and as
discussed in this paper is observable in laboratory
experiments.
In this paper (in [22], some of these results were
announced), we examine in detail the water wave
approximation to the NLS equation and consider
the following questions: (1) Are chaotic evolutions
in deep water physically observable? If so, then the
NLS is not an adequate description, and the per-
turbations to the NLS equation are critical to the
evolution. (2) What characterizes the wave train
418 M.J. Ablowitz et al. / Physica D 152153 (2001) 416433
evolution? (3) Are the dominant features of the evolu-
tion captured by the HONLS model? We show exper-
imentally and analytically that periodically modulated
nonlinear Stokes water waves are, in fact, tempo-
rally chaotic, non-symmetric and not reproducible. In
contrast, when the envelope of the slow modulation
is taken to be that of a soliton, the experiments are
reproducible.
The Floquet spectral theory of the NLS equation
is used to determine the characteristics of the evo-
lution of the wave train. This type of normal-mode
analysis, obtained by projecting numerical data
onto integrable nonlinear modes, was rst applied
to certain perturbations of the Toda chain such as
the FermiPastaUlam chain [10] and later to per-
turbations of the sine-Gordon and NLS equations,
e.g. [11,17] amongst others. Here we compute the
spectrum, in particular the discrete eigenvalues, and
observe when they evolve into sensitive homoclinic
regimes which in turn indicates bifurcations in the
waveform between different physical states. The
spectral decomposition of the laboratory data demon-
strates that the leftright switching mechanism for
chaotic excitations occurs in the water wave problem.
In numerical experiments with the HONLS equation
we also nd numerous leftright homoclinic transi-
tions. The correlation between the results of the lab-
oratory and numerical experiments indicates that the
macroscopic or gross features of the wave evolution
are indeed modeled by the HONLS equation.
2. Analytical background
2.1. Governing equations
The equations governing the surface waves are
given by
2
= 0, for z ,
z
0 as z ,
where (x, z, t ), (x, t ) are the velocity potential and
free surface displacement, respectively, and by the
boundary conditions on the free surface z = (x, t ),

t
+
x

x
=
z
,
t
+g +
1
2
()
2
= 0. (1)
In the small amplitude approximation, the velocity po-
tential is expanded about z = 0. For slowly modulated
waves one assumes the ansatz
= (Ae
i+|k|z
+)
+
2
(

+A
2
e
2(i+|k|z)
+) + ,
= (B e
i
+) +
2
( +B
2
e
2i
+) + , (2)
where = kx t, denotes complex conjugate and
the deep water dispersion relation is used:
2
= g|k|
with k
0
= 0.44 rad/cm (neither damping or surface
tension is taken into account in the theory). The vari-
ables A,

, A
2
are assumed to be functions of X =
x, Z = z, T = t , and B, , B
2
are functions of X
and T only. is a dimensionless number, i.e. a mea-
sure of small amplitude and balances slow modula-
tion; = ka, where a is the size of the initial surface
displacement. Substituting the above ansatz into the
expanded form of the free surface equation (1) leads
to the following perturbed NLS equation on z = 0
([19], corrected for misprints)
2i
_
A
T
+

2k
A
X
_

_
_

2k
_
2
A
XX
+4k
4
|A|
2
A
_
=
2
_
i
2
8k
3
A
XXX
+2k
3
iA
2
A

X
12k
3
i|A|
2
A
X
+2k

X
A
_
.
Note that

satises
2

= 0 with the boundary
conditions:

Z
= (2k/g)(|A|
2
)
X
on z = 0 and

0 as z . The free surface amplitude is


obtained from = (/g)((iA + (/2k)A
x
)e
i

(k
2
/)A
2
e
2i
) +(). We introduce the following di-
mensionless and translating variables: T

= T, X

=
kX, Z

= kZ,

= k,

= (2k
2
/)

, u =
(2

2k
2
/)A, =
1
8
T

, = X


1
2
T

. Solving
Laplaces equation on Z 0 by Fourier meth-
ods for

in terms of A yields the following perturbed
nonlocal NLS equation (HONLS):
iu

+u

+2|u|
2
u +(
1
2
iu

6i|u|
2
u

+iu
2
u

+2u[H(|u|
2
)]

) = 0, (3)
where H(f ) represents the Hilbert transform of the
function f . The Fourier transform of the Hilbert trans-
form yields

H(f ) = i sign(k)

f (k). In the periodic
case (0, L), k is replaced by k
n
= 2n/L. These
Fourier relations are readily implemented numerically
M.J. Ablowitz et al. / Physica D 152153 (2001) 416433 419
when using a pseudo-spectral method, which is the
method we employ. We also note that on the innite
line the physical space realization of the Hilbert trans-
form is H(f ) = (1/)PV
_

(f ()/ x) d, where
PV denotes the Cauchy principal value integral.
2.2. Integrable theory of the NLS
Setting = 0 in (3) with x and t , we
obtain the focusing cubic NLS equation which can be
written in Hamiltonian form
i
t
_
u
u
_
= J
_
_
_
_
H
u
H
u
_
_
_
_
(4)
with
J =
_
0 1
1 0
_
,
and Hamiltonian
H(u, u) =
_
L
0
(|u
x
|
2
|u|
4
) dx. (5)
Zakharov and Shabat [8] established the complete in-
tegrability of the NLS equation by discovering its Lax
pair:

