Vdoc - Pub Nonlinear Pdes A Dynamical Systems Approach
Vdoc - Pub Nonlinear Pdes A Dynamical Systems Approach
I N M AT H E M AT I C S 182
Nonlinear PDEs
A Dynamical
Systems Approach
Guido Schneider
Hannes Uecker
Guido Schneider
Hannes Uecker
GRADUATE STUDIES
I N M AT H E M AT I C S 182
Nonlinear PDEs
A Dynamical
Systems Approach
Guido Schneider
Hannes Uecker
2010 Mathematics Subject Classification. Primary 35-01, 35Bxx, 35Qxx, 37Kxx, 37Lxx.
Copying and reprinting. Individual readers of this publication, and nonprofit libraries
acting for them, are permitted to make fair use of the material, such as to copy select pages for
use in teaching or research. Permission is granted to quote brief passages from this publication in
reviews, provided the customary acknowledgment of the source is given.
Republication, systematic copying, or multiple reproduction of any material in this publication
is permitted only under license from the American Mathematical Society. Permissions to reuse
portions of AMS publication content are handled by Copyright Clearance Center’s RightsLink
service. For more information, please visit: http://www.ams.org/rightslink.
Send requests for translation rights and licensed reprints to [email protected].
Excluded from these provisions is material for which the author holds copyright. In such cases,
requests for permission to reuse or reprint material should be addressed directly to the author(s).
Copyright ownership is indicated on the copyright page, or on the lower right-hand corner of the
first page of each article within proceedings volumes.
2017
c by Guido Schneider and Hannes Uecker. All rights reserved.
The American Mathematical Society retains all rights
except those granted to the United States Government.
Printed in the United States of America.
∞ The paper used in this book is acid-free and falls within the guidelines
established to ensure permanence and durability.
Visit the AMS home page at http://www.ams.org/
10 9 8 7 6 5 4 3 2 1 22 21 20 19 18 17
To Daniela, Max and Jonas
and Anja, Franka and Henrike
Contents
Preface xi
Chapter 1. Introduction 1
§1.1. The three classical linear PDEs 1
§1.2. Nonlinear PDEs 4
§1.3. Our choice of equations and the idea of modulation equations 6
§1.4. Overview 11
vii
viii Contents
“If you want to build a ship, don’t herd people together to collect wood and
don’t assign them tasks and work, but rather teach them to long for the
endless immensity of the sea”
Antoine de Saint-Exupery
xi
xii Preface
known as the Chafee-Infante problem. The second one is a very basic intro-
duction to the Navier-Stokes equations, with a focus on periodic boundary
conditions.
Genuine PDE phenomena such as transport, diffusion, and dispersion
can hardly be understood by the interpretation of PDEs as systems of in-
finitely many ODEs. In Part III we consider PDEs which are posed on
the real line. We start with the linear heat equation, and then turn to
nonlinear problems. For famous model equations such as the Kolmogorov-
Petrovsky-Piskounov or Fisher equation, the Korteweg-de Vries (KdV) equa-
tion, the Nonlinear Schrödinger (NLS) equation, and the Ginzburg-Landau
(GL) equation, we discuss the local existence and uniqueness of solutions,
special solutions as fronts and pulses, their stability and instability, soliton
dynamics, the construction of attractors, and some related results.
The equations from Part III all play an important role in mathematics
and have entire monographs devoted to each. Moreover, they have many
connections to physics and other fields of applications, where they are of-
ten used as simplest possible models for the description of some real world
phenomena. In Part IV we explore these connections from a mathematical
perspective. The scalar equations from Part III occur as asymptotic effective
models, or more specifically as modulation equations, for the more compli-
cated systems from physics considered in Part IV. Examples are pattern
forming systems which can be described by the GL equation, light pulses in
nonlinear optics which can be described by the NLS equation, or long waves
in dispersive systems which can be described by the KdV equation. We dis-
cuss how the dynamics of the reduced model equations transfer to the more
complicated systems. Thus, in Part IV we give a mathematically rigorous
presentation of the formalism of modulation equations in the context of real
world applications. While this last part is close to recent research, it is still
in textbook style, and often we do not prove the sharpest or most general
result possible, but instead refer to the literature for extensions.
All chapters are kept as self-contained as possible, such that the reader
can start to read directly about his or her favorite equation. Having a
good background in linear ODEs, cf. §2.1, a starting point for our goals and
objectives are §2.2-§2.3 about basic nonlinear ODE dynamics combined with
Part III. There are other possible combinations, for instance the sections
about dissipative dynamics or the sections about conservative dynamics.
Nevertheless the reader can also read the book from the beginning to the
end. See the Grasshopper’s Guide on page 12 for detailed proposals. All
chapters contain exercises which we strongly recommend not to skip.
Preface xiii
This book grew out of our manuscripts for the lectures and seminars we
gave about ODEs and PDEs at the universities of Bayreuth, Karlsruhe, Old-
enburg, and Stuttgart. We thank the students who attended our lectures
and seminars and urged us to keep the presentation simple and accessi-
ble. Moreover, we thank all friends and colleagues with whom we coop-
erated over the years, mainly on topics from Part IV, in particular, Dirk
Blömker, Tom Bridges, Kurt Busch, Martina Chirilus-Bruckner, Christo-
pher Chong, Walter Craig, Markus Daub, Hannes de Witt, Arjen Doel-
man, Tomas Dohnal, Wolf-Patrick Düll, Wiktor Eckhaus, Jean-Pierre Eck-
mann, Bernold Fiedler, Thierry Gallay, Dieter Grass, Daniel Grieser, Mark
Groves, Tobias Häcker, Mariana Haragus, Ronald Imbihl, Ralf Kaiser, Tasso
Kaper, Klaus Kirchgässner, Markus Kunze, David Lannes, Vincent Lescar-
ret, Karsten Matthies, Ian Melbourne, Andreas Melcher, Johannes Müller,
Robert Pego, Dmitry Pelinovsky, Jens Rademacher, Björn Sandstede, Arnd
Scheel, Zarif Sobirov, Aart van Harten, C. Eugene Wayne, Daniel Wetzel,
Peter Wittwer, and Dominik Zimmermann. We thank Stefanie Siegert and
the unknown referees for a number of additional proposals to improve the
presentation. Especially we thank Alexander Mielke from whom we learned
about nonlinear dynamics and PDEs.
Finally we would like to thank Ina Mette from the AMS for her never
ending motivation to go on with this book project and many helpful com-
ments to transform our lecture notes into a book.
Introduction
1
2 1. Introduction
Γ1
window
door Γ2
heating
well developed numerical schemes such as the finite element method (FEM)
or the boundary element method (BEM), which are available for computing
approximate solutions of such systems. Only in very special cases solutions
can be found analytically.
Example 1.1.2. The heat equation or diffusion equation
(1.2) ∂t u = Δu,
with u = u(x, t) and u : Ω × R+ → R, where t denotes time and x denotes
space, describes the evolution of quantities such as heat, chemical concentra-
tions, or the probability distribution of a particle obeying Brownian motion.
It can be derived as follows. Let V ⊂ Ω be an arbitrary subset with
smooth boundary. The change of the total quantity within V equals the
flux through ∂V , i.e.,
d
u dx = −
F, n dS = − ∇ · F dx
dt V ∂V V
with F the flux density,
·, · the scalar product in Rd , n : ∂Ω → Rd again
the outer normal, and where we used the Gauss’ integral theorem. Since
this relation is true for all sets V , we find
∂t u = −∇ · F.
Very often the flux density F is proportional to the gradient ∇u of the
concentration u, i.e., F = −a∇u with a constant a > 0. By rescaling time
we finally come to the diffusion equation (1.2).
In order to solve this equation uniquely in a domain Ω × R+ additional
conditions are needed. As in Example 1.1.1 we need boundary conditions,
but also the temperature field at time t = 0 has to be known, i.e., we need
the initial condition u|t=0 = u0 with u0 : Ω → R. Stationary solutions, i.e.,
time-independent solutions, satisfy the Laplace equation (1.1). The heat
equation is the prototype parabolic PDE. There is an extensive theory for
equations of the form ∂t u = Lu with an elliptic operator −L.
Example 1.1.3. The linear wave equation
(1.3) ∂t2 u = Δu,
u = u(x, t) with u : Ω × R → R, is a simple model for, e.g., oscillations of a
string (Ω ⊂ R) or of a membrane (Ω ⊂ R2 ), or the propagation of light in
vacuum. In order to solve this equation uniquely in a domain Ω × (t0 , t1 ),
t0 < 0 < t1 we again need boundary and initial conditions. Like for scalar
second order ODEs we need two initial conditions, namely u|t=0 = u0 and
∂t u|t=0 = u1 with u0 : Ω → R and u1 : Ω → R. The Dirichlet boundary
condition u|∂Ω corresponds to a membrane which is fixed at the boundary.
In this case, the boundary will reflect the waves.
4 1. Introduction
For the wave equation the eigenmodes play a crucial role. An eigenmode
is a solution u(x, t) = eiωt v(x). This yields the eigenvalue problem
−Δv = ω 2 v.
Such problems play an important role in applications, especially in elasticity
theory, where the evolution equations of linear elasticity
∂t2 u = μΔu + (λ + μ)∇(∇ · u)
yield to the eigenvalue problem
−μΔv + (λ + μ)∇(∇ · v) = ω 2 v.
If Ω is a bounded set then under suitable boundary conditions there are
countably many real eigenvalues λn = ωn2 with 0 ≤ λ1 ≤ λ2 ≤ . . . → ∞.
In the construction of cars, bridges, planes, etc., one has to take care that
these so called resonant modes are not periodically excited. Hence, there is a
big industry using FEM and BEM in order to solve these elliptic eigenvalue
problems. The wave equation is the prototype hyperbolic PDE. There is
an extensive theory for equations of the form ∂t2 u = Lu with an elliptic
operator −L.
For reasons explained below we will focus on other examples than the
three classical ones. The fundamental Examples 1.1.1-1.1.3 cannot be and
will not be avoided. However, they will only occur as subproblems which
will help to understand the nonlinear problems under consideration.
The world is instationary, i.e., almost all systems evolve with time.
Typical examples are a vibrating beam, the daily change of the weather, or
the motion of the planets in the solar system. Hence, from the beginning
we will consider nonlinear time-dependent systems.
A mathematical concept which is basic to the analytical understanding
of all ODEs and PDEs is the concept of Dynamical Systems. Until the
beginning of the 1960s, Laplace’s principle that with the knowledge of all
physical rules and the present state of the world, the past and future be-
havior of the world for all times can be computed, was widely accepted as a
relevant philosophical foundation of science. Starting already with the work
of H. Poincaré in the 1890s, cf. [Poi57], this principle was finally observed
to be practically useless at the beginning of the 1960s, for instance by the
work of the meteorologist E. Lorenz in 1963 [Lor63]. He observed with
an analog computer for a three-dimensional model for the weather that the
possible time for predictions goes logarithmically with the precision of the
initial conditions, i.e., that long-time weather-forecasts are practically im-
possible. See Figure 1.2 for an illustration of the so called Lorenz attractor
and of the sensitivity of solutions w.r.t. the initial conditions.
Certain ODEs and PDEs, or, more general, dynamical systems, can be
classified as chaotic. The visualization of chaotic dynamical systems was
in fashion in the 1980s. Famous examples are the Mandelbrot and the Julia
sets. In this book, chaos will not play a central role, but one should keep in
mind its existence already in low-dimensional dynamical systems.
20
0
xp=0
x =1e−3
p
x =1e−5
p
−20
0 10 20
Example 1.3.2. At the end of the 20th century a new generation of high
speed ferries has caused serious problems, especially those that cross the
Channel between England and France and those operating in the Marl-
borough sound in New Zealand. The waves created by these ferries can
propagate without loss of energy over large distances, and thus retain the
potential to create enormous havoc when they come ashore, and as a conse-
quence of a fatal accident and other damage there are now speed limits for
these ferries [Ham99].
Again a modulation equation gives an idea to understand these phenom-
ena. The Korteweg-de Vries (KdV) equation
∂τ A = ν1 ∂ξ3 A + ν2 A∂ξ A,
with τ ∈ R, ξ ∈ R, A(ξ, τ ) ∈ R and coefficients νj ∈ R can be derived
with the help of a perturbation ansatz. The KdV equation is a universal
modulation equation which describes long waves of small amplitude, where
the original system condenses to the coefficients νj ∈ R.
Like the NLS equation, this famous nonlinear equation possesses soliton
solutions, very robust solitary waves. These waves interact like particles,
i.e., after some nonlinear interaction they reform and look exactly as before
the interaction. This observation, made in the middle of the 1960s, that
solutions of a PDE show simultaneously the behavior of a particle and a
wave, had a big influence on nonlinear science due to the similarity with the
particle-wave dualism in quantum mechanics.
For a long time the KdV equation has also been suggested as a model
for the description of tsunamis, water waves of only a few meters height, but
with a length of up to 100km, i.e., in the ocean they cannot be observed by
eye. In the 5000m deep pacific ocean they move with a very high velocity
of around 700km/h. If they approach land they become slower and steeper,
and cause serious floodings. However, data which is now available from the
tsunami at Christmas 2004 in the Indian Ocean seem to indicate that soliton
dynamics had played at least for this tsunami no role on the open sea. The
validity of the KdV equation will be discussed in Chapter 12.
Example 1.3.3. Since the 1960s, systems near the onset of a finite wave
length instability have been analyzed in detail using modulation equations.
These amplitude modulations describe slow changes in time and space of
the envelope of the finite wave length pattern close to the first instability.
The most famous and generic of such equations is the Ginzburg-Landau
(GL) equation
∂τ A = ν2 ∂ξ2 A + ν0 A + ν3 A|A|2 ,
with τ ≥ 0, ξ ∈ R, A(ξ, τ ) ∈ C and coefficients νj ∈ C. Famous pat-
tern forming systems which can be described with the GL equation are
10 1. Introduction
1
∂t u = −∂x3 u − ∂x (u2 ) − ε(∂x2 + ∂x4 )u,
2
√
with t ≥ 0, x ∈ R, u = u(x, t) ∈ R, and where 0 < ε ≈ θ − θc 1 is a
small parameter. Therefore, complicated dynamics that are present in this
equation occur directly at the first instability of the inclined plane problem.
The dynamics is dominated by traveling pulse trains consisting of unstable
pulses. Time series of the position of the pulses indicate the occurrence of
chaotic dynamics. This situation is relevant for cooling units. Again the 3D
Navier-Stokes equations for the water flowing down the unit is replaced by
a simpler model still containing very complicated dynamics.
Another situation where the Burgers equation arises as a modulation
equations are phase or wave number modulations of stable periodic pattern
in a pattern forming system, while phase (or wave number) modulations of
unstable pattern are generically described by Kuramoto-Shivashinsky type
of equations.
1.4. Overview
In order to keep the book as an introductory text and as self-contained as
possible, in Part I we explain basic dynamical systems concepts for ODEs,
such as phase space, fixed points, periodic solutions, attractors, stability and
instability, bifurcations and amplitude equations.
12 1. Introduction
1.4.1. Grasshopper’s Guide. To some extent the four parts of this book
are intended to be independent. Moreover, the chapters are kept as self-
contained as possible, such that the reader may start to read directly about
his or her favorite equation. Therefore, we also give the following guide.
Part I can obviously be read independently of the rest of the book. It is
an example-oriented basic course on finite-dimensional dynamical systems
which together with Chapters 5 and 6 (and possibly Chapter 13) yields a two
semester course about finite- and infinite-dimensional dynamical systems.
Chapters 7 and 8 of Part III can subsequently serve as a basis for a seminar.
An alternative one or two semester course is given by §2.2-§2.3 about
basic nonlinear ODE dynamics combined with (parts or all of) Part III and
some parts of Part IV, for instance the beginning of Chapter 10. Other
chapters of Part IV can then serve as a basis for a seminar.
1.4. Overview 13
1.4.3. Software. There are many software packages for the numerical so-
lution of ODEs and the graphical presentation of solutions. Matlab, Maple,
and Mathematica have built in facilities, and there are various simple to
use Java applets available. We strongly encourage the reader to do own
experiments with any of these programs.
From the above remarks about the very different types of PDEs it readily
follows that there cannot be a general tool for all types of PDEs. However,
tools for specific types of PDEs, both commercial and free are widely avail-
able. We use some short self-written matlab scripts to illustrate some PDE
dynamics, mostly for model problems. However, we do not discuss any
numerical methods behind these programs and refer to [Uec09] and the
references therein. For the computation of so called bifurcation diagrams
we refer to AUTO [Doe07, Dea16] and pde2path [UWR14].