x
= L
(x)
,
t
= L
(t )
, (6)
where
L
(x)
=
_
i iu
i u i
_
,
L
(t )
=
_
i[2
2
u u] 2iu +u
x
2i u u
x
i[2
2
u u]
_
. (7)
The compatibility condition for the pair of linear
systems (7) is satised for all values of the complex
spectral parameter if the coefcient u(x, t ) satises
the NLS equation (3) at = 0. If one imposes period-
icity by requiring the potential u(x, t ) to be periodic
in the spatial variable x with period L, then one can
characterize u (for any xed time t ) in terms of its
Floquet spectrum
(L
(x)
(u)) := { C|L
(x)
vvv = 0, |vvv| bounded x}.
(8)
Given a fundamental solution matrix M(x; ) (such
that M(0; ) = I) for the Lax pair (6) with potential
u(x), one denes the Floquet discriminant to be the
trace of the transfer matrix M(L; ) across one period:
(u; ) := Tr[M(L; )]. (9)
Since det[M(x; )] = 1, the Floquet spectrum can be
characterized as follows:
(L
(x)
(u)) := { C|(u; ) is real and
2 (u; ) 2}. (10)
The properties of the Floquet discriminant that we
will use in the following sections are:
1. (u; ) is constant under the NLS evolution.
2. Floquet discriminants at different values of the
spectral parameter Poisson commute, for =

,
{(u; ), (u;

)},
where the Poisson bracket is dened as
{F, G} = i
_
L
0
_
F
u
G
u

F
u
G
u
_
dx.
These properties show that (u; ) encodes the in-
nite family of NLS constants of motion (in fact,
parameterized by C).
Within the discrete spectrum, periodic/antiperiodic
eigenvalues are the roots of (u; ) = 2. We also
distinguish the following points of the spectrum:
1. The simple periodic/antiperiodic spectrum

s
=
_

s
j
|(, u) = 2,
d
d
= 0
_
. (11)
2. Critical points of spectrum
c
j
, specied by the con-
dition d(u; )/d|
=
c = 0.
3. Double points of the periodic/antiperiodic spectrum

d
=
_

d
j
|(, u) = 2,
d
d
= 0,
d
2

d
2
= 0
_
.
(12)
The nonlinear spectral transform is used to represent
solutions in terms of a set of nonlinear modes whose
structure and dynamical stability is determined by
420 M.J. Ablowitz et al. / Physica D 152153 (2001) 416433
the location of the corresponding element of the pe-
riodic/antiperiodic spectrum [11]. Generic multiple
points have multiplicity 2 and the location of the
double points plays a particularly important role in
the geometry of the phase space. Real double points
label inactive nonlinear modes whereas complex dou-
ble points are in general associated with linearized
instabilities of the NLS equation and label the orbits
homoclinic to the unstable solution [12].
2.3. Modulational instability and homoclinic
solutions
An issue of main importance is the stability of
the periodic Stokes wave. For the nondimensionalized
NLS equation the Stokes wave, or plane wave, is given
by u
0
(x, t ) = a e
2i|a|
2
t
, where for convenience, a is
assumed to be real. After dimensionalizing and trans-
forming to the surface displacement , the exponent
corresponds to the nonlinear frequency shift found by
Stokes that includes the term 4k
2
|A|
2
in Eq. (2). Its
stability can be examined by considering perturbations
of the form u(x, t ) = u
0
(1 + (x, t )) and linearizing
for small . Assuming that (x, t ) =
n
(0) e
i
n
x+i
n
t
+

n
(0) e
i
n
xi
n
t
,
n
= 2n/L, it is found that the
growth rate
n
is given by
2
n
=
2
n
(
2
n
4a
2
). Thus,
the plane wave is unstable to long wavelength pertur-
bations, i.e. provided 0 < (n/L)
2
< |a|
2
and the
number of unstable modes is the largest integer M sat-
isfying 0 < M < |a|L/. This stability criterion is ap-
plicable to the modulated periodic wave train produced
via
p
(t ) described in the laboratory experiments.
Another approach to examining instabilities is using
the associated Floquet theory. The discriminant for the
plane wave u
0
(x, t ) is readily computed to be
(a, ) = 2 cos(
_
a
2
+
2
L). (13)
Then, the associated Floquet spectrum consists of con-
tinuous bands R

[ia, ia], and a discrete part con-


taining the simple periodic/antiperiodic eigenvalues
i|a|, and the innite number of double points

2
n
=
_
n
L
_
2
a
2
, n Z. (14)
The complex double points, 0 < (n/L)
2
< |a|
2
(this
is the same condition as for linear instability), are as-
sociated with critical saddle-like level sets and can be
used to label their homoclinic orbits. The remaining

n
s for |n| > M are real double points. If [aL/] =
M, there are 2M complex (pure imaginary) double
points. Fig. 1a shows the spectrum of the plane wave
for the case M = 1. Although homoclinic orbits of the
modulationally unstable plane wave can be explicitly
constructed, here we simply provide the waveform of
the homoclinic solution along with its associated spec-
trum (see Fig. 1a). The homoclinic solution is char-
acterized by a single mode which limits to the plane
wave as t . The isospectral set of the plane
wave includes the homoclinic orbit. In the numerical
experiments, we use initial data which are small per-
turbations of unstable plane waves with M = 1, 3, 5
complex double points or unstable modes. The exper-
imental data specically corresponds to perturbations
of the Stokes wave in the three unstable mode regime.
Note that since the real axis is always spectrum
and (L) has the symmetry, (L) then


(L), only the spectrum in the upper-half -plane is
displayed in the spectral plots throughout the paper.
Numerical experiments involving perturbed NLS
and sine-Gordon equations have shown that the exis-
tence of instabilities and their associated homoclinic
orbits can generate a variety of interesting phenomena,
including temporally chaotic evolutions. We summa-
rize the results of [17] which are relevant in interpret-
ing the laboratory and numerical experiments.
2.4. Nearby states
We give a brief description of the multi-phase so-
lutions whose Floquet spectra are O() close to one
of the modulationally unstable plane wave with one
complex double point. The spectral congurations
are computed at t = 0. In [17], initial data of the
form u(x, 0) = u
0
+ u
1
= a + [e
i
1
cos(x) +
r e
i
2
sin(x)] are considered, where 0 < 1, r
and
i
s, for i = 1, 2, are real parameters and a
real frequency to be selected so that the perturbation
affects one or more specic complex double points.
The selection criteria are found to be: for a = m/L,
M.J. Ablowitz et al. / Physica D 152153 (2001) 416433 421
Fig. 1. The surface |u(x, t )| and the nonlinear spectrum with (a) one double point for a homoclinic solution of NLS,
(1)
= 0;
(b) one imaginary gap for a standing wave solution of NLS, u
0
= 0.5(1 + 0.1 cos x); (c) one cross for a standing wave so-
lution of NLS locked in the center and wings, u
0
= 0.5(1 + 0.1i cos x); (d) right state: a right-traveling wave solution of NLS,
u
0
= 0.5(1+0.05(e
90i
cos x+e
0i
sin x)); (e) left state: a left-traveling wave solution of NLS, u
0
= 0.5(1+0.05(e
0i
cos x+e
30i
sin x)).
if =
j
= 2j/L, 1 m 1, then the jth
complex double point splits into two simple points
and a complex band of spectrum or a gap is created.
The parameters
1
and
2
govern the symmetry of the
solution: if
1
=
2
+n, then the perturbed spectrum
exhibits the symmetry (this corresponds to
the solution being even in the spatial variable, i.e.
q(x) = q(x), a constraint commonly imposed in
the study of perturbations of the NLS equation). In
this case the double point splits along the imaginary
axis (gap conguration, see Fig. 1b) or symmetrically
about the imaginary axis (cross state, see Fig. 1c). In
422 M.J. Ablowitz et al. / Physica D 152153 (2001) 416433
the symmetric case, the homoclinic orbit separates the
symmetric subspace into disjoint invariant subman-
ifolds. Due to the analyticity of the discriminant ,
under small perturbations it is possible to evolve from
one conguration to the other, while maintaining the
even symmetry, only by passing through the complex
double point, i.e. by crossing the homoclinic manifold.
This even symmetry constraint, rst considered in
[11], has been used almost exclusively in the literature
so that homoclinic crossings can be easily identied.
On the other hand, for generic
1
and
2
, the se-
lected complex double point splits asymmetrically in
the complex plane. (For a complete analysis of the
O() splitting of the complex double points in the
noneven regime, see [17].) When one complex double
point is present the two basic spectral congurations
associated with noneven perturbations are as follows:
(1) The resulting upper band of spectrum lies in the
rst quadrant and the lower band lies in the second
quadrant. The wave form is characterized by a single
mode traveling to the right (right state, see Fig. 1d).
(2) The resulting upper band of spectrum lies in the
second quadrant and the lower band lies in the rst
quadrant. The wave form is characterized by a single
mode traveling to the left (left state, see Fig. 1e). The
Fig. 2. (a) Surface for u
0
= 0.5(1 + 0.01(e
0.9i
cos x + e
60i
sin x)) obtained with the difference scheme DDNLS for 0 < t < 500 and
5000 < t < 5500; (b) The nonlinear spectrum at three time slices.
main result obtained for near-integrable dynamics in
the noneven regime is that the spectrum can evolve
between the two distinct congurations for left- and
right-traveling waves without passing through a com-
plex double point; the homoclinic orbit does not sep-
arate the full NLS phase space.
2.5. Temporal chaos in the noneven regime
In the one double point regime, generically the
spectral conguration evolves from the right state to
the left state by crossing a nearby cross state or gap
state, not by moving through a double point. Similarly,
when more complex double points are involved, the
evolution of the spectrum from one conguration to
another can be accomplished by executing an appro-
priate sequence of crossings of gap and cross states,
and again does not entail homoclinic crossings. So
do chaotic evolutions develop in the noneven regime
in nearby systems? The answer is an emphatic yes,
but they are no longer characterized by homoclinic
crossings; instead they are produced by a new mech-
anism involving random bifurcations through nearby
standing waves (see Fig. 2a). As an example, consider
the initial data for a left-traveling modulated 3-phase
M.J. Ablowitz et al. / Physica D 152153 (2001) 416433 423
solution of the NLS equation
u
0
= 0.5(1+0.01(e
0.90i
cos x +e
60i
sin x)) (15)
with L = 2