14 1. Introduction
Exercises
1.1. Classify the following PDEs as linear or nonlinear.
d
d
a) ∂t u = ∂xi ∂xj (aij u)+ ∂xi (bi u), aij , bi : Rd → R smooth functions.
i,j=1 i=1
The operator L is called elliptic if the eigenvalues of the symmetric matrix A = (aij )
are strictly positive. It is called hyperbolic if they are nonzero, but have different
signs. It is called parabolic if the associated quadratic form (∂x → ξ, ∂y → η)
defines a parabola. Classify
a) 3∂x2 u + 10∂x ∂y u + 15∂y2 u + 36∂x u + 12∂y u + 17 = 0;
b) 3∂x2 u + 4∂x u + ∂y u + 2 = 0.
1.3. Consider the PDE ∂t u = ∂x u for u = u(x, t).
a) Find the general solution for x ∈ R.
b) Solve the PDE for x ∈ (0, 1) with the initial condition u(x, 0) = 1 for
x ∈ (0, 1) under the boundary condition u(1, t) = cos t.
c) Is it possible to solve the PDE for x ∈ (0, 1) with the initial condition
u(x, 0) = 1 for x ∈ (0, 1) and the boundary condition u(0, t) = cos t?
1.4. Consider a membrane Ω = (0, 1)2 which is fixed at the boundary ∂Ω, i.e.,
u|∂Ω = 0.
a) Make an ansatz u(x, y, t) = v(t) sin(mπx) sin(nπy), (n, m ∈ N) for the
solutions of ∂t2 u = Δu. Which equation is satisfied by v?
b) Solve the equation for v with the initial conditions v(0) = 0 and v̇(0) = 1.
c) Sketch for fixed m, n ∈ N the set of (x, y) ∈ Ω, for which u(x, y, t) = 0 for
all t ∈ R.
Chapter 2
15
16 2. Basic ODE dynamics
are the local and global existence and uniqueness of solutions, special so-
lutions such as fixed points, periodic solutions, homoclinic and heteroclinic
orbits, and further concepts such as stability and instability, invariant mani-
folds, ω-limit-sets, attractors, and chaotic dynamics. Many of these concepts
will later be transferred to nonlinear PDEs. Moreover, the search for special
solutions, such as front or pulse solutions for the PDEs, in later chapters
very often lead to ODE problems as they are considered here.
The behavior and the analysis of an ODE or of a PDE strongly differ
between dissipative and conservative systems. In Chapter 3 we provide the
strategy and the tools to tackle dissipative systems. Such systems are typi-
cally characterized by the existence of compact absorbing sets, i.e., compact
sets into which all solutions finally enter. In dissipative systems very often
more complicated and eventually chaotic dynamics occur through bifurca-
tions if some external parameter is varied. After introducing a number of
elementary bifurcations for one- and two-dimensional systems we introduce
with the Lyapunov-Schmidt reduction and the center manifold theorem two
reduction methods which allow us to find these elementary bifurcations in
higher dimensional systems, too. Chapter 3 is closed by presenting some
routes of bifurcations to chaotic behavior in dissipative systems.
The systems considered in Chapter 3 change the volume in phase space,
but many systems in nature preserve the volume in phase space, especially
those of classical mechanics. Thus, Chapter 4 is devoted to Hamiltonian
ODE dynamics. We provide some tools for their analysis and explain basic
facts about their behavior, which shows fundamental differences compared
to that of the systems of Chapter 3. For instance, a globally attracting fixed
point cannot exist in conservative or volume-preserving systems. Therefore,
stability and instability proofs or the route to chaotic behavior must be
completely different. The starting point of the bifurcation analysis is not a
globally attracting fixed point, but a so called completely integrable system.
In Chapter 4 we also discuss KAM theory which allows to understand the
behavior of systems which are small perturbations of completely integrable
Hamiltonian systems.
The ideas presented in this first part will reappear in subsequent sections.
For instance, Chapter 3 about dissipative ODE dynamics contains basic tools
which will be used in Chapters 5-7, §8.3-Chapter 10, and Chapter 14 about
dissipative PDE dynamics. Similarly, Chapter 4 about conservative ODE
dynamics contains basic tools which will help to understand §8.1, §8.2, and
Chapters 11-12 about conservative PDE dynamics.
We emphazise that the purpose of Part I is not to give a comprehen-
sive overview about ODEs. Rather we present the basic ideas of nonlinear
dynamics as needed in subsequent parts of this book in the analysis of PDEs.
2.1. Linear systems 17
2.1.2. Local existence and uniqueness. We briefly recall that the initial
value problem for (2.2) has a unique solution, and that the solutions of
(2.2) form a d-dimensional vector space. The following local existence and
uniqueness result and many other results in this book are based on the
contraction mapping principle which is absolutely fundamental in nonlinear
20 2. Basic ODE dynamics
Lemma 2.1.5 only asserts local existence and uniqueness. The way to
show existence and uniqueness of solutions beyond t0 + δ is to prove bounds
on u(t0 + δ). The key tool is Gronwall’s inequality which will be used in
many proofs below. We first restrict to a simple version [Ver96, Theorem
1.2].
Lemma 2.1.7. (Gronwall’s inequality) For t ∈ (t0 , t0 + a) with a > 0,
and φ and ψ non-negative continuous functions assume that
t
(2.3) φ(t) ≤ ψ(s)φ(s) ds + δ.
t0
Proof. Dividing (2.3) by its right-hand side and multiplication of both sides
with ψ(t) yields after integration that
t t
ψ(τ )φ(τ )
τ dτ ≤ ψ(τ ) dτ
t0 t0 ψ(s)φ(s) ds + δ t0
t
t
which implies that ln t0 ψ(s)φ(s) ds + δ − ln δ ≤ t0 ψ(τ ) dτ and finally
that
t t
ψ(s)φ(s) ds + δ ≤ δ exp ψ(τ ) dτ .
t0 t0
By assumption φ(t) is smaller than the expression on the left-hand side.
From the integrated ODE
t
u(t) = u0 + A(s)u(s) ds
t0
we find the inequality
t
u(t)Rd ≤ u0 Rd + A(s) u(s)Rd ds
t0
22 2. Basic ODE dynamics
Proof. Since the solutions of (2.2) depend linearly and one-to-one on the
initial conditions u0 we have that the set of solutions of (2.2) is isomorphic
via u → u(t0 ) to the space of initial conditions, i.e., Rd .
Definition 2.1.10. The matrix-valued function t → φ(t) is called funda-
mental matrix if φ̇(t) = A(t)φ(t) and if φ(t0 ) has full rank for a t0 ∈ R.
Remark 2.1.11. From the local existence and uniqueness Theorem 2.1.8,
it immediately follows that φ(t) has full rank for all t ∈ R. If t → φ(t)
is a fundamental matrix then u(t) = φ(t)φ(t0 )−1 u0 solves the initial value
problem with u|t=t0 = u0 . If t → ψ(t) is another fundamental matrix
then φ(t)φ(t0 )−1 = ψ(t)ψ(t0 )−1 such that there exists an invertible matrix
C = φ(t0 )−1 ψ(t0 ) which is independent of t with ψ(t) = φ(t)C.
2.1.4. The exponential matrix. In general, (2.2) can only be solved ex-
plicitly in case d = 1. For d ≥ 2, if a solution is known of the d-dimensional
problem, then the dimension of the problem can be reduced by one, i.e.,
after the reduction a linear ODE in d − 1 space dimensions has to be solved,
cf. [Cod61, Page 118]. However, in case
(2.7) u̇ = Au, u|t=0 = u0 ,
with A ∈ Rd×d independent of t, all solutions can be computed explicitly.
The ansatz u(t) = eλt u yields the eigenvalue problem Au = λ u, and in
case that A has d linearly independent eigenvectors φ1 , . . . , φd ∈ Rd with
eigenvalues λ1 , . . . , λd , the general solution reads u(t) = di=1 ci eλi t φi with
ci , . . . , cd ∈ R. In case that there are complex eigenvalues or Jordan blocks
this formula becomes slightly more complicated. Equation (2.7) appears
as linearization around fixed points of nonlinear systems and hence plays a
crucial role.
Remark 2.1.12. Linear scalar equations of nth order
n−1
(n)
y (t) + aj y (j) (t) = 0,
j=0
24 2. Basic ODE dynamics
with aj ∈ R, can always be written as a linear first order system. For the
construction of the explicit solution
this is not necessary. The r different ze-
roes λk with multiplicity kr , i.e. rk=1 kr = n, of the characteristic equation
λn + n−1 j
j=0 aj λ = 0 allow to construct the general solution, namely
r −1
r k
y(t) = ck,j eλk t tj
k=1 j=0
Obviously this also holds for the series differentiated w.r.t. t. Similarly, we
obtain the estimate
etA u0 Rd ≤ etA u0 Rd .
where the time derivative and the infinite sums can be interchanged due
to the uniform convergence w.r.t. t on compact intervals. Moreover, A and
the infinite sum can be interchanged due to the boundedness and hence
continuity of A.
The solution operator etA can be expressed in terms of the Jordan normal
form J of A. The change of coordinates u = Sy in u̇ = Au yields ẏ =
S −1 ASy = Jy. We have
or equivalently
−1 tA −1 t2 A2
S e S =S 1 + tA + + ... S
2
−1 t2 S −1 ASS −1 AS
= 1 + tS AS + + ...
2
t2 J 2
= 1 + tJ + + . . . = etJ .
2
Since
⎛⎛ ⎞ ⎞
⎛ tJ ⎞
J1 0 e 1 0
⎜⎜ .. ⎟ ⎟ ⎜ .. ⎟
exp(tJ) = exp ⎝⎝ . ⎠ t⎠ = ⎝ . ⎠
0 Jr 0 etJr
it is sufficient to consider
We have
Nkμ = (δi,j−μ ), for μ = 0, ..., k − 1,
Nkμ = 0, for μ = k, k + 1, ...
26 2. Basic ODE dynamics
and so finally
⎛ tk−1
⎞
1 t. . . . . . (k−1)!
⎜ ⎟
⎜0 1 . . . tk−2 ⎟
⎜ (k−2)! ⎟
⎜ .. ⎟
etNk = ⎜ .. . . . . . . . . .
. ⎟
.
⎜ . ⎟
⎜ .. ⎟
⎝ . 1 t ⎠
0 ... 0 1
This representation formula immediately yields a statement about the sta-
bility of the fixed point u = 0 of the ODE u̇ = Au.
Definition 2.1.13. a) The fixed point u = 0 is called asymptotically
stable for u̇ = Au if for all u0 ∈ Rd we have u(t) = etA u0 → 0 for
t → ∞.
b) The fixed point u = 0 is called stable for u̇ = Au if for all u0 ∈ Rd
we have that u(t) = etA u0 stays bounded for all t ≥ 0.
c) In all other cases the origin u = 0 is called unstable.
Theorem 2.1.14. a) If all eigenvalues of A have strictly negative real
parts, then u = 0 is asymptotically stable.
b) If A possesses no eigenvalue with positive real part and if all eigen-
values with real part zero possess the same algebraic and geometric
multiplicity, i.e., no non-trivial Jordan block, then the origin u = 0
is stable.
c) In all other cases, i.e., if A possesses at least one eigenvalue with
strictly positive real part or at least one eigenvalue with real part
zero with algebraic multiplicity bigger than the geometric multiplic-
ity, then the origin u = 0 is unstable.
3) We plot the flow, i.e., a number of chosen orbits. These are curves
which are defined by the solutions t → u(t) ∈ R2 .
A combination of 2) and 3) is called a phase portrait. The choice of the size
of h and of the orbits depends on the problem, and also is a matter of taste.
Remark 2.1.15. The vector field and the direction field are both tangent
vectors of the solution t → u(t). Hence, the differential equations u̇ = f (u)
and u̇ = f (u)/f (u)R2 have the same orbits, i.e., solution curves in the
phase plane, although their dynamics are very different.
One more option is to plot the nullclines. These are the sets
Nj := {(u1 , u2 ) ∈ R2 : fj (u1 , u2 ) = 0},
for j = 1, 2, where the vector field is vertical, respectively horizontal. The
intersection points of N1 and N2 give the fixed points u∗ of the ODE, i.e.,
points with f (u∗ ) = 0. If the solution starts in a fixed point, the solution
stays in that fixed point, i.e., u(t) = u∗ for all t. Often, nullclines at least
partially coincide with coordinate axis, and, moreover, for (non-degenerate)
linear systems we have N1 ∩ N2 = {(0, 0)} as the only fixed point.
Due to §2.1.4 it is sufficient to consider (2.2) with A ∈ R2×2 in Jordan
normal form. There are the following cases.
a) The eigenvalues have the same algebraic and geometric multiplicity,
i.e.,
λ1 0
A = with a1) 0 = λj ∈ R or a2) 0 = λ1 = λ2 . In
0 λ2
case a1) we distinguish three subcases i) λ1 = λ2 , ii) λ1 > λ2 > 0
and iii) λ1 > 0 > λ2 . All other cases are obtained from i)-iii) by a
reversal of time t → −t.
b) The eigenvalue has geometric
multiplicity one and algebraic multi-
λ 1
plicity two, i.e., A = is a Jordan block with λ ∈ R.
0 λ
c) The degenerate case of at least one eigenvalue λ = 0. Besides the
trivial case A = 0 there are the cases c1) 0 = λ1 < λ2 and c2)
λ = 0 with geometric multiplicity one and algebraic multiplicity
two.
In the following we consider a number of examples to visualize these cases,
see Figure 2.1.5.
1 0
a1 i): Let A = , i.e., u̇1 = u1 , u̇2 = u2 . For the solutions
0 1
we find u1 (t) = et u1 (0), u2 (t) = et u2 (0). The orbits are straight
lines since u1 (t)/u2 (t) = u1 (0)/u2 (0) = const.. The nullclines are
28 2. Basic ODE dynamics
1 1 1 1
0 0 0 0
−1 −1 −1 −1
−1 0 1 −1 0 1 −1 0 1 −1 0 1
1 1 1 1
0 0 0 0
−1 −1 −1 −1
−1 0 1 −1 0 1 −1 0 1 −1 0 1
Figure 2.1. Phase portraits for a1i), a1ii), a1iii), and b) in the first
row, for a2i), a2ii), c1) and c2) in the second row.
Proof. Since
φ̇(t + T ) = A(t + T )φ(t + T ) = A(t)φ(t + T )
with φ(t) also φ(t + T ) is a fundamental matrix. Hence, there exists an
invertible d × d-matrix C, such that
φ(t + T ) = φ(t)C.
Each invertible d × d-matrix C can be written as C = eT B with B a non-
unique d ×d-matrix. As example consider C = diag(λ1 , . . . , λd ) with λj > 0.
Then a logarithm is given by B = diag(ln λ1 , . . . , ln λd ). For the general case
use −1 = eiπ and the expansion of ln(1 + λ) in case of Jordan blocks. See
Exercise 2.5. For P (t) = φ(t)e−tB we obtain
P (t + T ) = φ(t + T )e−(t+T )B = φ(t)Ce−T B e−tB = φ(t)e−tB = P(t).
Definition 2.1.18. The matrix C = eT B is called monodromy matrix. The
eigenvalues of C are called Floquet multipliers, and the eigenvalues of B are
called Floquet exponents.
Im
Im
1
1
Re Re
2 −1
Example 2.1.20. Consider the iteration xn+1 = Cxn , with C = .
0 1
The solution
can be computed explicitly by the transformation
x =Sy, with
1 1 −1 2 0
S= . We find xn+1 = SB n+1 S x0 , with B = .
0 1 0 1
Example 2.1.21. Consider the 1-periodic ODE u̇(t) = cos2 (2πt)u(t) for
u(t) ∈ R. Using cos2 (2πt) = 1/2 + cos(4πt)/2 the solution with initial
condition u(0) = u0 is given by
t sin 4πt
u(t, 0, u0 ) = u0 exp +
2 8π
32 2. Basic ODE dynamics
and therefore
sin 4πt t
P (t) = exp and etB = exp .
8π 2
We find the Floquet multiplier e1/2 and the Floquet exponent 1/2. The
time-one-map is given through Φ1 u0 = e1/2 u0 .
The following example [MY60] shows that in the periodic case the eigen-
values of the matrix A(t) have no significance for the stability of u = 0.
cf. Remark 2.1.12. However, we might also try an expansion w.r.t. ε, i.e.,
u(t) = u0 (t) + εu1 (t) + O(ε2 ). Plugging this ansatz into (2.12) and sorting
w.r.t. powers in ε yields
O(ε0 ) : ü0 (0) + u0 = 0, u0 (0) = a, u̇0 (0) = 0
⇒ u0 (t) = a cos t,
O(ε1 ) : ü1 + u1 = 2a sin t, u1 (0) = 0, u̇1 (0) = 0
⇒ u1 (t) = −at cos t + a sin t,
and hence uapp1 (t) = a cos t − εta cos t + εa sin t + O(ε2 ). Comparing with
u shows that the expansion only makes sense for t = O(1), and becomes
completely useless after that.
With some physical insight, we may however directly see from (2.12)
that 2ε∂t u corresponds to a weak damping, and hence we suspect that there
are two time scales involved in (2.12). Thus we may try a multi-scale ansatz
of the form
(2.13) u(t) = A(εt)eiω0 t + c.c.,
with ω0 ∈ R an a priori unknown (fast) frequency, and where A = A(τ ) ∈ C
d
is a slowly varying (complex valued) amplitude. Then, e.g, dt u = (iω0 +
d iω t
ε dτ )Ae 0 + c.c., and plugging into (2.12) we obtain
O(ε0 ) : − ω02 + 1 = 0, A(0) = a/2 ⇒ ω0 = 1,
d
O(ε1 ) : 0 = −2i( A + A)eit + c.c..
dτ
This yields A(τ ) = e−τ A(0), and thus
uapp2 (t) = A(τ )eit + c.c. + O(ε) = ae−εt cos(t) + O(ε),
which at least is a much better approximation of the true solution than
uapp1 , see Fig. 2.3.
The equation dτ d
A = −A is called the amplitude equation for the ansatz
(2.13) for the system (2.12), and here can be solved explicitly, like the orig-
inal system. However, already in simple nonlinear ODEs in general neither
the original equation nor the amplitude equation can be solved explicitly.
Moreover, although the amplitude equation is usually a bit “simpler”, this
is not the essential characteristic. The main points are that the amplitude
equation often falls into some universality class, and that it describes the
system on long scales. Thus, if one has to use numerical methods, then the
numerical costs are greatly reduced. For instance, in the present example
we would reduce the numerical costs by a factor 1/ε, e.g., by factor 10 if
ε = 0.1. More drastic cost reductions may occur for PDEs, see Part IV of
this book.
34 2. Basic ODE dynamics
u(t)
u_app1(t)
u_app2(t)
2
-1
-2
0 5 10 15 20 25 30
Figure 2.3. Exact solution and the two approximations for (2.12); ε =
0.1, a = 1.
Proof. Similar to the proof of Theorem 2.1.5 we apply the contraction map-
ping Theorem 2.1.3 to the integrated ODE
t
(2.15) u(t) = u0 + f (u(s), s) ds =: F (u)(t),
t0
where F : M → M with
M = C 0 ([t0 − T0 , t0 + T0 ], {u ∈ Rd : u − u0 Rd ≤ C1 })
which is equipped with the metric
d(u, v) = sup u(t) − v(t)Rd =: u − vM .
t∈[t0 −T0 ,t0 +T0 ]
We have
t
F (u) − u0 M ≤ sup f (u(s), s) dsRd
t∈[t0 −T0 ,t0 +T0 ] t0
t
≤ sup | f (u(s), s)Rd ds| ≤ T0 C3 ≤ C1
t∈[t0 −T0 ,t0 +T0 ] t0
Proof. If u(t)Rd is finite, then the local existence and uniqueness Theorem
2.2.1 applies and also u(t − T0 )Rd and u(t + T0 )Rd are finite for some
T0 > 0.
The following two examples show that f being only continuous is not
sufficient for uniqueness, and that solutions in general do not exist globally.
Example 2.2.4. Consider the one-dimensional ODE u̇ = |u| with initial
value u|t=0 = 0. The right-hand side is not Lipschitz-continuous at u = 0.
This initial value problem has the solution u = 0, but also infinitely many
other solutions, namely
0, for 0 ≤ t ≤ τ
uτ (t) =
(t − τ )2 /4, for τ ≤ t,
solves the ODE for each τ > 0, i.e., there is no uniqueness of solutions.
Example 2.2.5. Consider the one-dimensional ODE u̇ = 1 + u2 with the
initial value u|t=0 = 0. This initial value problem has the solution u(t) =
tan t, i.e., the solution explodes, or blows up, in finite time, it becomes ∞
for t = π/2. Hence, the solution does not exist for all t ∈ [0, ∞), i.e., there
is no global existence of solutions.
Our next goal is to prove the continuity of the solutions w.r.t. the initial
conditions. In order to do so we use Gronwall’s inequality, cf. Lemma 2.1.7.
Lemma 2.2.6. Let f : Rd × R → Rd be continuous and locally Lipschitz-
continuous w.r.t. the first variable. Then each solution u(t, t0 , u0 ) is
Lipschitz-continuous w.r.t. u0 in the following sense: For every T0 > 0
there exist δ > 0 and L > 0 such that for all u1 ∈ Rd with u0 − u1 Rd ≤ δ
we have, for all t ∈ [t0 − T0 , t0 + T0 ],
(2.16) u(t, t0 , u0 ) − u(t, t0 , u1 )Rd ≤ Lu0 − u1 Rd .
Proof. We have
u(t, t0 , u0 ) − u(t, t0 , u1 )Rd
t
≤ u0 − u1 Rd + f (u(τ, t0 , u0 ), τ ) − f (u(τ, t0 , u1 ), τ )Rd dτ
t0
t
≤ u0 − u1 Rd + |C2 u(τ, t0 , u0 ) − u(τ, t0 , u1 )Rd dτ |.
t0
Hence, by Gronwall’s inequality we find
u(t, t0 , u0 ) − u(t, t0 , u1 )Rd ≤ u0 − u1 Rd eC2 |t−t0 | .
2.2. Local existence and uniqueness for nonlinear systems 37
As long as they exist, solutions of ODEs (2.14) have the trivial, but
fundamental property
(2.17) u(t + s, t0 , u0 ) = u(t, s, u(s, t0 , u0 )), u(t0 , t0 , u0 ) = u0 .
For autonomous systems we have
u(t, t0 , u0 ) = u(t − t0 , 0, u0 ) =: u(t − t0 , u0 ),
i.e., w.l.o.g. we can always choose the initial time t0 = 0. Then (2.17)
transfers into
(2.18) u(t + s, u0 ) = u(t, u(s, u0 )), u(0, u0 ) = u0 .
Thus, it makes no difference whether we solve the ODE until the time t + s,
or if we solve the ODE until the time s, start again, and solve until the
time t + s. A similar structure occurs for iterations un+1 = f (un ). In the
following we focus on the autonomous case.
38 2. Basic ODE dynamics
Since with this strict definition, only ODEs (2.14) with solutions existing
globally forward in time in case I = R+ , respectively, forward and backward
in time in case I = R define continuous dynamical systems in the phase
space Rd , we shall not be that strict in the following and call any map u
which fulfills (2.18) a dynamical system.
For ODEs (and PDEs) the dynamical systems property (2.18) expressed
in terms of the family of (nonlinear) solution operators (St )t∈I is given by
(2.19) St+s = St Ss , S0 = I,
where St is defined by St u0 := u(t, u0 ), and where I is here the identity
on M . Due to (2.19), the family of solution operators (St )t∈I is called a
semigroup in case I = R+ .
Remark 2.2.11. Except for the points which are mapped to infinity the
map u0 → u(t, u0 ) is bijective due to u(t, u(−t, u0 ))=u(t − t, u0 )=u0 in case
I = R or I = Z. Since additionally u(t, u0 ) depends continuously on u0
dynamical systems can be interpreted as a flow of homeomorphisms, i.e., as
flow of bijective bi-continuous maps from Rd into Rd . If f is C k then also
u(t, u0 ) is C k w.r.t. u0 , i.e., then the dynamical system can be interpreted
as a flow of C k -diffeomorphisms.
For linear systems the statements of Theorem 2.1.14 remain true with
the more general Definition 2.3.3. The following theorem guarantees in
many situations that stability or instability in the linearized system implies
stability or instability for the full system (2.14).
40 2. Basic ODE dynamics
Theorem 2.3.4. Let u∗ be a fixed point for (2.14). Let A = ∂u f (u∗ ) ∈ Rd×d
be the linearization of f in u∗ .
a) If all eigenvalues λj of A satisfy Re λj < 0, then u∗ is asymptoti-
cally stable.
b) If A has an eigenvalue λ with Re λ > 0, then u∗ is unstable.
Gronwall’s inequality, cf. Lemma 2.2.8, applied to eμ0 t u(t)Rd finally im-
plies eμ0 t u(t)Rd ≤ C0 u0 Rd eC0 bt , respectively
u(t)Rd ≤ C0 u0 Rd e(C0 b−μ0 )t .
2.3. Special solutions 41
Figure 2.5. The phase portrait and the sector in the unstable case
in a typical situation.
system we find
d k k
d d
2R R = (R2 ) = |uj |2 = 2Re uj ∂t uj
dt dt dt
j=1 j=1
k
=2Re uj (λj uj + ajm um + gj ).
j=1 m=j
Using
k
k
2Re uj λj uj ≥ 2σ uj uj = 2σR2
j=1 j=1
and
k
k
k
|2Re uj ajm um | ≤ 2( |ajm |2 )1/2 R2 ≤ 2γR2 ,
j=1 m=j j=1 m=1,m=j
together with
k
|2Re uj gj | ≤ 2Rb ρ2 + R2 ≤ 2Rb(ρ + R)
j=1
yields
d
R ≥ 2σR2 − 2γR2 − 2bR(ρ + R).
2R
dt
Choosing b = σ/10 yields
d
R ≥ σR/2 − bρ.
dt
Similarly, we find
d
ρ ≤ σρ/20 + b(ρ + R).
dt
where we used 2Re dj=k+1 uj λj uj ≤ 0. Since
σR/2 − bρ − σρ/20 − b(ρ + R) ≥ σ(R − ρ)/4
we finally obtain
d
(R − ρ) ≥ σ(R − ρ)/4
dt
and as consequence
R(t) − ρ(t) ≥ (R(0) − ρ(0))eσt/4 .
For solutions with R(0) = 2ρ(0) it follows that R(t) ≥ ρ(0)eσt/4 . However,
this contradicts the assumption of stability, since R(t)+ρ(t) ≤ ε for all t ≥ 0
is not possible, independent of how small ρ(0) > 0 or δ > 0 has been chosen.
2.3. Special solutions 43
The following theorem states that near a hyperbolic fixed point the flow
can be completely linearized by a change of coordinates h. For a proof we
refer to [Tes12, Theorem 9.9].
−1
−1 0 1
Figure 2.6. Flow for u̇1 = −u2 + u1 (1 − u21 − u22 ) and u̇2 = u1 +
u2 (1 − u21 − u22 ).
From the phase portrait we find that all solutions converge towards the
circle r = 1 which is a periodic solution with the minimal period T = 2π.
Moreover, the origin r = 0 is an unstable fixed point in the r equation.
As an exercise, we may consider the linearization around (u1 , u2 ) = 0. We
obtain
u̇1 u1 1 −1
=A with A = ,
u̇2 u2 1 1
with eigenvalues λ1,2 = 1 ± i. Since Re λ1,2 = 1 > 0 we also have with
Theorem 2.3.4 the instability of the origin. On the other hand, from the
phase portrait the periodic solution r = 1 seems to be asymptotically stable.
However, as we see in a moment we have to be more precise when we talk
about stability of periodic solutions.
This fact and the fact that the intersection point u∗ of the periodic
solution is a fixed point of the Poincaré map Π, i.e., Π(u∗ ) = u∗ , lead to the
following definition.
Proof. The proof goes along the lines of the proof of Theorem 2.3.4.
Example 2.3.15. In order to prove the stability of the periodic solution
r = 1 in Example 2.3.11 it remains to compute the Floquet multipliers.
Since r = r − 1 satisfies r˙ = −2
r + O(
r2 ) and since Π(u0 ) = u(2π, u0 ) for
u0 ∈ S due to φ solving φ̇ = 1, we find
DΠ((1, 0)) = e−2·2π ∈ R1×1 .
Thus we have one Floquet multiplier with |e−4π | < 1, which implies the
stability of the periodic solution r = 1.
The following theorem guarantees that the sets Ws and Wu from Defini-
tion 2.3.17 are smooth manifolds and that they are invariant under the flow
of the ODE.
{−1, 1} for k ∈ 2Z + 1
which yields the eigenvalues λ ∈ implying sad-
{−i, i} for k ∈ 2Z
dles for odd
k and centers
for even k. The eigenvectors at the saddles are
1 1
ϕ1 = , ϕ2 = , in the stable direction, and unstable direction,
1 −1
respectively. Physical intuition lets us suspect the existence of, for instance,
heteroclinic orbits
4
f1=0
3 f =0
2
2
−1
−2
−3
−4
−4 −2 0 2 4
The whole interior of the “eye” is filled with periodic solutions. There-
fore, each of them is stable in the sense of Definition 2.3.12 with Floquet
multiplier 1.
Stable, center, and unstable manifolds can also be generalized from fixed
points to more complicated objects, for instance to periodic solutions. They
exist for discrete dynamical systems, too.
Remark 2.3.22. Let u∗ be a fixed point of the iteration un+1 = Πun . The
set
Ws = {us ∈ Rd : ∃ β > 0 : lim Πn (us ) − u∗ Rd eβn = 0}
n→∞
is called the stable manifold of u∗ . The set
Wu = {uu ∈ R : ∃ β > 0 :
d
lim Πn (uu ) − u∗ Rd eβ|n| = 0}
n→−∞
2.4.1. ω-limit sets. Given some initial condition u0 for a dynamical sys-
tem St u0 = u(t, u0 ) on X, the behavior of the solution for t → ∞ is described
by the ω-limit set, defined by
(2.24) ω(u0 ) = {v ∈ X : ∃ (tn )n∈N with tn → ∞ and lim u(tn , u0 ) = v}.
n→∞
Thus, the ω-limit set of u0 consists of all limit points of the forward orbit
through u0 . Hence, an equivalent characterization is
(2.25) ω(u0 ) = ∩t≥0 ∪s≥t u(s, u0 ).
If γ+ = γ+ (u0 ) is the (forward) orbit through u0 and v ∈ γ+ then ω(v) =
ω(u0 ) such that we also write ω(γ+ ) := ω(u0 ).
Theorem 2.4.1. The set ω(γ+ ) is closed and invariant. If X = Rd and γ+
is bounded, then ω(γ+ ) is compact, connected and non-empty. In general,
if ∪t≥t0 St u0 is compact for some t0 ≥ 0, then ω(γ+ ) is compact, connected
and non-empty.
In the general case, i.e., with X some Banach space, the proof works
the same way, with ω(γ+ ) compact as a closed subset of the compact set
∪t≥t0 St u0 . Note that this compactness argument is used in d) and e).
Proof. The fact that ω(B) is nonempty, compact, invariant and connected
follows as in Theorem 2.4.1. To show that A is the maximal compact in-
variant set, let Y be a compact and invariant set. Then St Y = Y and,
since B is absorbing, St Y ⊂ B for t ≥ t0 , hence Y ⊂ B and therefore
ω(Y ) = Y ⊂ A = ω(B). This shows that A is the maximal compact invari-
ant set.
52 2. Basic ODE dynamics
Next we show that A attracts all bounded sets. Suppose that this is not
the case, then there is a bounded set Y , a δ > 0, and a sequence tn → ∞
with
dist(Stn Y, A) ≥ δ.
Thus, there are un ∈ Y with dist(Stn un , A) ≥ δ/2. Since Y is bounded and
B absorbing we have Stn un ∈ B for n large enough, and as B is compact
there is a subsequence with
Theorem 2.4.6. Let A be the global attractor for the flow St and let u0 ∈ X.
For all ε > 0 and T > 0 there exists a τ = τ (ε, T ) > 0 and a point v0 ∈ A
such that
Sτ +t u0 − St v0 X ≤ ε for all 0 ≤ t ≤ T.
Proof. From the continuous dependence on the initial conditions for given
ε, T > 0 there exists a δ = δ(ε, T ) such that
2
6
0 4
1 2
−2 0
0
0 1 0 500 1000 1500 2000
Figure 2.9. Left: Two orbits for (2.28) approaching the circle S :=
{(x, y, z) ∈ R3 : x2 + y 2 = 1} of fixed points from above and below,
respectively. Right: illustration of the notion of pseudo orbits for
(2.28), here consisting of fixed points on S.
-1
-2
-2 -1 0 1 2
Figure 2.10. The “vector field” for Example 2.4.9 for ε > 0.
Then
dist(Aμ , A0 ) → 0 as μ → 0.
Proof. Let ε > 0. Since A0 attracts Q there exists a t > 0 such that St0 Q is
a subset of the ε/2-neighborhood N (A0 , ε/2) of A0 , i.e., St0 Q ⊂ N (A0 , ε/2).
Next, for μ > 0 sufficiently small, we have
sup Stμ u − St0 u ≤ ε/2
u∈Q
This yields that the ω-limit set of an orbit γ intersects every Poincaré
section in only one point. If there is no fixed point in ω(γ), then ω(γ) is a
periodic orbit.
Example 2.4.12. Using the Poincaré-Bendixon theorem allows us to prove
the existence of a periodic solution for
(2.29) ẋ = y and ẏ = −x + y(1 − x2 − 2y 2 ).