2, = 2/L. Fig. 2a shows the


surface (0 < t < 500) obtained with the follow-
ing Hamiltonian discretization (DDNLS): i u
n
=
(u
n+1
+u
n1
2u
n
)/h
2
+2|u
n
|
2
u
n
= 0 for N = 24.
Initially, the waveform travels to the left; as time
evolves the perturbation induced by the discretiza-
tion causes the waveform to jump between left- and
right-traveling waves. We observe that this bifurca-
tion occurs randomly and intermittently throughout
the entire time series for 0 < t < 10 000.
The bifurcation between left- and right-traveling
waves occurs when the spectrum evolves through a
nearby cross state (corresponding to a standing wave
solution) and not by executing homoclinic crossings.
In the surface plot (Fig. 2a), the rst bifurcation from
a left- to right-traveling wave occurs at t = 131.8. The
evolution of the spectrum due to the perturbation in-
duced by DDNLS is shown in Fig. 2b at three succes-
sive time slices in this transition region when the wave-
form bifurcates from a left state (t = 131.5) to a
right state (t = 132.1) by evolving through a cross
state (t = 131.8). The evolution of spectrum was
computed to 5000 and it showed random intermittent
bursts in which the spectral conguration bifurcates
between the left state and the right state. (For a com-
plete presentation of the spectral results and descrip-
tion of the behavior in the noneven regime, see [17].)
Although it is still an open question as to what
analytical object under perturbation gives rise to the
chaos in the noneven regime and how to rigorously
prove the mechanism, it seems intuitively clear that
the random switching between left- and right-running
waves is similar to the switching between standing
wave states observed in the even regime. In the non-
even regime the signature of the chaos is different
as there are no homoclinic crossings (observed both in
discrete problems and now in water waves, as shown
in Section 4). Even so, a persisting homoclinic struc-
ture seems a likely candidate to play a role in the en-
suing chaos. For the results of a Melnikov analysis of
a simpler symmetry-breaking perturbation of the NLS
carried out to explain the chaotic evolutions in terms
of a transversal intersection of the homoclinic mani-
folds of persisting hyperbolic sets, see [23].
In the asymptotic theory, the HONLS equation is
a more accurate approximation to the water wave
dynamics and since it destroys even symmetry, ho-
moclinic crossings are not allowed. However, the
leftright switching mechanism for chaotic excitations
is available and is in fact observed in the laboratory
experiments. The question at hand is: In the descrip-
tion of deep water waves, when is the NLS equation
adequate and when is the HONLS equation critical to
the description?
3. Laboratory results
3.1. Experimental apparatus and procedures
The experimental apparatus comprised a wavetank
with motion-controlled instrumentation carriage, wa-
ter supply, wavepaddle, wavegages, and computer
system. The glass wavetank is 14.3 m long by 25.4 cm
wide and 20 cm in depth with a sanded, glass beach
at the downstream end for energy absorption. At
this depth it is sufcient to use the deep water limit.
Mounted on the top are siderails that support the
instrumentation carriage that is attached to a belt.
The belt is driven by a servo-controlled motor that
allows the instrumentation carriage to translate down-
stream of the wavemaker at the group velocity of the
underlying wave train.
The tank is cleaned with alcohol before it is lled
with water to a depth greater than the desired 20 cm.
The water is distilled in-house, then run through a l-
tering system that removes ions, organics and parti-
cles. A lm of baby powder is spread on the surface
(following Scott, 1979). The carriage is tted with a
brass rod and then translated down the tank to scrape
the surface lm toward the end of the tank. This lm
moves as a rigid body in front of the carriage, un-
til it reaches the end of the tank where it is vacu-
umed with a wet-vac. This procedure provides for a
reproducible water surface, for which the linear damp-
ing rate of the 3.33 Hz wave train is 6 10
4
cm
1
.
This value is more than the standard boundary layer
424 M.J. Ablowitz et al. / Physica D 152153 (2001) 416433
calculation of clean surface damping, but less than that
of a contaminated surface calculation.
The water is allowed to rest for 15 min before ex-
periments are conducted; the surface is cleaned at
least every 2 h. Waves are generated using a vertically
oscillated, anodized wedge with both position and ve-
locity control. The programmed velocity for both sets
of experiments is the linear, uid particle velocity;
the programmed position is the desired free-surface
displacement.
The theoretical formulation of Section 2 does not
include the effects of viscous damping. To obtain a
measure of viscous decay in the experiments, we mea-
sured the maximum amplitude of the envelope soliton
at 23 positions down the tank. This amplitude decayed
exponentially at a rate of 6 10
4
cm
1
/s.
The surface displacement is measured using ve
capacitance-type wave gages. Gage 0 is xed at 40 cm
downstream from the wavemaker. It is an intrusive,
capacitance-type wavegage. Gages 14 were mounted
40 cm apart on the carriage; they are non-intrusive,
capacitance wave gages that span the width of the
tank, and thus average out any surface motion in the
direction perpendicular to wave propagation.
The purpose of gage 0 is to measure the water sur-
face displacement near the wavemaker to insure that
the surface displacement there is reproducible. A me-
chanical cam is used to close a switch that begins the
data collection; it has a few milliseconds of slop. To
adjust for that, we shift the total time series of all ve
gages by a few milliseconds based on the correlation
coefcients between the time series obtained by gage
0. That is, we compare the time series from gage 0
from every experiment with the same initial conditions
to that obtained for the rst such experiment. We shift
the time series of all the gages by the small amount
necessary to give the maximum correlation coefcient
between the measurements obtained at gage 0. This
small shift (a few milliseconds out of a 65 s time se-
ries) insures that the starting point of each set of time
series is the same from experiment to experiment.
The time series from gage 0 have correlation coef-
cients among experiments of 0.99 or better, except for
two experiments (among 40) for which the correlation
coefcients were 0.96 or 0.97. This high correlation
indicates that, indeed, the initial conditions, i.e. the
water surface displacement near the wavemaker, was
reproducible within a small noise level.
Gages 14 are mounted on the carriage so that two
types of time series are obtained from them. The rst
are xed measurements in which the gages are located
at a particular position along the tank. The second
are measurements obtained in a reference frame that
translates at the linear group speed of the underlying
wave train. The carriage that supports the gages is at
rest for 21.28 s to allow for transient motions to pass
and then set in motion at the desired speed.
In the temporal measurements we compare time se-
ries obtained from different experiments with identi-
cal initial conditions. These measurements are graphed
against each other to produce a phase plane diag-
nostic for reproducibility. If the results of the two
experiments are identical the graph will be the 45

line. In particular, the time series from gage 0 near


the wavemaker produces a 45

line with a very slight


width, indicating a 12% noise level. This indicates
what should be expected from time series showing the
waveeld evolution.
3.2. Envelope solitons
The control experiment is performed with the
soliton solution of NLS since theoretically it
should evolve reproducibly (see below). The wave-
maker was programmed to oscillate as
s
(t ) =
a sin(
0
t )sech(a
0
t /