A direct consequence of the Poincaré-Bendixon theorem is that a posi-
tively invariant bounded set for ẋ=f (x), x∈R2 , which does not contain
a fixed point, must contain a periodic orbit. The set A = {(x, y) ∈ R2 :
1/4< x2 + y 2 <1} is positively invariant for (2.29). This follows by looking
at the sign of
d
(x(t)2 + y(t)2 ) = 2y 2 (1 − x2 − 2y 2 )
dt
56 2. Basic ODE dynamics
1.5
0.5
−0.5
−1
−1.5
−2
−2 −1 0 1 2
Proof. We have
d d
(2.31) V (u(t)) = (∇V (u(t))) T
u(t) = −∇V (u(t))2Rd < 0
dt dt
except in case that u is a fixed point. Suppose that there exists a non-trivial
periodic solution with u(t) = u(t + T ) with minimal period T > 0. Then by
(2.31) we have V (u(t)) > V (u(t + T )) = V (u(t)), which is a contradiction.
Remark 2.4.14. The linearization A ∈ Rd×d at a fixed point in a gradient
system is a symmetric matrix. Therefore, all eigenvalues are real.
2.4. ω-limit sets and attractors 57
Saddle
Minimum
Minimum
Figure 2.13. The solutions decay along the gradient of the potential.
It turns out that with only a few additional assumptions the ω-limit sets
and the global attractor of gradient systems can be completely described.
Theorem 2.4.15. Suppose that V (u) → ∞ for uRd → ∞, that the set E
of fixed points is finite, and that the fixed points are all hyperbolic. Then,
for all u0 ∈ Rd we have ω(u0 ) = u∗ for some fixed point u∗ . Moreover, for
(2.30) there exists a compact absorbing set B, and the attractor A = ω(B)
consists of finitely many fixed points and the connecting orbits between the
fixed points.
See Exercise 2.18 for an example of a map where conjugacy to the shift
on Σ+2 can be shown explicitly. A related and famous 1D iteration for which
chaotic behaviour can be shown for certain parameters is the logistic map,
see, e.g., [Dev89], and §3.4.1.
V0 V1
H1
H0
Proof. Each horizontal strip Hi is mapped through f into the vertical strip
Vi = f (Hi ). We consider Vi ∩ Hj which is the image of some thin horizontal
strips Hij . We obtain vertical strips Vij = f 2 (Hij ) by two iterations of f .
See Figure 2.15.
V00 V10 V 11 V 01
H 10
H 11
f
f
H
01
H
00
a) b) c)
Σ1
q q
Σ0
R
p p
and assume that these sets are contained in the previous neighborhood.
The solutions go from Σ0 to Σ1 according to Figure 2.16 b). The inner map
ψint : Σ0 → Σ1 , which maps a point a ∈ Σ0 into the first intersection point
of the associated solution with Σ1 , maps vertical vertical lines from Σ0 into
a logarithmic spiral in Σ1 . The outer map ψout transports a neighborhood
of q through the homoclinic solution into
0 = {(x, y, z) : x2 + y 2 = r2 , |z| < z1 }.
Σ 0
See Figure 2.16 c). The map ψ is defined by ψ = ψout ◦ ψint for all points
X ∈ Σ0 with ψ(X) ∈ Σ0 and has the same asymptotic behavior as ψint for
z → 0 since for z → 0 the time needed by the solution to come from Σ0
to Σ1 becomes infinite, whereas the time needed by the solution to come
from Σ1 to Σ0 stays finite. Hence, a rectangular set R around the entrance
point p of the homoclinic orbit is mapped into a spiral like structure. See
Figure 2.17. The assumption | Re ω| < λ is necessary that this picture really
occurs, for more details see [GH83, §6.5.1]. Therefore, graphically we have
found a Smale’s horseshoe for ψ.
64 2. Basic ODE dynamics
ψ(W)
z
W
p θ
2.6. Examples
The following series of examples is intended to give some familiarity with
the notions and ideas introduced so far.
Example 2.6.1. a) For the potential V (x, y) = (x2 − 1)2 + y 2 we find
−∇V (x, y) = −(4x(x2 − 1), 2y), leading to the fixed point (x, y) = (0, 0),
which is a saddle point of V , and to the fixed points (±1, 0), which are
minima of V . For every r > 1 the set {(x, y) : x2 + y 2 ≤ r2 } is absorbing.
The attractor A is given by [−1, 1] × {0}, consisting of the three fixed points
and the heteroclinic connections between the unstable fixed point (0, 0) and
the stable fixed points (±1, 0).
b) For the potential V (x, y) = (x2 − 1)2 + (y 2 − 1)2 we find that any
neighborhood of the unit square Q = [−1, 1] × [−1, 1] is an absorbing set.
The global attractor is Q = W u ((0, 0)).
(a) (b)
1
−1
−1 0 1
1 1 1
0 0 0
−1 −1 −1
−1 0 1 −1 0 1 −1 0 1
As examples consider
(1) u̇ = u(1 − u − v), v̇ = v(u − v),
(2) u̇ = u(1 − u − v), v̇ = v(1 − u − αv),
(3) u̇ = u(2 arctan(2v) − u), v̇ = v(3 arctan(2u) − v).
In (2), α > 0 is some parameter. Clearly, (1)=(PP), (2)=(C), (3)=(S). After
determining the unique fixed points (u, v)∗ with uv > 0 for (1) and (3), the
phase portraits can be conveniently sketched by considering the signs of the
growth rates M, N . For (1) we may additionally use the fact that, e.g.,
[1/4, 3/4] × [1/4, 3/4] is positively invariant.
For (2) we note that for α = 1 we have M = N , and thus a line {u + v =
1} of fixed points. For α = 1 we again have a unique non-trivial fixed point.
In particular, for α > 1 (α < 1) the v species (the u species) dies out. For
α > 1 the biological interpretation is that for u = v the growth rate N of v
is smaller than that of u, due to higher damping (faster saturation) of the
growth of v by itself, hence u “wins”.
0.4 2 M>0
N>0
0.2 M>0 1 M<0
N>0 N>0
0 0
0 0.5 1 0 1 2 3 4 5
Exercises
Exercises 2.11 and 2.12 should be done with some software for ODE phase
portraits, e.g., xppaut or pplane, and we also recommend to use such software for
illustration after the analysis for the other planar ODEs, e.g., in Exercises 2.10,
2.15, 2.19, and 2.20.
2.6. Examples 71
2 3
1.5
2
1
1
0.5
0 0
−0.5
−1
−1
eps=0.1
−2
−1.5 eps=0.5
eps=1
−2 −3
0 10 20 30 −3 −2 −1 0 1 2 3
2 4
0, if u = 0,
a) u̇ = u − u2 , b) u̇ = −u + 4u3 − u5 , c) u̇ =
−u3 sin(1/u), if u = 0.
Find the fixed points and compute their linearization. Which fixed points are stable
and which fixed points are unstable? Sketch the phase portraits.
2.9. Consider f : R3 → R3 , f (u) = Au + g(u) with
Au = (au1 , acu2 , cu3 ), g(u) = (0, acu1 u3 , 0),
−1
where a > 1 > c > a > 0. According to the discrete Hartman-Grobman theorem,
cf. Remark 2.3.9, there exists a homeomorphism h such that h−1 ◦ f ◦ h = A. Show
that h cannot be Lipschitz-continuous.
Hint: Clearly h−1 ◦ f n ◦ h = An . Show that this implies
c−n h2 (u1 , 0, cn u3 ) − an h2 (a−n u1 , 0, u3 ) = nh1 (u1 , 0, cn u3 )h3 (a−n u1 , 0, u3 ).
Next show that h2 (u1 , 0, 0) = 0 and h2 (0, 0, u3 ) = 0, if h is Lipschitz-continuous.
Then obtain a contradiction for n → ∞.
2.10. Consider ẋ = y, ẏ = −cy − x + x3 , with c ∈ R a parameter. Find the fixed
points and compute their linearization. Which fixed points are stable and which
fixed points are unstable? Sketch the phase portrait for different values of c. Find
the stable, the unstable and the center manifolds for the fixed points.
2.11. Find the possible ω-limit sets for ẋ = y, ẏ = x + εy − x3 + 0.1x2 y, with
ε ∈ [−0.09, −0.07]. Compute values of ε ∈ [−0.09, −0.07] where a qualitative
change of the periodic orbits occurs. Hint: Unstable objects can be found by
t → −t.
2.12. Consider
2
ẋ 1 1 x (x + αy 2 )x
= − .
ẏ −1 1 y (αx2 + y 2 )y
Plot the phase portrait for α = 1, 5, 10. Find the fixed points and the periodic
solutions. Which of them are stable? Find the maximal α∗ > 1, such that there
exists a non-trivial stable periodic solution for all α ∈ [1, α∗ ]. (Hint: Consider the
phase portrait for α ∈ (10, 12) by computing the ω-limit set for the initial condition
(x, y) = (0.1, 0.1).) Let α = 12. Find the fixed points and the associated stable
and unstable manifolds.
2.13. Consider u̇ = f (u) with f ∈ C 1 (R2 , R2 ). Let Ω ⊂ R2 be open and simply
connected. Assume the existence of a b ∈ C 1 (R2 , R) with div(bf ) > 0 in Ω. Use
the integral law of Gauss to show the non-existence of a periodic orbit in Ω.
2.14. Use the idea from Example 2.4.12 to prove that ẋ = x−y −x3 , ẏ = x+y −y 3 ;
has a periodic solution.
2.15. Discuss the stability of the fixed point (x, y) = (0, 0) and sketch the phase
portraits for the following systems; compare with ẋ = y, ẏ = −x and explain the
qualitative differences.
a) ẋ = y, ẏ = −x3 . b) ẋ = y 1999 , ẏ = −x1999 .
Hint for a) Consider V (x, y) = αx4 + y 2 with suitable α.
2.6. Examples 73
2.16. Consider ẍ + δ(x)ẋ + 25x = 0 with δ(x) = 8 for |x| > 1 and δ(x) = −6 for
|x| ≤ 1. In order to show the existence of a periodic orbit consider the Poincaré
map G1 : S1 → S2 and G2 : S2 → S3 , where S1 = {(x, ẋ) : x = −1, ẋ ≥ 0},
S2 = {(x, ẋ) : x = 1, ẋ ≥ 0} and S3 = {(x, ẋ) : x = 1, ẋ ≤ 0}. Use then the
symmetry of the problem.
2.17. Let Σ be the set of all 0 − 1 sequences (sj )j∈N with the following property.
If sj = 0, then sj+1 = 1, i.e., Σ consists of all sequences without two succeeding
zeroes. Prove that:
a) the shift σ maps Σ into itself;
b) there exists a dense orbit in Σ ;
c) the set of periodic orbits is dense in Σ .
Dissipative dynamics
In this chapter we provide the strategy and the tools to tackle dissipative
systems, which are characterized by the existence of a compact absorbing
set. In such systems very often through so called bifurcations complicated
and eventually chaotic dynamics occur if some external parameter is varied.
In applications such an external parameter can be for instance an external
heating or the concentration of a chemical substance. A typical scenario is as
follows. For small values of this parameter all solutions are attracted to some
asymptotically stable fixed point. If the value of the parameter is increased
the fixed point becomes unstable. Then more complicated dynamics can
be expected in a neighborhood of the unstable fixed point, for instance
new fixed points or time-periodic solutions may bifurcate, i.e., appear in a
neighborhood of the first unstable fixed point. A further increase of the
external parameter leads to instabilities of the bifurcating solutions. Then
quasi-periodic solutions can occur. The next bifurcation may already lead
to chaotic dynamics.
75
76 3. Dissipative dynamics
3.1. Bifurcations
We present a number of elementary bifurcations and explain how the implicit
function theorem and the Lyapunov-Schmidt reduction can be used to prove
their occurrence in more complicated systems.
ẋ = f (x, μ) = μx − x3 ,
x
x
μ
μ
There are two other elementary bifurcations of fixed points, namely the
transcritical bifurcation and the saddle-node bifurcation.
Example 3.1.2. (Transcritical bifurcation of fixed points) Consider
ẋ = μx − x2 ,
with x = x(t) ∈ R and μ ∈ R. The trivial fixed point x = x∗1 = 0 is
asymptotically stable for μ < 0 and unstable for μ > 0. For μ = 0 a
real eigenvalue crosses the imaginary axis. There exists another fixed point
x∗2 = μ, which coincides with the trivial solution x∗1 = 0 for μ = 0. Since
in general we know a priori only the trivial solution, we say that the fixed
point x∗2 = μ bifurcates from the trivial solution x∗1 = 0. For the transcritical
bifurcation an exchange of stability takes place: for μ < 0, x∗1 = 0 is stable
and x∗2 = μ is unstable; for μ > 0, x∗1 = 0 is unstable and x∗2 = μ is stable.
See Figure 3.3.
x
Im Im
Re μ
Re
μ <μ 0 μ>μ0
Im Im 1
Re −1
Re
−1 0 1
Figure 3.5. Two complex conjugate eigenvalues cross the imagi-
nary axis and the phase portrait for (3.1) for μ > 0.
In §3.3 we shall see that this bifurcation occurs generically when a fixed
point loses stability due to two complex conjugate eigenvalues crossing the
imaginary axis.
3.1. Bifurcations 79
|| r ||
of (x0 , μ0 ) with the same argument until the assumption ∂x f (x(μ), μ)) = 0
is no longer satisfied. In such a point (x0 , μ0 ) a new branch of solutions can
bifurcate from this family of solutions η → (x, μ)(η). This so called bifur-
cation point can be analyzed with the Newton polygon which is explained
subsequently. It turns out that generically only two situations for the bifur-
cations of fixed points can occur, namely the transcritical bifurcation from
Example 3.1.2 or the pitchfork bifurcation from Example 3.1.1.
Scaling arguments. One way to establish the existence of bifurcating
solutions in the general case are scaling arguments and the implicit function
theorem.
Example 3.1.6. Let f (x, μ) = μx + x2 + sin x. Then f (0, μ) = 0 for all
μ ∈ R. Hence, x = x∗1 = 0 is the trivial solution for all μ ∈ R. For all values
of μ ∈ R we have ∂μ f |x=0 = 0. Hence, it is sufficient to consider
∂x f |x=0 = (μ − 2x + cos x)|x=0 = μ + 1.
Thus, a bifurcation can only take place when μ + 1 = 0. Therefore, we
introduce the small bifurcation parameter α = μ + 1. In order to find
Example 3.1.2 in the present problem we rescale x = αy and introduce
F (y, α) = α−2 f (αy, 1 + α) = y + y 2 + O(α).
Thus, for α = 0 we have the simple equation F (y, 0) = y + y 2 having the
solutions y1∗ = 0 and y2∗ = −1. According to
∂y F |(y,α)=(yj∗ ,0) = (1 + 2y)|(y,α)=(yj∗,0) = 0
We have the trivial solution (0, 0) for which the assumptions of the implicit
function theorem are not satisfied, i.e., ∂x f (0, 0) = ∂ε f (0, 0) = 0. Again we
are interested in non-trivial solutions x = x(ε) near the origin.
We make the ansatz x(ε) = εv(ε) and obtain
F (v, ε) = ε−2 f (εv, ε) = v 2 + v + ε = 0.
For ε = 0 we find the non-trivial solution v1∗ = −1. We additionally have
∂v F (−1, 0) = −1 = 0 such that we can apply the implicit function theorem
and obtain a smooth solution v = v1∗ (ε) = −1 + O(ε). Hence, we find
a non-trivial solution x∗1 = −ε + O(ε2 ) for f = 0. However, the ansatz
x(ε) = ε2 v(ε) yields
F (v, ε) = ε−3 f (ε2 v, ε) = O(ε) + v + 1 = 0.
For ε = 0, we find the non-trivial solution v2∗ = −1 and ∂v F (−1, 0) = 1 = 0.
Hence, we can apply the implicit function theorem and obtain a smooth
solution v = v2∗ (ε) = −1 + O(ε). Therefore, we found a second curve of non-
trivial solutions x∗2 = −ε2 +O(ε3 ) for f = 0. The expansions correspond to
the solutions
ε ε2
x1,2 (ε) = − ± − ε3
2 4
of (3.3), which only can be computed explicitly since (3.3) is a second order
polynomial w.r.t. x.
The Newton-polygon. In the last example there exist at least two curves
of non-trivial solutions. Since we have a polynomial in the example we can
be sure that we found all solutions. For non-polynomial problems the scaling
argument can be made rigorous with the help of the Weierstrass preparation
theorem which allows to bring analytic f into a polynomial form w.r.t. one
of the variables, cf. [CH82, §2.8]. With this preparation it is then clear
that the solutions which we will find with the scaling arguments are the only
non-trivial solutions near (x, ε) = (0, 0).