2) where
0
= 20.94 rad/s,
a = 0.2 cm, k
0
= 0.44 rad/cm,

(k
0
) = 24.4 cm/s
(group velocity), g = 980 cm/s
2
and T = 71.9 dynes/
cm (surface tension).
Fig. 3 shows time series obtained by the ve wave
gages when the carriage is near the wavemaker. There
is evidence of some dispersion in the envelope soliton.
The effects of viscous decay are observed as well, e.g.
in Fig. 4, which shows an envelope soliton near the
wavemaker and 800 cm downstream of it.
To determine if the soliton experiment is repro-
ducible, the experiment was run every 15 min over a
2.5 h period. Time series are obtained at gage 4 when
the carriage was 800 cm downstream of the wave-
maker. Fig. 5a graphs the water surface displacements
M.J. Ablowitz et al. / Physica D 152153 (2001) 416433 425
Fig. 3. Time series from gages (a) 0, (b) 1, (c) 2, (d) 3, and (e) 4 obtained when the carriage was xed so that gage 1 was 40 cm from gage 0.
measured in the rst and last experiments against each
other. There was a slight phase shift between the two
data sets. Shifting each data point in one of the data
sets the same amount (0.00921 s), produces the nearly
perfect 45

line (Fig. 5a). The experiments using soli-


tons indicate that we are able to conduct a repro-
ducible experiment when the initial conditions are
such that the waveeld evolution is predicted to be
reproducible. The soliton experiment reects a stable
nonchaotic evolution which the NLS equation ade-
quately describes.
3.3. Modulated periodic wave trains
For modulated wave trains the position of
the wavemaker is programmed to be
p
(t ) =
a sin(
0
t )(1 +
E
sin
p
t ) where a = 0.5 cm,
p
=
1.047 rad/s,
E
= 0.1, the values
0
, g, k
0
, T are the
same as in the soliton case and the corresponding
unperturbed periodic wavelength of the modulation is
L = 147 cm.
Fig. 6 shows output from the ve wavegages for
the modulated wave train. The output from gage
0, shown in Fig. 6a, shows a periodic modulated
wave train. The output from gages 14 are shown
in Figs. 6be. For times less than 21.3 s the carriage
supporting these gages is at rest and so the periodicity
is the same as that near the wavemaker. At 21.3 s, the
carriage moves with the group speed of the waves
as evidenced by the Doppler shift that occurs at this
time. One would expect that when the gage starts to
move at the group speed, the amplitude at the time
426 M.J. Ablowitz et al. / Physica D 152153 (2001) 416433
Fig. 4. Time series from gages (a) 0, (b) 1, (c) 2, (d) 3, and (e) 4 obtained when the carriage was xed so that gage 1 was 840 cm from
gage 0.
the carriage begins motion would remain constant,
except for some viscous decay, as happens in Fig. 6e.
However, this does not occur in the traces of Figs.
6bd, presumably because of a mismatch in linear
and actual group speed. We note that evidence of
reections occur after about 61 s.
For the modulated wave train initial data, phase
plane plots show that the wave is reproducible near
the wavemaker. However, as the wave travels down
the tank we obtain Lissajous-type gures (see, e.g.
Fig. 5b) indicating that a complicated phase shift de-
velops between the waves of two experiments and
cannot be simply removed. Indeed subsequent exper-
iments show that different Lissajous gures, with di-
verging phase trajectories are obtained, each of which
correspond to different complicated phase shifts. The
phase shifts are sensitive to small changes in initial
data. Unlike the soliton, the two time series start to
diverge indicating the experiment is irreproducible.
Spatial data associated with modulated periodic
wave trains was also obtained yielding a different
perspective of the evolution. Some of the spatial data
were used in the numerical experiments, discussed
below, as initial conditions. The spatial envelope is
reconstructed by concatenating 40 sets of data for
each of the four gages. In the 40 experiments, the
initial location of the carriage differs by 1 cm suc-
cessively for each experiment. The result is 160 time
series of the water surface spaced 1 cm apart which
are used to reconstruct the spatial prole of the water
surface, 160 cm long, by concatenating the data sets.
Our ability to measure a spatial envelope by conduct-
ing 40 experiments requires the experiments to be
reproducible.
Fig. 7 shows six spatial proles obtained in this
way. Fig. 7a shows the spatial prole near the wave-
maker, before the wave has reached the wavegages.
This prole provides a benchmark for the level of
M.J. Ablowitz et al. / Physica D 152153 (2001) 416433 427
Fig. 5. Phase plane plots for: (a) experimental soliton data; (b)
experimental modulated wave train data.
noise as indicated by the miniscule blips in what
should actually be a at surface level. At t = 15 s,
the waveform is somewhat close to the wavemaker
and the blips in the data are not signicant (Fig. 7c).
Further down the tank additional crests start to form.
The blips become signicant and no longer represent
a simple non-smoothness in the wave prole. By 40 s,
the periodicity of the underlying wave train is lost
(Fig. 7f). The degeneration of the spatial coherence
of the waveeld indicates that the experiments are not
reproducible. This experimental irreproducibility and
the theory presented below are evidence that modu-
lated Stokes wave trains evolve chaotically for certain
parameter regimes.
4. Numerical experiments
We have shown experimentally that periodically
modulated nonlinear Stokes waves can evolve chaoti-
cally. Armed with this information and the analytical
background, in this section we turn to the issue of
determining the qualitative features of the evolution
and whether it can be successfully modeled with the
HONLS equation. This is accomplished by the fol-
lowing: (1) we do some additional post-processing of
the data from the laboratory experiments. We calcu-
late the Floquet spectral decomposition of the labora-
tory data at sampled times. This establishes that the
water wave dynamics is characterized by leftright
homoclinic transitions. (2) We numerically examine
the long-time dynamics of the HONLS equation. We
rst establish the parameter regimes for which the
HONLS exhibits chaotic behavior (if at all) using
model initial data. When solutions to the HONLS
equation are regular and O() close to NLS solutions
for the same initial data we consider the NLS equation
to adequately describe the dynamics in that regime.
Once the basic HONLS dynamics is understood,
we examine the HONLS using experimental data as
initial data, to allow a closer comparison with the
laboratory experiments. The diagnostics indicate that
the numerical experiments compare very well with
the laboratory experiments. For certain regimes (i.e.
when a higher number of unstable modes are present)
the higher-order nonlinear terms become critical and
for these cases the dominant features of the chaotic
behavior is captured by the HONLS equation.
In the numerical experiments the parameter values
are specied in the nondimensionalized framework
and have been carefully matched with the parame-
ters used in the laboratory experiments. We use a
fourth-order pseudo-spectral code for integrating the
HONLS equation (3) with N = 512 Fourier modes
in space and a fourth-order adaptive Runge-Kutta
scheme in time. As in the laboratory experiments,
reproducibility is studied using phase plane plots.
In the phase plane plots the evolution of the sur-
face displacement obtained using initial data
u(x, 0) is graphed against the evolution of the sur-
face displacement

obtained using u

(x, 0), where


u

(x, 0)=u(x, 0)(1 + u


r
(x)), is on the order of
experimental error (1%) and u
r
(x) is taken to be a
random eld. The formula for the reconstruction of
(the phase plane plot is the one diagnostic where
dimensional coordinates are presented) in terms of u
428 M.J. Ablowitz et al. / Physica D 152153 (2001) 416433
Fig. 6. Time series from gages (a) 0, (b) 1, (c) 2, (d) 3, (e) 4. Gage 0 is at rest, 40 cm from the wavemaker. Gages 14 are mounted on a
carriage that translates downstream at the linear group velocity starting at 21.3 s.
can be found in Section 2. The associated nonlinear
spectral theory of the NLS equation is also used to
investigate the dynamics. The data provided by both
the physical and numerical experiments is projected
onto the nonlinear spectrum of the NLS and we fol-
low its evolution in time to determine changes in the
nonlinear mode content. Although in the experiments
the spectrum is computed every dt = 0.1, it is only
shown at sampled times.
The benchmark case is the soliton case using model
initial data of the form u(x, 0) = sech(x), i.e. a soli-
ton with zero velocity. The spacetime evolution of
the waveform obtained using the HONLS equation
with = 0.14, for 0 < t < 5, is given in Fig. 8a.
The soliton develops an O() velocity and sheds a
small amount of radiation off the front of the soli-
ton, but the dynamics is regular. The phase plane plot,
for 0 < t < 10, remains close to the 45

line with
little spread (Fig. 8b), as was observed in the labo-
ratory experiments. Although the spectrum is not in-
variant, the spectral plots show that the spectrum does
not change conguration and that there are only small
O() changes in the amplitude and speed of the soli-
ton. This indicates that we obtain a reproducible ex-
periment when the initial conditions are solitons and
supports the notion that, for envelope solitons for the
time scales under consideration, the unperturbed NLS
equation adequately describes the long-time dynamics.
The other model initial data we considered cor-
responded to modulationally unstable periodic wave
trains and is of the form
u(x, 0) = a(1 +
T
cos
n
x) (16)
with
T
= 0.1. We varied the amplitude a in the
M.J. Ablowitz et al. / Physica D 152153 (2001) 416433 429
Fig. 7. Spatial proles corresponding to times of (a) 1.00 s, (b) 12.00 s, (c) 15.00 s, (d) 29.00 s, (e) 33.00 s, and (f) 40.00 s after the start
of data collection. The dots represent the experimental data.
Fig. 8. (a) Surface for u(x, 0) = sech x obtained using the HONLS equation with = 0.14 for 0 < t < 5; (b) Phase plane plot of the
surface amplitude (mm) vs.

for model soliton initial data for 0 < t < 10.


430 M.J. Ablowitz et al. / Physica D 152153 (2001) 416433
instability criterion to have M = 1, 3, 5 unstable
modes nearby (and we denote this as the M unsta-
ble mode regime). In [21], we have shown that for
data T there are 2n + 1 simple imaginary points in
the spectrum:
0
and {
n+
,
n
}, n = 1, 2, . . . , M.
When
T
is asymptotically small, the simple points

n+
,
n
of the spectrum of L get successively closer
to each other with their distance from a double point
being O(
n
T
), n = 1, 2 . . . , M. The distance from a
double point can be made arbitrarily small by taking
M sufciently large. Thus, the evolution can exhibit
homoclinic transitions in which case the dynamics is
irregular and chaotic.
For initial data in the one unstable mode regime (e.g.
(16) with a = 0.5,
T
= 0.1 and L = 2