A systematic approach to find the scalings is as follows. Assume that
f can be expanded
∞ in some convergent power series near the origin, i.e.,
mn ∈ R is nonzero,
f (x, ε) = a x m εn . Whenever the coefficient a
m,n mn
make a dot at (m, n) ∈ N0 × N0 . Then take the lower convex hull of all dots
in the N0 × N0 -plane. This hull is the so called Newton-polygon with finitely
many line segments with endpoints (mi , ni ) and (mi+1 , ni+1 ) and slopes −αi .
Associated with each of these lines there are pi solutions x∗i (ε) = εαi vi∗ (ε)
of f (x, ε) = 0, where pi = mi − mi−1 .
Example 3.1.9. The Newton polygon for
f (x, ε) = x3 + 3x2 ε + 2xε2 + ε5 = 0
82 3. Dissipative dynamics
3 5
2
3
2
1
0 0
0 1 2 0 1 2 3
Figure 3.7. Newton polygons for f (x, ε) = x2 +xε+ε3 and for f (x, ε) =
x3 + 3x2 ε + 2xε2 + ε5 .
−1 −1
and hence s = −g01 g20 such that μ = −g01 g20 x2 +O(x3 ), i.e., depending on
−1
the sign of g01 g20 a sub- or a supercritical pitchfork bifurcation occurs. More
coefficients can vanish, but this is a degenerated situation, which requires
more symmetries.
By a small perturbation the vanishing coefficients can be made non-
zero. This is called unfolding of the bifurcation, cf. [GS85, Chapter III].
Such unfoldings are robust w.r.t. other small perturbations, i.e., additional
parameters different from the unfolding parameters will not change the so-
lution set qualitatively.
Dividing this equation by x gives ε+x2 +O ε2 x4 = 0 which can be analyzed
by the Newton polygon. We find √ a subcritical pitchfork bifurcation, i.e.,
non-trivial solutions x∗ (ε) = ± −ε + O(ε) and y ∗ (ε) = O(ε) for ε < 0.
(a) (b)
Figure 3.9. When a fixed point becomes unstable the bifurcating so-
lutions can be found on the exponentially attracting center manifold.
(a),(b) before and after the bifurcation of a stable periodic orbit on Mc .
Some parts of the theorem, namely the invariance, the existence, and
Lipschitz-continuity instead of r-times differentiability of the center manifold
will be proven in §13.1. Here we will discuss some of the assertions of the
theorem and concentrate on its application by giving a number of examples.
With the first example we explain how center manifold theory can be used
to handle bifurcation problems although the central eigenvalues are only on
the imaginary axis for one particular value of the bifurcation parameter
Example 3.2.2. For μ close to zero consider the trivial decoupled system
(3.6) ẋ = μx − x3 , ẏ = −y.
3.2. Center manifold theory 87
For μ < 0, the origin (x, y) = (0, 0) is stable. For μ = 0, we find the one-
dimensional center manifold Wc = {(x, y) ∈ R2 : y = 0}. In order to handle
non-zero values of μ with the center manifold theorem the above system is
extended to
ẋ = μx − x3 , ẏ = −y, μ̇ = 0.
For this extended system we find the two-dimensional center manifold Wc =
{(μ, x, y) ∈ R3 : y = 0}. Note that after introducing μ̇ = 0 the term μx is
no longer a linear, but a nonlinear term. Since μ̇ = 0 implies that μ is a
constant, the two-dimensional center manifold is foliated by one-dimensional
invariant manifolds. See Figure 3.10. Hence, the additional equation μ̇ = 0
can be canceled again and on the two-dimensional center manifold μ can be
considered again as a parameter. Therefore, by applying the center manifold
theorem in a sloppy way the two-dimensional bifurcation problem (3.6) can
be reduced to
ẋ = μx − x3
in the one-dimensional center manifold Wc = {(x, y) ∈ R2 : y = 0}. Obvi-
ously, the reduction is trivial in this case, i.e., h = 0. In summary, bifurcation
problems can be handled with the help of the center manifold theorem by
introducing the equation μ̇ = 0.
y
μ
x x
The following two examples are about the non-uniqueness and the non-
smoothness of center manifolds.
Example 3.2.4. In order to illustrate the non-uniqueness of the center
manifold we consider
(3.7) ẋ = x2 , ẏ = −y.
Obviously, the central subspace Ec is given by the x-axis. The solutions of
the ODEs are given by x(t) = 1−tx x0
0
and y(t) = y0 e−t . Elimination of time
t yields y(x) = (y0 e−1/x0 )e1/x . For x < 0 we have limx→0,x<0 y (n) (x) = 0,
i.e., every solution approaches the origin in a flat way. For x > 0 we find
that y = 0 is the only solution which approaches the origin. Thus, we find
infinitely many different C ∞ -center manifolds which are tangential to Ec at
the origin by glueing together the obits in the left half plane with the positive
real axis. This shows that center manifolds are non-unique in general. The
only analytic center manifold, i.e., with a convergent power series, is the
x-axis. See Figure 3.11 for the phase portrait of (3.7).
0.5
−0.5
−1
−1 −0.5 0 0.5 1
(3.8) ẋ = ε + x2 + y 2 and ẏ = −y + x2
with small ε. For ε = 0 we have the fixed point (x, y) = (0, 0) with eigenval-
ues 0, −1 with neutral direction (1, 0). Thus we expand the center manifold
as y = h(x, ε) = ax2 + bxε + cε2 + . . ., which yields a = 1, b = −2, c = 2
and hence ẋ = ε + x2 + O(x4 ) as reduced
√ equation. We
√ have a saddle-node
bifurcation with two fixed points − −ε (stable) and −ε (saddle) for ε < 0
and no fixed point for ε > 0. Thus we do not actually need h. See Figure
3.12 for the phase portrait.
0.4
y=x2+0.2x+0.01
0.3
0.2
0.1
−0.1
−0.5 0 0.5
As already said, some parts of the center manifold theorem, cf. Theorem
3.2.1, namely the invariance, the existence, and the Lipschitz-continuity of
the center manifold will be proven in §13.1. There we will explain that center
manifold theory is not restricted at all to the finite-dimensional situation.
In Part IV it is used for the construction of bifurcating spatially periodic
solutions of pattern forming systems, but also in the construction of traveling
wave solutions in unbounded cylindrical domains.
3.3. The Hopf bifurcation 91
Proof. For the somewhat lengthy proof we introduce the new bifurcation
parameter ε = μ − μ0 and extend the ODE system with ε̇ = 0. Then we
apply the center manifold theorem and reduce the full system to a system
on the three-dimensional center manifold associated to the eigenvalues λ±
and the variable ε.
On the center manifold Mc for arbitrary coordinates (y, z) ∈ R2 the
reduced system can be written as
ẏ = a11 y+a12 z+a120 y 2 +a111 yz+a102 z 2 +a130 y 3 +a121 y 2 z
+a112 yz 2 +a003 z 3 +O(ε2 (|x|+|y|)+|y|4 +|z|4 ),
(3.10) ż = a21 y+a22 z+a220 y 2 +a211 yz+a202 z 2 +a230 y 3 +a221 y 2 z
+a212 yz 2 +a203 z 3 +O(ε2 (|x|+|y|)+|y|4 +|z|4 ),
ε̇ = 0,
where the values of the real-valued coefficients a· = a· (ε) depends on our
a11 a12
choice of basis. The only restriction so far is that the matrix A =
a21 a22
92 3. Dissipative dynamics
ẋ = Ax + f (x)
Πε exists for ε > 0, too. Associated to this fixed point are periodic solutions
of (3.13), (3.11), and finally of (3.10). Therefore, we are done.
q
Hopf line
sn sf uf un
p
saddle points
The transfer of Figure 3.13 to the fixed point (α0 , β0 ) and to the pa-
rameters μ, κ yields to the solution of a number of algebraic equations. For
instance, the Hopf line is given by solving the 4th order equation
trace A = −(μ4 − (1 − 2κ)μ2 + κ/(1 + κ))/(μ2 + κ) = 0,
hence
1
1/2
μ1,2 (κ) = √ (1 − 2κ) ± (1 − 8κ)1/2 .
2
In summary, we obtain the bifurcation diagram plotted in Figure 3.14, while
Figure 3.15 shows two selected phase portraits. We will come back to such
systems in Chapter 9.
1/4
sn
sf
uf
un
1 2 μ
Hopf line
f1=0 2.5
2 f =0
1
f2=0 f =0
2 2
1.5
1.5
1
1
0.5 0.5
0
0 0 0.5 1 1.5 2 2.5
0 0.5 1 1.5
Figure 3.15. Phase portraits for (μ, κ) = (0.5, 0.2) (left) and (μ, κ) =
(0.5, 0.1) (right)
μ = μ(x), i.e., we have a smooth curve (x, μ) = (x, μ)(s) parameterized with
s until simultaneously ∂x f (x(s), μ(s)) = ∂μ f (x(s), μ(s)) = 0.
There are variants of this idea [Kel77, Kuz04, Doe07, Sey10] which
automatically allow the continuation of branches around folds and beyond
bifurcation points, the detection and localization of bifurcation points, and
branch switching at bifurcation points. One standard method is so called
(pseudo-)arclength continuation, which is implemented in the package Auto,
[Doe07, Dea16], see also xppaut, [Erm02]. Many of these methods can
also be applied to bifurcation problems in PDEs and are important tools
there. A recent package specifically designed for elliptic systems in two
space dimensions is pde2path, [UWR14, DRUW14].
98 3. Dissipative dynamics
−1. The fixed point becomes also unstable for the second iterate of the
Poincaré map Π2 , but now via a real Floquet multiplier crossing the unit
circle at 1. On the center manifold associated to the Floquet multiplier −1
we have the following situation. If
Πx = −x + αx2 + O(x3 ), then Π2 x = x + βx3 + O(x4 ),
such that for Π2 a pitchfork bifurcation occurs. The two stable bifurcating
fixed points for Π2 corresponds to a 2-periodic solution for Π itself, since for
Π no bifurcation of fixed points occurs. Hence, a new periodic orbit with
twice the period is bifurcating from the old one. Assuming that this new
periodic orbit becomes unstable in the same way and that this procedure
goes on and on we finally come to chaotic dynamics. A famous ODE example
showing this behavior is by Rössler [Rös76].
Example 3.4.1. The Rössler system. Consider the ODE
⎛ ⎞ ⎛ ⎞
x −(y + z)
∂t ⎝y ⎠ = ⎝ x + by ⎠ ,
z b + z(x − a)
where typically b = 0.1 and a ∈ R serves as a bifurcation parameter. Starting
from an asymptotically stable periodic orbit for a=4 we find for increasing a
a period-doubling sequence, cf. Figure 3.17. See, e.g., [PJS92, §3.3] and the
references therein for a more detailed introduction to the Rössler system.
3.4.2. The logistic map. There is a discrete model problem for the period-
doubling route to chaos, namely the logistic map
xn+1 = μxn (1 − xn ) = F (xn )
with μ ≥ 0 and xn ∈ R. We have for the nth iteration F n (x) → −∞ for
n → ∞ if x < 0 or x > 1. For μ ∈ [0, 4] the map F maps the interval [0, 1]
into itself. In the following we restrict ourselves to values μ and x0 in these
sets. More details can be found in [Dev89], including a discussion of chaos
in the strict
√ sense of Definition 2.5.2 in the logistic map for parameter values
μ > 2 + 5.
The condition F (x) = μx(1 − x) = x gives the fixed points x∗1 = 0 and
x∗2 = 1 − 1/μ. At μ0 = 1 a transcritical bifurcation of fixed points occurs.
The linearization around the fixed point x∗ is given by yn+1 = F (x∗ )yn
where F (x∗ ) = μ(1 − 2x∗ ). For x∗ = 1 − 1/μ we obtain
F (x∗ ) = μ(1 − 2(1 − 1/μ)) = 2 − μ.
Hence, this fixed point is stable for μ ∈ (1, 3) and becomes unstable at
μ1 = 3 via some period-doubling. A stable two-periodic solution appears.
See Figure 3.18.
100 3. Dissipative dynamics
(a) a = 4 (b) a = 6
2.5
6
2
1.5 4
z
z
1
0.5 2
5
5 5 10
0 0 5
0 −5 0
−5 −5 −5
y x y x
15
12
10
8 10
z
6
4 5
2
10 10
0 10 0 10
0 0
−10 −10 −10 −10
y x y x
n
value μ = μn . We find a pitchfork bifurcation of F 2 and so the occurrence
of a 2n -periodic solution for F .
Even more interestingly, the period-doublings show some asymptotic
behavior. It can be proved rigorously by computer-assisted proofs that the
limit
μn − μn−1
lim ≈ 4.6692,
n→∞ μn+1 − μn
called the Feigenbaum constant, exists, cf. [CE80]. As a consequence we
have the existence of μ∞ = limn→∞ μn ≈ 3.57, too. A recent overview about
the theoretical background of the occurrence of these limits is [Avi11]. For
most values μ > μ∞ the system exhibits chaotic behavior. In Figure 3.19
the ω-limit set for starting point x0 = 1/2 is plotted as a function over the
bifurcation parameter μ. There are isolated regions on the μ-axis where
no attractive chaotic
√ behavior occurs, the so called windows of stability.
Beginning at 1+ 8 ≈ 3.83 there is for instance a range of parameters μ with
a stable 3-periodic solution. There is a general theory [Dev89, §1.10] that
for maps from R to R solutions of period 3 imply the existence of periodic
solutions of every period m ∈ N, known as the Theorem of Sarkovskii.
Figure 3.19. The ω-limit set for starting point x0 = 1/2 is plotted as a
function over the bifurcation parameter μ. For every fixed μ the iterates
xN , . . . xN +M with N and M sufficiently large are plotted.
15
10
z
5 10
0
−10 −5 0 5 −10
10 y
x
13.91 where the two parts S1 and S2 are connected with the two-dimensional
stable manifold of the origin and form two homoclinic connections, see Figure
3.20(b). For ρ < ρglobal the part S1 connects to X1∗ and S2 to X2∗ , see Figure
3.20(a), and vice versa for ρ > ρglob . This behavior is the origin of a so called
homoclinic explosion, cf. [Wig88], and creates the chaotic behavior in the
system. Therefore, chaotic behavior is present in the system already for ρ
close to ρglobal , but becomes only attractive for larger values of ρ. Figures
3.20(c) and (d)) illustrate the different behavior for ρ = 24 < ρHopf and
ρ = 27 (Figure 3.20(d)).
An elementary but more detailed introduction to the Lorenz system
can be found in [Str94, §9], including explanations of simple mechanical
and electronic systems able to simulate the Lorenz system, together with
applications to send encrypted messages.
Exercises
3.1. For the following ODEs ẋ=f (x) determine all fixed points and their stability
in dependence of the parameter μ∈R. What bifurcations occur at what μ?
a) ẋ=μ + 6 + 4x − x2 , b) ẋ = 2 − μ + x(μ−4) + 3x2 − x3 , c) ẋ=μ + x(μ−1) − x2 .
3.2. Compute the non-trivial solutions close to the origin for
εx − yx2 − x4
f (x, y, ε) = = 0.
y + x2
3.3. Check the stability of (x, y) = (0, 0) for the following systems by calculating
the center manifold, cf. [Wig03].
(a) ẋ = −xy − x6 , ẏ = −y + x2 , (b) ẋ = x2 y − x5 , ẏ = −y + x2 , and
xn+1 0 1 xn 0
(c) = + .
yn+1 − 12 23 yn −yn3
3.4. Which of the following systems has periodic orbits close to z = (x, y) = (0, 0)?
Does Theorem 3.3.1 apply?
2
−μ 1
2 z − |z| z,
2
(a) ż =
−12 −μ
−μ 1
(b) ż = z − |z|2 z, μ ∈ R, with small |μ|.
−1 −μ
2
0 −1 x − xy
(c) ż = + .
1 0 xy
3.5. This exercise (from [Str94]) brings together a number of concepts treated
above, namely bifurcation, center manifold calculations, and non-trivial gradient
dynamics. The system
ϑ̇1 = f1 (ϑ) := k sin(ϑ1 − ϑ2 ) − sin ϑ1 ,
ϑ̇2 = f2 (ϑ) := k sin(ϑ2 − ϑ1 ) − sin ϑ2 ,
with ϑ = (ϑ1 , ϑ2 ) and parameter k > 0 describes two rotating magnets between
two fixed magnets, in a geometry as follows:
3.4. Routes to chaos 105
a) The system has exactly nine fixed points in [−π, π]2 for k < 1/2. Determine
these and their stability.
b) What bifurcation occurs at k = 1/2? What new fixed points emerge?
c) Find a potential V such that ϑ̇ = −∇V (ϑ).
d) Sketch a phase portrait for 0 < k < 1/2 and a phase portrait for k > 1/2.
Remark. Doing c) first and then the rest is a good idea.
3.6. The Chemostat. The chemostat is an industrially used “predator-prey sys-
tem” to cultivate bacteria. In case of 3 species the system is modeled by
m1 s(t)
ṡ(t) = (1 − s(t)) − x1 (t),
a1 + s(t)
m1 s(t) m2 x2 (t)
ẋ1 (t) = x1 (t) −1− ,
a1 + s(t) a2 + x1 (t)
m2 x1 (t) m3 x3 (t)
ẋ2 (t) = x2 (t) −1− ,
a2 + x1 (t) a3 + x2 (t)
m3 x2 (t)
ẋ3 (t) = x3 (t) −1 .
a3 + x2 (t)
a) Explain the modeling.
b) Let σ(t) = 1 − s(t) − 3k=1 xk (t). Show that σ̇(t) = −σ(t) and use this to prove
that the ω-limit set of any solution (s(t), x1 (t), . . . , x3 (t)) is contained in
3
Ω = {(s, x1 , . . . , x3 ) : s + xk = 1}.
k=1
3
c) Substitute s = 1 − k=1 xk into the equations for ẋk and try to reproduce the
period-doubling shown in Figure 3.21.
3.7. Let Qc (x) = x2 + c. Prove that for all c < 14 there exists a unique μ > 1 such
that Qc is conjugated to Fμ (x) = μx(1 − x) through the map h(x) = αx + β.
3.8. Consider the iteration xn+1 = Tλ (xn ), where
$
2x for x ∈ [0, 1/2],
Tλ (x) = λ
2 − 2x, for x ∈ (1/2, 1].
a) Prove, x∗ = 0 is an asymptotically stable fixed point for λ ≥ 0 sufficiently
small.
b) Compute Tλ2 = Tλ ◦ Tλ and find graphically the 2-periodic solutions,
i.e., solve Tλ2 (x) = x by finding the intersection points of the functions
x → Tλ2 (x) and x → x.
106 3. Dissipative dynamics
0.5
0.4 0.4 0.4
0.3 0.3
3
0.3
x
x
0.2 0.2
0.2 0.1
0.1
0.6
0.4 0.6
0.4 0.8 0.8
0.3 0.6 0.6 0.4 0.6
0.2 0.4 0.2 0.4 0.2 0.4
0.1 0.2 0.2
0.2 x x x x
2 1 2 1
0.4 0.4
0.3 0.3
Hamiltonian dynamics
109
110 4. Hamiltonian dynamics
defined by −F (x6 ) = −F (x5 ), where all kinetic energy has been transformed
to potential energy again. The ball will now roll back, and all together
we obtain a periodic orbit γ1 . Similar periodic orbits are obtained for all
starting positions (x, 0) with x2 < x < x3 , or, equivalently for all (x, ẋ)
inside the region bounded by the homoclinic orbit γ2 to (x3 , 0) and passing
through (x7 , 0). In a similar way all orbits can be constructed graphically.
For instance, the orbit γ3 corresponds to a ball coming from the far left with
some large positive speed and rolling through the potential all the way to
a position x8 on the (far) right where it reaches some maximal potential
energy −F (x8 ) and then rolls back.
−F(x)
f(x)
0
γ3
4
γ1 2
γ2
0
-2
0 1
x1 x2 x3 x4 -2 -1
x7 x5 x6 -4 -4 -3
The Jacobi matrix Y (t) = Du0 u(t, u0 ) satisfies the differential equation
d
Y (t) = ((D(J∇H))(u(t, u0 )))Y (t).
dt
Furthermore, the determinant |Y (t)| satisfies
d
|Y (t)| = trace(D(J∇H))|Y (t)|.
dt
See Exercise 4.1 for a proof of this formula in R2 . Using Y (0) = I, and
2d
2d
2d
2d
2d
trace(DJ∇H) = δik ∂ uk Jij ∂uj H = Jij ∂ui ∂uj H = 0
i=1 k=1 j=1 i=1 j=1
due to the skew symmetry of J and due to the symmetry of the matrix
(∂ui ∂uj H)i,j , we obtain the assertion.
The invariant Lebesgue measure in phase space is called Liouville mea-
sure. The theory of measure preserving dynamical systems is the subject
of ergodic theory, see, e.g., [Wal82, Kre85]. Complicated dynamical be-
haviour in Hamiltonian systems is described statistically in this theory.
Then
1 1
2d 2d
∂ uk H = (mij δik uj + mij ui δjk ) = (mik + mki )ui ,
2 2
i,j=1 i=1
For general ODEs in case that all eigenvalues lie on the imaginary axis
the nonlinear terms decide about stability. For Hamiltonian systems the
quadratic approximation of the Hamiltonian at the fixed point gives addi-
tional information.
Theorem 4.1.6. Let H(u) = 12 uT Au + O(u3 ) with A strictly positive (or
strictly negative) definite. Then u = 0 is stable.
Proof. Let
ρ0 (r) = min{H(u) : |u| = r}, ρ1 (r) = max{H(u) : |u| = r}.
Then
1 1
ρ0 (r) = λmin r2 + O(r3 ), ρ1 (r) = λmax r2 + O(r3 ),
2 2
where λmin > 0, respectively λmax > 0, is the smallest, respectively the
largest eigenvalue, of the positive definite matrix A. Then there exists an
r0 > 0, such that
1
ρ0 (r) ≥ λmin r2 and ρ1 (r) ≤ λmax r2 .
4
for all r ∈ [0, r0 ]. Given ε > 0 we choose
λmin
0≤δ≤ min(ε, r0 ).
4λmax
Then
4 4
|u(t, u0 )|2 ≤ ρ0 (u(t, u0 )) ≤ H(u(t, u0 ))
λmin λmin
4 4 4λmax 4λmax 2
= H(u0 ) ≤ ρ1 (u0 ) ≤ |u0 |2 ≤ δ ≤ ε2 .
λmin λmin λmin λmin
114 4. Hamiltonian dynamics
2π
e ω JA − I. Due to the non-resonance assumption, Dx∗ F (0, 0) has only non-
zero eigenvalues and hence is invertible. Then, by the implicit function
theorem, F ( x∗ , ε) = 0 can be solved for x
∗ = x
∗ (ε). The associated solution
) is non-trivial since I = 0. The period is
(I, φ, x
T 2π 2π 2π
dφ dφ dφ 2π
dt = = = = + O(ε).
0 0 φ̇ 0 ∂I H 0 ω + O(ε) ω
This family of periodic solutions is tangential to span{I, φ} since
x =
O(ε).
Then
cos φ − sin φ
p := q̇ = ṙ + rφ̇ ,
sin φ cos φ
and C and H in polar coordinates become
1 (r).
(4.7) C = r2 φ̇ = 0, H = (ṙ2 + r2 φ̇2 ) + U
2
From (4.7) we obtain φ̇ = 0 and ṙ2 = 2H − r2 φ̇2 − 2U (r). This yields a scalar
first order equation as follows. Locally we can assume r = r(φ); then
d d d ṙ
ṙ = r(φ) = ( r)φ̇ ⇒ r := r= .
dt dφ dφ φ̇
In particular
2(H − U ) 2(H − U )r4
(r )2 = − r2 = − r2 ,
φ̇2 C2
which yields the Clairaut ODE
r = ±g(r), with g(r) = 2C −2 (H − U (r))r4 − r2 .
Instead of r we use the inverse radius σ=1/r, which yields the so called
fundamental equation of the 1-body problem
σ = g(1/σ)σ 2 .
For the gravitational potential U (r) = −Ar−1 we obtain
(4.8) σ = − −σ 2 + ασ + β, with α = 2A/C 2 , β = 2H/C 2 .
Lemma 4.2.1. a) We have H ≥ −A2 /(2C 2 ) (lower energy bound).
b) For H = −A2 /(2C 2 ) the orbit is a circle with radius C 2 /A.
√ √
Let σ(0) = σ0 with max{0, α−2 δ } < σ0 < α+2 δ . Then we have a local
√
solution as g is locally Lipschitz near σ0 . Substituting u = 12 δx + α2 we
obtain
σ σ σ
du du dx
φ= = = √
σ0 −g(u) σ0 − −u + αu + β
2 0 − 1 − x
σ
2
− arccos σ
= arccos σ 0 ,
0 = √2 (σ0 − α2 ) and σ
with σ = √2 (σ − α2 ). From −1 < σ
0 < 1 we obtain
δ δ
σ0 ) < π and thus
0 < arccos(
2
(4.10) √ (σ − α/2) = σ = cos(φ + b) with b = arccos(σ0 ).
σ
W.l.o.g. we choose the initial condition σ0 = 1/rmin and obtain
p
(4.11) r(φ) = with p = C 2 /A, e = 1 + 2HC 2 /A2 .
1 + e cos(φ)
We distinguish three cases.
(1) 0 ≤ e < 1 ⇔ H < 0: Then r(φ) is defined for all φ ∈ R and
2π-periodic in φ. Going back to cartesian coordinates we obtain
x q1 cos φ
= =r ,
y q2 sin φ
and cos φ = (p/r − 1)/e yields (p − ex)2 = r2 = x2 + y 2 , and thus
2
ep y2 p2
x+ + = .
1 − e2 1 − e2 (1 − e2 )2
This is Kepler’s first law: The body moves on an ellipse with focal points
(0, 0) and (−(2ep)/(1 − e2 ), 0), numerical eccentricity e and major semi-axis
a = −A/(2H). The point (rmin , 0) is called peri center (perihel for a planet
in the solar system) and (rmax , π) is called apo center (apohel). Examples for
numerical eccentricities e are e = 0.0167 for Earth, e = 0.2056 for Mercury,
and e = 0.9673 for Halley’s comet. The relatively large eccentricity of
Mercury is of great importance historically since already in the 19th century
it allowed the observation of the perihel precession of Mercury: after each
elliptical orbit Mercury’s perihel is shifted by a few angular seconds. This
contradicts the above (newtonian) calculations, but could be explained by
Einstein’s relativity theory.
(2) e = 1 ⇔ H = 0: the existence interval is −π < φ < π, and
geometrically the orbit is a parabola opening to the left, y 2 = −px + p2 .
(3) e > 1 ⇔ H > 0: the orbit is the hyperbola y 2 = (e2 −1)x2 −2epx+p2 .
Thus we found the orbits for the 1-body problem in implicit form and
without time dependence, determined by parameters H, A and C. Next, the
orbits can be characterized via initial conditions and the time-dependence
4.2. Some celestial mechanics 119
F2 F2
The solution of the associated differential equations, and the associated ques-
tion about the mechanical stability of our solar system, have been considered
as essential for mankind. However, it turned out that only the two body
problem (N = 2), see above and Exercise 4.9, can be solved explicitly and
already the three body problem shows chaotic behavior.
There is one intermediate problem, namely the so called restricted three
body problem. There it is assumed that the third body K3 has a very small
mass compared to the other two bodies K1 and K2 . The restricted three
body problem is obtained by neglecting the forces of K3 on K1 and K2 , such
that their motion is not affected by K3 , i.e., they move on Kepler ellipses
around their center of mass.
In a coordinate system which moves with the two larger bodies of reduced
masses μ = m1m+m1
2
and 1−μ, which lie fixed in −μ and 1−μ, the Hamiltonian
for the motion of the third body is given by
1
H(q, p) = (p21 + p22 + p23 ) + q2 p1 − q1 p2 + U (q),
2
where
q2 q2 1−μ μ
U (q) = − 1 − 2 − − .
2 2 2 2
(q1 + μ) + q2 + q3 2 (q1 − 1 + μ)2 + q22 + q32
120 4. Hamiltonian dynamics
The second and third term in H and the first term in U come from the
Coriolis force in the rotating coordinate system. There are five equilibria,
called Lagrangian points, shown in Figure 4.3.
q2
P4
P1 −μ P2 1−μ P3 q1
P
5
For the linearisation JM in these points (especially ∂qj ∂q3 U |Pi = 0 for j =
1, 2 and i = 1, . . . , 5) we thus obtain
⎛ ⎞
0 1 0 1 0 0
⎜ −1 0 0 0 1 0 ⎟
⎜ ⎟
⎜ 0 0 0 0 0 1 ⎟
JM = ⎜ ⎜ ⎟.
⎜ −∂q1 U
2 −∂q1 ∂q2 U 0 0 1 0 ⎟ ⎟
⎝ −∂q1 ∂q2 U −∂q22 U 0 −1 0 0 ⎠
0 0 −∂q23 U 0 0 0
We find that the q3 , p3 -part decouples and leads to the eigenvalue problem
0 = λ2 + ∂q23 U.
Since ∂q23 U > 0 we have λ1,2 ∈ iR. For the remaining eigenvalues we have
It turns out that the points P1 , P2 and P3 are saddles and therefore unstable.
In the points P4 and P5 we find
√ √
1 3 2 3 2 9 3 3 1
q1 = − μ, q2 = ± , ∂q1 U = − , ∂q2 U = − , (∂q1 ∂q2 U ) = −
2
( − μ).
2 2 4 4 2 2
For 4( 12 −μ)2 < 1 the eigenvalues are purely imaginary, i.e., the points P4 and
P5 are linearly stable. Unfortunately M is indefinite, such that we cannot
conclude on the nonlinear stability of P4 and P5 with the above theorem.
Nevertheless P4 and P5 are realized in nature and play an important role
for space missions. For instance, Sun and Jupiter can be taken as the big
bodies, and in an angle of 60 degrees before and after Jupiter on his orbit
there are the so the called Greeks and Trojans, some families of asteroids.
Clearly this system is the direct sum of d Hamiltonian systems with d inde-
ω
pendent Hamiltonians Hj = 2j (qj2 + p2j ). The Ij = 2ω1 j Hj are conserved also
d
for the flow of (4.12), i.e., dt Ij (x(t)) = 0 for solutions x = x(t) of (4.12).
th
For the j system the orbits are circles, i.e.,
qj + ipj = 2Ij eiφj with φj (t) = φj (0) + ωj t mod 2π.
For (4.12) the phase space decomposes into d-dimensional tori
{u ∈ Rd : I1 = c1 , . . . , Id = cd }.
For one or more vanishing Ij s we have dimensions of the tori between 1 and
d. The d-dimensional tori contain so called quasi-periodic solutions
x(t) = g(ω1 t, . . . , ωd t)
122 4. Hamiltonian dynamics
Example 4.3.3. Let F = F (I, φ). Then the map induced by φ = ∂ F and
I
I = ∂φ F is symplectic, cf. [Arn78, §48].
=H0 (I) I,
+ ε[∂ H0 · ∂ f + H1 (φ, 0)] + O(ε2 ).
I φ
The idea is to eliminate the terms of order O(ε) by finding f such that
I,
∂IH0 · ∂φf + H1 (φ, 0) = 0.
If we find such an f , then we can go on and find in the next step another
symplectic transformation which then eliminates the O(ε2 ) terms, etc., until
finally all perturbations are eliminated. Before we do so we look at the prob-
lem to eliminate the terms of order O(ε) in more detail. Given H1 (φ, I,
0)
we seek f : Td × Rd → R, such that
and H0 (I)
(4.17) ∂IH0 (I) I)
· ∂ f (φ, I,
= −H1 (φ, 0).
φ
of (4.17) is not changed. If there are no resonances until the nth step, then
the perturbation up to terms of order O(εn ) can be removed. We have the
following approximation theorem.
Theorem 4.4.1. If the normal form transformations allow to remove all
terms up to order O(εn ), i.e., if ∂φH = O(εn+1 ), then there exist C1 and
ε0 > 0 such that for all ε ∈ (0, ε0 ) we have in the original coordinates
sup I(t) − I(0) ≤ C1 ε.
t∈(−1/εn ,1/εn )
˙
I) ≤ C2 ε for a C2 > 0, and, since I ≤
Proof. We have I − I(φ, = ∂φH
C3 ε n+1 for a C3 > 0, we have
− I(0)
I(t) ≤ C3 |t|εn+1 ≤ C3 ε
for all |t| ≤ 1/εn . This yields
I(t) − I(0) ≤I(t) − I(t) − I(0)
+ I(t)
+ I(0) − I(0)
≤(2C2 + C3 )ε =: C1 ε.
So far we did not consider the convergence of the above Fourier series
in the normal form transforms. This turns out to be complicated due to
so called small divisor problems. This means that for given ω ∈ Rd and
(arbitrary small) δ > 0 there always is a k ∈ Zd such that
|k · ω| < δ.
Hence, the divisors in the series for f become arbitrarily small and the
convergence of the Fourier series is a serious problem. The problem is solved
by restricting the set of possible frequencies.
Definition 4.4.2. A vector ω ∈ Rd is called of type (L, γ) if for all k ∈
Zd \ {(0, . . . , 0)} we have
|k · ω| ≥ L|k|−γ
We remark that for given γ > d and almost all ω ∈ Rd there exists
a L > 0 such that ω is of type (L, γ), cf. [Arn88, page 114]. To study
the analytic properties of the generating function F we use the following
functions spaces.