2), we
found that the solution of the HONLS equation with
= 0.14, 0 < t < 2.5, is well approximated by the
solution obtained using the NLS equation. The wave-
form does not display any temporal irregularities and
homoclinic transitions do not occur; stable nonchaotic
dynamics ensue. The phase plane plot is very close
to the 45

line. The spectral conguration is the least


complicated of all the modulated wave train cases. The
initial data is for a standing wave gap state. As time
evolves, the discrete eigenvalues remain well separated
and do not evolve into sensitive regions. By t = 2.5,
Fig. 9. (a) Surface for initial data (16) with a = 0.7 and L = 4

2 obtained using the HONLS equation with = 0.28 for 0 < t < 10;
(b) Phase plane plot of the surface amplitude (mm) vs.

for model soliton initial data for 0 < t < 10.


the upper band of spectrum has moved slightly to the
right of the imaginary axis which is reected by the
waveform developing a very small speed. Temporal
irregularities or homoclinic transitions do not occur.
As in the soliton case, for the one unstable mode case
the NLS equation yields a satisfactory description of
the dynamics as the HONLS terms do not signicantly
alter the dynamics.
As the number of unstable modes is increased,
the NLS equation loses this ability. Fig. 9a shows
the surface obtained using the HONLS equation with
= 0.28, 0 < t < 10 for initial data (16) with
a = 0.7,
T
= 0.1 and L = 4

2. In this higher un-


stable mode regime (M = 3), the behavior of HONLS
solutions is considerably more complicated than what
was previously observed for the discrete systems with
noneven initial data (cf. Section 2). Here, each of
the unstable modes is exhibiting leftright switching
leading to a very complex waveform. In addition to
displaying temporal chaos, the solution appears to
be spatially irregular. The phase plane plot is spread
signicantly away from the 45

line (see Fig. 9b).


Additionally, numerous homoclinic transitions are
observed in the spectrum. These transitions are sim-
ilar to those observed for the experimental initial
data case (also in the three unstable mode regime).
M.J. Ablowitz et al. / Physica D 152153 (2001) 416433 431
Fig. 10. (a) Phase plane plot of vs.

for experimental initial data for 0 < t < 5. Notice the similarity to Fig. 5b. Spectral plots
corresponding to times of (b) 0.0, (c) 0.5, (d) 1.0, (e) 2.0, (f) 2.1, and (g) 2.6. Solid darkened curves are curves of spectrum and we have
included some of the curves of real (the dashed curves) to give an indication of the topological changes in the spectrum.
We refer the reader to the spectral plots for the ex-
perimental data (Figs. 10bf) for a sample of the
changes in the spectral conguration that occur in
this regime. This is a generalization of the basic
mechanism for chaos in the noneven regime that we
observed for one unstable mode (cf. Section 2). In
the spectral plots only the spectrum related to the
dominant low modes is shown. The amplitude of
the higher modes is very small and the spectrum of
these radiative states is not depicted as the radiation
432 M.J. Ablowitz et al. / Physica D 152153 (2001) 416433
modes are not signicant in the description of the
chaotic state. For the case M = 5, we found that the
phase plane plot diverges from the 45

line even more


strongly than for M = 3. The spectrum evolved sig-
nicantly and we found more numerous homoclinic
transitions than with M = 3. The cases M = 3, 5
yield strong temporal irregularities and, for the time
scales under consideration, chaotic dynamics. The ex-
amination of the model data demonstrates that when
there are a higher number of unstable modes present
initially, the NLS is inadequate and the HONLS per-
turbations make a signicant difference to the nal
evolution.
The numerical results for the HONLS using the
experimental data as initial data provide evidence of
chaotic evolutions consistent with the laboratory re-
sults. For the experimental data, denoted as E we
use u(x, 0) = the spatial envelope of the modulated
wave train obtained near the wavemaker (see Fig. 7b).
This data corresponds to a multi-phase solution in the
three unstable mode regime. Using initial data E, for
short times the experiment is reproducible and the
phase plane plot stays close to the 45