Definition 4.4.3. For n ∈ N define the spaces
1,n = {a : Zd → C : a1,n = |a0 | + |ak ||k|n < ∞}.
k∈Zd
which is equipped with the norm f Cbn = n|j|=0 ∂φj f C 0 , where f C 0 =
b b
supφ∈Td |f (φ)|, and where F −1 : a → f is defined by f (φ) = k∈Zd ak eik·φ ,
cf. §5.1.
Lemma 4.4.5. Let ω ∈ Rd be of type (L, γ) and b ∈ 1,n with b0 = 0. Then
ibk
a, defined by ak = k·ω for k = 0, a0 = 0, is in 1,n−γ .
Proof. We have
bk
a1,n−γ = |ak |k| |= ≤ L−1 b .
k · ω |k|
n−γ n−γ
1,n
k∈Zd \{0} k∈Zd \{0}
is invertible. For all δ > 0 there is an ε0 > 0, such that for all ε ∈ (0, ε0 )
there is a set Pε ⊂ V × T d with the following properties. The Lebesgue
measure μ of (V × T d ) \ Pε is less than δ, and for all (I0 , φ0 ) ∈ Pε the orbit
through (I0 , φ0 ) is quasi-periodic.
At certain values there are gaps in the distribution, and these correspond to
low resonances between the periods of the asteroids and Jupiter.
Wu
p q
Ws
f k (S)
Exercises
4.1. Prove that ẏ = (traceM )y for y = detY , where Y (t) ∈ R2×2 satisfies Ẏ = M Y
for M = M (t) ∈ R2×2 .
4.2. The “6 − 12 Lennart-Jones potential” models the forces between two neutral
particles (atoms or molecules), namely an attractive van der Waals force at long
ranges and a repulsive force at short ranges due to overlapping electron orbitals.
In a simple (dimensionless) form it is given by F (u) = au−12 − bu−6 where u is
the distance between the particles and a, b > 0 are suitable constants. Choose
a = 0.001 and b = 1 and discuss the phase portrait of the system ü = −F (u).
4.3. Consider the pair ẍ + ω 2 x = 0, ÿ + μ2 y = 0 of (uncoupled) harmonic oscilla-
tors. Write this as a Hamiltonian system. Find two integrals in polar coordinates.
Discuss the cases (i) ω/μ rational and (ii) ω/μ irrational.
4.4. Given r, μ > 0, write down explicitly a circular solution of the 1-body problem
q̈ = −μq/q3 , q ∈ R2 , i.e., find initial conditions q0 , q̇0 such that the solution
satisfies q(t) = r for all t ∈ R.
4.5. In dimensionless form, the first “Post-Newtonian” approximation for the orbit
of a planet around the sun is
∂θ2 u + u = α + εu2 ,
where u = 1/r and (r, θ) are the polar coordinates of the planet and α, ε > 0 are
parameters. Discuss the phase portrait of this system.
4.6. Let M ∈ Rn×n be nonsingular and symmetric and F : Rn → R be smooth.
Write the Newtonian equation M ẍ + ∇F (x) = 0 as a Hamiltonian system.
4.7. Write the 4th order ODE u + qu + f (u) = 0 as a Hamiltonian system
for (u, u , u , u ). Hint. Let z=(u, u ) and derive a system T z +∇V (z)=0 with
non-singular T ∈R2×2 .
4.5. Homoclinic chaos 131
d q 1
4.8. Consider the 1-body problem = J∇H(q, p), H(q, p) = U (q) + |p|2 .
dt p 2
Show that the angular momentum C = q × q̇ is constant.
4.9. (The 2-body problem) Consider two mass points with positions qj ∈ R3
and masses mj that move under mutual gravitational attraction. The equations
are
m1 q̈1 = F21 , m2 q̈2 = F12 ,
with
Fij = mi mj g(|qi − qj |)(qi − qj ), g(r) = G/r 3 .
This problem can be completely reduced to the 1 body problem. For this consider
the center of mass qs = (m1 q1 + m2 q2 )/ms , with ms = m1 + m2 . Find the ODE
for qs and express the orbits q1,2 via qs and orbits of the 1 body problem for the
distance q = q2 − q1 .
4.10. Let F, G, H : R2n → R be smooth. Show that (a)
{F, {G, H}} + {G, {H, F }} + {H, {F, G}} = 0.
(b) F is an integral of u̇ = J∇H iff {F, H} = 0. (c) d
dt F (u) = {F, H}.
φ). Show that the map induced by φ = ∂ F and I = ∂φ F is
4.11. Let F = F (I, I
symplectic.
Chapter 5
PDEs on an interval
The second part of this book is about nonlinear dynamics in countably many
dimensions. It contains this chapter about PDEs on an interval and Chapter
6 about the Navier-Stokes equations.
We start with ordinary differential equations in R∞ , where R∞ stands
for the spaces RN or RZ , i.e., for the spaces of real or complex (identifying C
with R2 ) sequences (aj )j∈N or (aj )j∈Z . In this book these countably many
ODEs arise from PDEs, for which the spatial variable lives in a bounded
domain. For function spaces on such domains very often a countable basis
exists. By an expansion of the PDE w.r.t. this basis, for instance by an
expansion into Fourier series in case of rectangular domains and suitable
boundary conditions, the PDE can be transformed into an ODE in R∞
Example. Consider the linear heat equation ∂t u = ∂x2 u for x∈[0, π]
with boundary condition u(0, t)=u(π, t)=0. Expanding
u(x, t)= k (t) sin(kx),
u
k∈N
133
134 5. PDEs on an interval
and more generally up := ( dj=1 |uj |p )1/p , 1 ≤ p ≤ ∞, and finally u∞ =
maxj=1,...,d |uj |. We have for instance
u∞ ≤ up ≤ d1/p u∞ .
In infinite dimensions there are infinitely many non-equivalent norms. The
norms which we use in this section are as follows.
Definition 5.1.1. For p ∈ [1, ∞) and θ ∈ R let
1/p
up,θ (RZ ) = |un |p max(1, |n|)pθ .
n∈Z
For p = ∞ and θ ∈ R let
u∞,θ (RZ ) = sup |un | max(1, |n|)θ .
n∈Z
We set
p,θ (RZ ) = {u : Z → R : up,θ (RZ ) < ∞}.
Similarly, we define ·p,θ (RN ) and p,θ (RN ). We use the abbreviations
·p,θ and p,θ for ·p,θ (R∞ ) and p,θ (R∞ ).
We briefly recall the basic notions of semigroup theory which is the abstract
version of the subsequent analysis.
136 5. PDEs on an interval
lim T (t) − I = 0
t↓0
Theorem 5.1.8. Let supk∈N |λk | = α < ∞. Then for every θ ∈ R and
p ∈ [1, ∞] the associated semigroup is operator-continuous in p,θ for all
t ∈ R.
Theorem 5.1.10. For every θ ∈ R, p ∈ [1, ∞), and u(0) ∈ p,θ , the curve
t → u(t) is continuous in p,θ for t ≥ 0 if and only if supj∈N Reλj = α < ∞.
Proof. Let ε > 0. Using the triangle inequality in p,θ we have that
u(t) − u(0)p,θ
N ∞
=( |(e λn t
− 1)un (0)| |n| )
p pθ 1/p
+( |(eλn t − 1)un (0)|p |n|pθ )1/p
n=1 n=N +1
= s1 + s2
138 5. PDEs on an interval
for a N ∈ N suitably chosen in the following. In order to prove that s1 +s2 <
ε for t > 0 sufficiently small we first estimate s2 by choosing N so big that
∞
s2 ≤ (eαt + 1)( |un (0)|p |n|pθ )1/p < ε/2.
n=N +1
For this N we then find a t0 > 0 such that for all t ∈ (0, t0 ) we can estimate
s1 ≤ ( max |eλn t − 1|)u(0)p,θ < ε/2.
n=1,...,N
This means that the curve is one time differentiable for t ∈ (1, 2],
two times differentiable for t ∈ (2, 3], etc.
5.1. From finitely to infinitely many dimensions 139
n∈N n∈N
which holds if |ti | < tr . Such generators are called sectorial and play
a major role in the analysis of dissipative systems. The associated
semigroup (eλn t )n∈N is called analytic. We come back to this in
§6.3.
Figure 5.1. The picture shows the regions where the spectrum of the
generators must be contained in to have a continuous (left panel), a
differentiable (middle panel), or an analytic (right panel) semigroup.
1, if y = x2 , x = 0
f (x) =
0, elsewhere
is a finite-dimensional example of a function for which every directional
derivative exists, but which is not differentiable. In infinite-dimensional
spaces less ’exotic’ examples are possible.
Example 5.1.14. Consider X = Y = L2 (0, 1) and F (u)(x) = sin(u(x)).
We show that F is Gâteaux differentiable, but not differentiable at u = 0.
We have
F (u + τ v) − F (u) sin(τ v(x))
lim = lim = cos(0)v(x) = v(x)
τ →0 τ τ →0 τ
due to the differentiability of sin : R → R. For the Fréchet differentiability
we can vary v not only along lines. Due to the above computed Gâteaux
derivative the only possible candidate for the derivative A(0) is the identity.
We set
Proof. We have to show that for every ε > 0 there exists a δ > 0 such that
for all (ξ 1 , P1 ), (ξ 2 , P2 ) with P1 ≤ δ and P2 ≤ δ we have S(ξ 1 , P1 ) −
S(ξ 2 , P2 )X ≤ ε.
We set P3 = P1 ∪ P2 and choose an arbitrary ξ 3 . Then by the triangle
inequality we have
S(ξ 1 , P1 )−S(ξ 2 , P2 )X ≤S(ξ 1 , P1 )−S(ξ 3 , P3 )X + S(ξ 3 , P3 −S(ξ 2 , P2 )X
≤ε/2 + ε/2 ≤ ε.
142 5. PDEs on an interval
where we used
N1
N3
S(ξ , P1 ) − S(ξ , P3 )X =
1 3
f (ξj1 )(
xj −x
j−1 ) − f (ξj3 )(xj − xj−1 )X
j=1 j=1
N1
α(j)−1
1
(5.1) ≤ f (ξj ) − f (ξ 3 |xk+1 − xk | ,
k+1 ) X
j=1 k=α(j−1)
%α(j)−1
xj−1 , x
where [ j ] = k=α(j−1) [xk , xk+1 ]. See Figure 5.2.
x0 x x2 x3 x x5 x6 x7
1 4
~
x0 ~
x ~x
1 2
Figure 5.2. The partition P1 is drawn below the line and P3 above the
line. In this example we have α(0) = 0, α(1) = 3, α(2) = 7, . . ..
Theorem 5.1.20. Assume (5.3) and (5.4). For all C1 > 0 there exists a
T0 > 0 such that for all w ∈ p,θ with wp,θ ≤ C1 we have a unique solution
u ∈ C([0, T0 ], p,θ ) of (5.2) with initial condition u(0) = w.
Proof. We fix a C2 > 0 and show that for T0 ∈ (0, 1) sufficiently small the
right-hand side of (5.5) is a contraction in the set
n∈N
obviously the eigenvalues are not contained in a sector and the semigroup is
not analytic (w.r.t. time t = tr + iti ) since
| = sup |e−n
2 +i(−1)n n3 )(t 2t n+1 n3 t
sup |e(−n r +iti ) r +(−1) i
|=∞
n∈N n∈N
for ti = 0. However, we have the decay estimate
2 ±in3 )t
(eλn t un )n∈N p,θ ≤ sup |e(−n nθ |up,0 ≤ Ct−θ/2 up,0 .
n∈N
≤ Cθ−r (1 − α)−1 T01−α eβT0 sup B(u(τ ) + v(τ ), u(τ ) − v(τ ))p,θ
t∈[0,T0 ]
Theorem 5.1.23. Assume (5.7) and (5.8). For all C1 > 0 there exists a
T0 > 0 such that for all w ∈ p,θ with wp,θ ≤ C1 we have a unique solution
u ∈ C([0, T0 ], p,θ ) of (5.6) with initial condition u(0) = w.
Remark 5.1.24. Both theorems obviously also hold if B is replaced by a
general locally Lipschitz-continuous map from p,θ into p,θ or p,r respec-
tively, i.e., for instance in the latter case that for all C1 there exists an L
such that max(up,θ , vp,θ ) ≤ C1 implies
N (u) − N (v)p,r ≤ Lu − vp,θ .
Every polynomial nonlinearity is locally Lipschitz-continuous in this sense.
Moreover, Theorem 5.1.20 and Theorem 5.1.23 are prototypes for other
local existence and uniqueness theorems for semi-linear evolutionary PDEs
below.
k ≤ C(
Proof. a) Since ρm ρm
k−l + ρ m m
l ) with C = 2 for ρ k = max(1, |k|) using
Lemma 5.1.26 it follows that
u ∗ v1,m = | k−l vl ρm
u k |≤ |
uk−l vl ρm
k |
k∈Zd l∈Zd k∈Zd l∈Zd
≤C uk vl |
(| ρm
k + |
uk vl |
ρm
l )
k∈Zd l∈Zd
u1,0
≤C( v 1,m +
u1,m
v 1,0 ) ≤ 2C
u1,m
v 1,m .
148 5. PDEs on an interval
k∈Zd l∈Zd
≤( ( |
uk−l vl |(
ρm m
k−l + ρ
2 1/2
l )) )
k∈Zd l∈Zd
≤C(|
u|
ρ ∗ |
v |2,0 + |
m
u| ∗ |
v |
ρm 2,0 )
u2,m
≤C( v 1,0 +
u1,0
v 2,m ).
v 1,0 ≤
The final assertion follows from Sobolev’s embedding theorem
v 2,m for m > d/2.
C
We now give a number of classical examples of nonlinear PDEs over in-
tervals with periodic boundary conditions. In fact, over unbounded domains
each of these equations will play an important role in this book. For the
modeling and physical background of the equations we refer in particular to
Part III.
Except for the KdV equation, for all equations from above the local
existence and uniqueness theory can be handled with Theorem 5.1.20 and
Theorem 5.1.23. The linear parts are given by the eigenvalues λk , k ∈ Z,
where
a), b) λk = −k 2 + 1, c) λk = −k 2 ,
d) λk = −ik 3 , e)λk = −ik 2 , f) λk = −(1 + iμ)k 2 + 1.
In a), b), e) and f) we only need that (eλk t )k∈Z : 2,θ → 2,θ is bounded for
fixed t. Since the nonlinear terms in a), b), e) and f) are bi- and trilinear
maps from 2,θ → 2,θ for θ > 1/2 we have the local existence and uniqueness
for these equations in 2,θ for θ > 1/2 according to Theorem 5.1.20. Since
for c) and d) the nonlinear terms are only bilinear maps from p,θ+1 into p,θ
we need an estimate
(5.11) (eλk t )k∈Z p,θ →p,s+1 ≤ C max(1, t−α )
with α ∈ [0, 1) for the semigroup in order to apply our local existence and
uniqueness result from Theorem 5.1.23. According to Example 5.1.21 such
an estimate is true for c) with α = 1/2, but not for d). The KdV equation
is a so called a quasilinear (hyperbolic) equation. There is local existence in
2,θ for θ = 3, for instance. However, the proof is more involved, cf. [Paz83,
§8, Theorem 5.6] or §8.2 for further remarks.
and with zj ∈ Rn0 for j = 1, . . . , N . We claim that the unit ball of p,θ
is contained in the union of the balls Bε/2 ((zj , 0)) of p,r (RN ) if we choose
(n0 + 1)r−θ ≤ ε/2. This follows since for u = ((uk )k=1,...,n0 , u∞ ) in the unit
ball of p,θ (RN ) we have a j ∈ {1, . . . , N } such that
where ((uk )k=1,...,n0 , 0) − (zj , 0)p,r ≤ ε/2 due to the construction of the
points zj and where (0, u∞ )p,r ≤ ε/2 due to
∞
∞
(0, u∞ )p,r ≤( |un | |n| )
p pr 1/p
≤ sup |n| r−θ
( |un |p |n|pθ )1/p
n0 +1 n=n0 +1,...,∞ n0 +1
≤(n0 + 1) r−θ
(0, u∞ )p,θ ≤ ε/2
From the definition it is clear that for u ∈ C m (Ω, R) we have uCbm < ∞,
if Ω is bounded. More generally, we define
Cbm (Ω, R) = {u ∈ C m (Ω, R) : uCbm < ∞}.
For Ω = Ω = R the function u(x) = x is in C 0 , but not in Cb0 . For Ω
bounded, Cbm is dense in Cb0 . For the treatment of unbounded Ω we define
m
Cb,unif (Ω, R) = {u : Ω → R :∂xj u is uniformly continuous for
|j| = 0, . . . , m, uCbm < ∞}.