line. As the
waveeld evolves, the experiment is rendered irrepro-
ducible. A complicated phase shift develops between
the experiments that changes with time resulting in a
phase plane plot (Fig. 10a) that is remarkably similar
to that of the laboratory data (Fig. 5b).
The spectral results are striking. The spectrum at
t = 0 (Fig. 10b) depicts four curves of spectrum
with seven simple eigenvalues as the end points of
spectrum. The three nearby complex double points
(labeling the three unstable modes) correspond to
the center locations between the simple eigenvalues.
We number them according to their distance from
the origin with the rst mode being farthest. The
spectrum evolves signicantly in time and a number
of homoclinic transitions between modes occur. As
mentioned previously, only the spectrum related to
the dominant low modes is shown in Fig. 10 as the
transitions occur within a xed, low-dimensional set
of nonlinear modes. Figs. 10bg give an overview of
the changes in the spectral conguration and show the
spectrum at six time slices, at (b) t = 0.0, (c) t = 0.5,
(d) t = 1.0, (e) t = 2.0, (f) t = 2.1, and (g) t = 2.6.
In Figs. 10b and c the transition in the orientation of
the third and fourth curves of spectrum indicates a
bifurcation in the third nonlinear mode of the solution
between left- and right-traveling. Similarly, between
Figs. 10c and d, there is a change in the orientation
of the rst and second curves indicating a homoclinic
transition for the rst mode. Later in the evolution,
the eigenvalues and homoclinic double points move
away from the imaginary axis (see Figs. 10eg). In
this sequence of plots the second and third curves
in the spectral conguration switch orientation and
back again indicating leftright ipping in the second
mode. For the timeframe examined, 0 < t < 5, there
are frequent random leftright homoclinic transitions
in all of the unstable nonlinear modes. Each of these
transitions corresponds to a change in the character-
istics of the nonlinear modes and leads to physical
changes in the wave eld.
Finally, we remark upon the spectral decomposition
of the laboratory data. Using the actual physical data
(not evolving with the HONLS equation) we compute
the spectrum. Fig. 10b and Figs. 11a and b provide
the spectrum at t = 12, 29 and 32 s, respectively. Sig-
nicantly, the evolution is unmistakably characterized
by leftright homoclinic transitions! These three time
slices demonstrate a switching in the orientation of
the rst and second bands of spectrum indicating that
the rst mode is switching from left- to right-traveling.
We also note that at later times (see Figs. 11a and b)
there are only two nonreal eigenvalues, as opposed to
three earlier, as indicated in Fig. 10b. We believe the
Fig. 11. Spectral plots for the laboratory data corresponding to
times of (a) 29.0 s, and (b) 32.0 s. Solid darkened curves are curves
of spectrum and we have included some of the curves of real
(the dashed curves) to give an indication of the topological
changes in the spectrum.
M.J. Ablowitz et al. / Physica D 152153 (2001) 416433 433
reason for the loss of an eigenvalue, i.e. reduction
from three unstable modes to two unstable modes, is
due to viscous damping which has not been incorpo-
rated into the theory. Nevertheless, it is remarkable that
the leftright switching scenario is still apparent. Thus,
the dissipation does not interact signicantly with the
chaotic mechanism. Based upon the results presented
in this paper, the long-time evolution of the modu-
lational instability and subsequent chaotic dynamics
is adequately described by the HONLS equation (3).
However, we note that (linear) damping should be
added to better describe amplitude changes. We shall
investigate this effect in the future.
Acknowledgements
This work was partially supported by the AFOSR
USAF, Grant No. F49620-00-1-0031 and the NSF,
Grant Nos. DMS-0070772, DMS-9803567 and
DMS-9972210.
References
[1] G.G. Stokes, Camb. Trans. 8 (1847) 441473.
[2] T. Levi-Civita, Math. Ann. XCIII (1925) 264.
[3] T.B. Benjamin, J.E. Feir, J. Fl. Mech. 27 (1967) 417430.
[4] V.E. Zakharov, Phys. J. Appl. Mech. Technol. Phys. 4 (1968)
190194.
[5] D.J. Benney, G.J. Roskes, Stud. Appl. Math. 48 (1969) 377
385.
[6] E.D. Belokolos, A.I. Bobenko, V.Z. Enolskii, A.R. Its,
V.B. Matveev, Algebro-geometric Approach to Nonlinear
Integrable Problems, Springer Series in Nonlinear Dynamics,
Springer, Berlin, 1994.
[7] M.J. Ablowitz, H. Segur, Solitons and the Inverse Scattering
Transform, SIAM, Philadelphia, PA, 1981.
[8] V.E. Zakharov, A.B. Shabat, Sov. Phys. JETP 34 (1972) 62
69.
[9] H.C. Yuen, B.M. Lake, Phys. Fluids 18 (1975) 956960.
[10] W.E. Ferguson, H. Flaschka, D.W. McLaughlin, J. Comput.
Phys. 45 (1982) 157209.
[11] A.R. Bishop, M.G. Forest, D.W. McLaughlin, E.A. Overman
II, Physica D 23 (1986) 293328.
[12] N. Ercolani, M.G. Forest, D.W. McLaughlin, Physica D 43
(1990) 349384.
[13] D.W. McLaughlin, E.A. Overman, Surv. Appl. Math. 1 (1995)
83203.
[14] M.J. Ablowitz, B.M. Herbst, Phys. Rev. Lett. 62 (1989) 2065
2068.
[15] D.W. McLaughlin, C.M. Schober, Physica D 57 (1992) 447
465.
[16] M.J. Ablowitz, B.M. Herbst, C.M. Schober, Phys. Rev. Lett.
71 (1993) 26832686.
[17] M.J. Ablowitz, B.M. Herbst, C.M. Schober, Physica A 228
(1996) 212235.
[18] E. Lo, C.C. Mei, J. Fluid Mech. 150 (1985) 395408.
[19] K.B. Dysthe, Proc. Roy. Soc. London A 369 (1979) 105114.
[20] K. Trulsen, K.B. Dysthe, Wave Motion 24 (1996) 281289.
[21] M.J. Ablowitz, C.M. Schober, Contemp. Math. 172 (1994)
253268.
[22] M.J. Ablowitz, J. Hammack, D. Henderson, C.M. Schober,
Phys. Rev. Lett. 84 (2000) 887890.
[23] A. Calini, C.M. Schober, Math. Comput. Simulation 55 (2001)
351364.
[24] Y. Li, D.W. McLaughlin, J. Shatah, S. Wiggins, Commun.
Pure Appl. Math. 49 (1996) 11751255.
[25] M.G. Forest, C.G. Goedde, A. Sinha, Math. Comput.
Simulation 37 (1994) 323339.
[26] M.G. Forest, C.G. Goedde, A. Sinha, Physica D 67 (1993)
347386.
[27] M.J. Ablowitz, B.M. Herbst, C.M. Schober, J. Comput. Phys.
126 (1996) 299314.
[28] A. Calini, N. Ercolani, D.W. McLaughlin, C.M. Schober,
Physica D 89 (1996) 227260.
[29] M.J. Ablowitz, B.M. Herbst, C.M. Schober, J. Comput. Phys.
131 (1997) 354367.

Potrebbero piacerti anche