For Ω = R the function u(x) = sin(x2 ) is in Cb0 , but not in Cb,unif
0 . For
similar reasons Cb (R, R) is not dense in Cb (R, R), but Cb,unif (R, R) in
n 0 n
0
Cb,unif (R, R). All these spaces are Banach spaces.
Hölder spaces. The spaces Cb0 and Cbm are not the optimal choice for
solving linear PDEs. Even for arbitrarily smooth boundary ∂Ω the boundary
value problem
(5.12) Δu = f in Ω, u|∂Ω = 0,
for f ∈ Cb0 in general does not possess a solution u with optimal regularity,
cf. Example 5.2.4 on page 157. Optimal regularity holds for the subsequently
defined Hölder-continuous functions and Sobolev functions, i.e., for instance
for (5.12) from f ∈ C 0,α it follows u ∈ C 2,α . For α ∈ (0, 1] we define
C 0,α (Ω, R) = {u : Ω → R : u is α -Hölder-continuous, uC 0,α < ∞}
equipped with the norm
|u(x) − u(y)|
uC 0,α = uC 0 + sup ,
b
x,y∈Ω,x=y,|x−y|≤1 |x − y|α
5.2. Basic function spaces and Fourier series 153
All these function spaces are Banach spaces, cf. Exercise 5.9. C 0,1 (Ω, R) is
the space of Lipschitz-continuous functions.
Lebesgue and Sobolev spaces. Unfortunately, the above spaces are
not equipped with a scalar product and so tools from linear algebra related
to orthogonality are not available. A natural choice of a scalar product for
functions would be
(5.13)
u, vL2 = u(x)v(x) dx.
Ω
However, if the above spaces are equipped with the above scalar product
they are not complete w.r.t. the induced norm. For instance the sequence
(un )n∈N with
⎧
⎨ 1, for |x| ≤ 1 − 1/n,
un (x) = 0, for |x| ≥ 1,
⎩
n(1 − |x|), for |x| ∈ (1 − 1/n, 1),
is a Cauchy sequence w.r.t. the norm induced by the L2 -scalar product.
However, the limit function is not in Cb0 although un ∈ Cb0 for all n ∈ N.
Since the limit of a Cauchy sequence of Riemann integrable functions
is in general no longer Riemann integrable the Riemann integral has to
be replaced by the Lebesgue integral in order to define complete function
spaces [Alt16, §A1]. In order to define the Lebesgue and Sobolev spaces we
introduce
C ∞ (Ω, R) = {u : Ω → R : u is arbitrarily many times differentiable}
and
C0∞ (Ω, R) = {u ∈ C ∞ (Ω, R) : u has compact support in Ω}
where the support of a function is defined by supp(u) = clRd {x ∈ Ω : u(x) =
0}. The Lebesgue spaces are defined by
1/p
∞
L (Ω, R) = cl·Lp (C0 (Ω, R)), where uLp =
p
|u(x)| dx
p
Ω
for all p ∈ [1, ∞). By construction all these spaces are Banach spaces consist-
ing of equivalence classes of Cauchy sequences, with two Cauchy sequences
154 5. PDEs on an interval
in the same class if their difference converges to zero. The Lp -spaces con-
structed in this way coincide with the spaces known from measure theory.
The space L2 (Ω, R), respectively L2 (Ω, C), is a Hilbert space equipped with
the scalar product (5.13).
For the solution of PDEs so called Sobolev spaces turn out to be useful.
For p ∈ [1, ∞) and Ω bounded we define
W m,p (Ω, R) = cl·W m,p (C ∞ (Ω, R)),
where
uW m,p = ( ∂xj upLp )1/p
|j|≤m
and
W0m,p (Ω, R) = cl·W m,p (C0∞ (Ω, R)),
for general Ω. Since the sum in the definition of · W m,p is finite there
are various equivalent norms such as uW m,p = |j|≤m ∂xj uLp . By con-
struction these spaces are Banach spaces, too. The spaces H m (Ω, R) =
W m,2 (Ω, R) and H0m (Ω, R) = W0m,2 (Ω, R) are Hilbert spaces equipped with
the scalar product
& '
u, vH m = ∂xj u, ∂xj v L2 .
|j|≤m
We introduce
m
uW m,∞ = ∂xα uL∞
|α|=0
and the space W m,∞ (Ω, R) as the space of all functions u : Ω → R for which
the weak derivatives ∂xα u exist for |α| = 0, . . . , m and for which uW m,∞ <
∞.
Remark 5.2.1. The concept of weak derivatives can be generalized to the
concept of distributional derivatives [RR04, Chapter 5]. A priori, the sets
C ∞ and C0∞ are just vector spaces. There is no norm for which these spaces
are complete. However, using inductive limits of semi-norms, C0∞ (Ω, R) can
be made to be a complete metric space D(Ω), called space of test functions,
where convergence un → u in D(Ω) means: a) There exists a compact
K ⊂ Ω such that supp(un ), supp(u) ⊂ K, b) limn→∞ ∂xα un (x) = ∂xα u(x)
uniformly in K for all α ∈ Nd . However, this convergence is not induced by
a norm.
The elements of the dual space of D(Ω, R) = C0∞ (Ω, R) are called dis-
tributions, i.e., a distribution T is a continuous linear map from D into the
real or complex numbers. This means that un → u in D implies T un → T u,
which is equivalent to the formulation that for all open bounded sets D there
is a constant C and a number m ∈ N such that
(5.14) |T (φ)| ≤ CφCbm for all φ ∈ C0∞ (D, R).
For a continuous function u ∈ Cb0 (Rd , R), or for u in one of the above other
spaces
Tu (φ) = u(x)φ(x) dx
Rd
defines the so called associated distribution, which is then called regular.
For the distribution associated to ∂xα u we find
T∂xα u (φ) = (∂xα u(x))φ(x) dx
Rd
=(−1)|α| u(x)(∂xα φ(x)) dx = (−1)|α| Tu (∂xα φ).
Rd
156 5. PDEs on an interval
Example 5.2.2. For u(x) = |x| we show that u ∈ H 1 ((−1, 1)) but u ∈
H 2 ((−1, 1)) by computing the weak derivatives ∂x u, ∂x2 u. For φ ∈ C0∞ ((−1, 1))
0 1 0 1
we have Tu (∂x φ) = −1 −x∂x φ dx + 0 x∂x φ dx = −1 φ dx + 0 −φ dx =
−Tg (φ) with
−1, x < 0,
g(x) = ∂x u(x) =
1, x > 0.
0 1
Similarly, Tu (∂x2 φ) = −1 ∂x φ(x) dx+ 0 −∂x φ(x) dx = 2φ(0) = 2δ0 (φ) where
δ0 is called the Dirac δ distribution. Thus, ∂x2 u ∈ L2 as there is no function
1
g such that −1 g(x)φ(x) dx = φ(0).
Lemma 5.2.3. Let −∞ < a < b < ∞. Then H 1 ((a, b)) ⊂ C 0,1/2 ((a, b))
and
1
(5.15) uL∞ ≤ 2uL2
2
uL2 + ∂x uL2 ,
b−a
√
(5.16) |u(x) − u(y)| ≤ x − y∂x u2L2 .
5.2. Basic function spaces and Fourier series 157
Proof. Since C 1 ((a, b)) is dense in H 1 ((a, b)) w.r.t. the · H 1 -norm, it is
sufficient to prove (5.15) and (5.16) for u ∈ C 1 ((a, b)). We have
x
2 d s−a 2
u (x) = u (s) ds
ds x − a
a x x
1 2 s−a
= u (s) ds + 2u(s)∂x u(s) ds
a x−a a x−a
1
≤ u2L2 + 2uL2 ∂x uL2
x−a
and similarly
b
d s−x 2 1
2
u (x) = u (s) ds ≤ u2L2 + 2uL2 ∂x uL2 .
x ds b − x b − x
Hence, u2 (x) ≤ min{ x−a
1 1
, b−x }u2L2 + 2uL2 ∂x uL2 . For the second esti-
mate we use the Cauchy-Schwarz inequality, namely
y y
|u(x) − u(y)| = ∂s u(s) ds ≤ 1|∂s u(s)| ds ≤ |x − y|∂x uL2 .
x x
0.5
0.5
0
-0.5
-0.5 0 0.5 -1
1
5.2.2. Fourier series. PDEs with periodic boundary conditions for the
spatial coordinates can be transferred to ODEs in R∞ with the help of
Fourier series. For notational simplicity we restrict ourselves first to the
torus Td = Rd /(2πZ)d . Let Cper ∞ be the space of functions u : Td → Rd , with
∞
a C periodic extension u : Rd → Rd satisfying
u (x1 + 2π, x2 , . . . , xd ) = u
(x1 , x2 , . . . , xd ) = u (x1 , x2 + 2π, . . . , xd )
(x1 , x2 , . . . , xd + 2π).
= ... = u
We define
m ∞
Hper = clos·H m (Td ) (Cper ).
The question is if and in what sense a function can be represented by
its Fourier series, or equivalently, in which norms Fourier series converge.
In L2 we have a simple answer which follows from the general theory of
orthonormal systems, and which for convenience we summarize here.
Definition 5.2.5. Let H be a Hilbert space with scalar product
·, · : H ×
H → C. A (finite or infinite) system (φj ) in H, j = 1, . . . , N or j ∈ N, is
called orthogonal system if
φi , φj = 0 for i = j. It is called orthonormal
if additionally
φj , φj = 1. It is called a complete orthonormal system
(complete ONS) or Hilbert basis if
u, φj = 0 for all j implies u = 0 for
u ∈ H.
and since (φj ) is complete this implies v = u. (ii)⇒(iii) again follows from
Lemma 5.2.6 a), and (iv) follows from (iii) with v = u. Finally, (iv)⇒(i)
since
u, φj = 0 for all j and (iv) imply u = 0, hence u = 0.
Due to the equivalence of (5.18) and (5.19), often both are called Par-
seval’s identity. Clearly, Lemma 5.2.6 and Lemma 5.2.7 also holds if se-
quences (φj )j∈N are replaced by sequences (φj )j∈Z , (φj )j∈Nd , (φj )j∈Zd , with
the respective replacements in the sums. The most important example are
classical Fourier series.
1
SN (x) = k e
u ik·x
, with k =
e
u ik·x
, uL2 = u(x)e−ik·x dx
(2π)d Td
|k|≤N
1
|
uk |2 = u2L2 .
(2π)d
k∈Zd
1, if k = m,
Proof. a) By direct calculation we find
φk , φm = i.e.,
0, else,
the (φk ) are an ONS. The completeness of this ONS can be shown with the
Weierstraß approximation theorem, see [Alt16, Satz 7.10].
b) follows since SN is the orthogonal projection of u on TN .
c) Parseval’s identity can be computed directly for finite sums. Going
to the limits shows the assertion.
5.2. Basic function spaces and Fourier series 161
The map u → ( u)k∈Zd will be abbreviated with F . By c), F is an
isometric isomorphism from L2 to 2 . Its inverse ( uk )k∈Z → u is denoted
by F −1 . By d) the smoothness of u is related to the decay of its Fourier
coefficients.
Formally we have ∂x u(x) = uk eikx , or equivalently F (∂x u) =
k∈Z ik
uk )k∈Z . It follows, that F is in fact an isomorphism between the Sobolev
(ik
m and the spaces of sequences
spaces Hper 2,m which have been introduced in
−1
§5.1. Moreover, F maps 1,m to Cb . m
Lemma 5.2.9. Let m ∈ N0 . a) There exists a C > 0, such that for all
∈ 1,m
u
uCbm ≤ C
u1,m .
∈ 2,m
b) There exist C1 , C2 > 0, such that for all u
C1
u2,m ≤ uH m ≤ C2
u2,m .
C1 uH m ≤
u2,m ≤ C2 uH m .
tively the space of finite sequences. The results then follow by continuous
extension, see the subsequent Lemma 5.2.10.
a) We have
j
uCbm ≤ C sup sup ∂x k eikx ≤ C sup sup
u |k|j |
uk | |eikx |
x∈R 0≤j≤m k∈Z
x∈R 0≤j≤m k∈Z
≤ C
u1,m .
Parseval 1
= F (∂xm u)22,0 = ∂ m u2 2 ≤ u2H m .
2π x L
Proof. The condition that the extension must be continuous leads to the
only possible extension of f , namely
f(x) = lim f (x ).
x ∈A,x →x
It remains to prove the existence of this limit, i.e., to prove that f is well
defined. In order to do so, let (xn )n∈N be a sequence with xn ∈ A and
limn→∞ xn = x. Hence, (xn )n∈N is a Cauchy sequence in X and from the
uniform continuity it follows that the image sequence (f (xn ))n∈N is a Cauchy
sequence in Y . Since Y is complete, we have the existence of
y = lim f (xn )
n→∞
Remark 5.2.13. (Real Fourier series) Besides the complex Fourier ex-
pansion also real Fourier polynomials and series of the form
∞
a0
u(x) = + [ak cos(k · x) + bk sin(k · x)],
2 d
k∈N
where
1
k =
u u(x)e−iωk ·x dx = (F u)k .
L1 L2 · · · Ld Ω
Again, F is an isomorphism between Hper
m (Ω) and
2,m .
164 5. PDEs on an interval
Td u dx = 0 we have
(5.20) |u|2 dx ≤ |∇u|2 dx.
Td Td
5.2. Basic function spaces and Fourier series 165
Lemma 5.2.18. For m − d/2 > n there exists a C > 0, such that
uC n (Td ) ≤ CuH m (Td ) .
Proof. The assertion follows from Lemma 5.2.9 and Lemma 5.1.27.
Analytic properties of the solution operator of a linear evolution equation
can be established with the help of Fourier series.
Example 5.2.19. We consider the solution operator T (t) defined via the
solution u(x, t) = T (t)u0 )(x) of the linear heat equation ∂t u = ∂x2 u, with x ∈
[0, π], under Dirichlet boundary condition u(0, t) = u(π, t) = 0 to the initial
value u(x, 0) = u0 (x). In order to prove that (T (t))t≥0 is a C0 -semigroup in
L2 ((0, π)) and in H m ((0, π))∩H01 ((0, π)) for every m ∈ N we make an odd 2π-
periodic extension of the functions with u(0, t) = u(π, t) = 0. The semigroup
in the space of 2π-periodic functions is denoted again by T (t). We proved
in §5.1.2 that T(t) = F T (t)F −1 defined by (T(t) u(0))k∈Z = (e−k t u
2
k (0))k∈Z
(0) ∈ 2,m we have
is continuous in 2,m , i.e., for every u
T(t) (0)2,m → 0
u(0) − u for t→0.
Due to the isomorphism property of F between Hper
m and
2,m , cf. Lemma
5.2.9 b) and c), it follows that
T (t)u(0) − u(0)Hper
m ≤ C1 F −1 T (t)u(0) − u
(0)2,m
= C1 T (t)
u(0) − u(0)2,m → 0 for t → 0 ,
m . The restriction of x to [0, π] gives
i.e., T (t) is a C0 -semigroup in Hper
the result. Moreover, from Example 5.1.21 it is known that for r ≥ 0 the
semigroup T(t) can be estimated by
2.
with C = C1 CC
166 5. PDEs on an interval
Example 5.2.20. We consider ∂t2 u = ∂x2 u with x ∈ (0, 2π) and periodic
boundary conditions. We rewrite this equation as first order system for
z = (∂t u, ∂x u) and obtain
d 0 ∂x
z(x, t) = Az(x, t), with A = ,
dt ∂x 0
or in Fourier space
d z (k, t), = 0 ik
z(k, t) = A with A .
dt ik 0
The general solution is given by
z(x, t) = c1 (k)eik(t+x) z1 + c2 (k)eik(−t+x) z2 =: etA z(·, 0)(x),
k∈Z
where
c1 (k) 1 1 1
= z(k, 0).
c2 (k) 2 1 −1
Hence, we have a uniformly bounded C0 -semigroup for z in, e.g., H m × H m ,
which however is not smoothing.
Theorem 5.2.22. For all C1 > 0 there exists a T0 > 0 such the following
holds. For u0 ∈ H θ with u0 H θ ≤ C1 there exists a unique solution u ∈
C([0, T0 ], H θ ) of (5.21) with u|t=0 = u0 .
5.3. The Chafee-Infante problem 167
Proof. The proof goes line for line as the proof of Theorem 2.3.4 a).
In order to check the assumptions for system (5.21) for a concrete non-
linear PDE we have to handle products of functions in physical space. The
θ -spaces are closed under multiplication if θ is sufficiently big.
Hper
Lemma 5.2.24. For all θ > d/2 there exists a C > 0, such that for all
u, v ∈ Hper
θ we have
u ∈ C([0, T0 ], Hper
θ ) with u|
t=0 = u0 , respectively, A ∈ C([0, T0 ], Hper ) with
θ
A|T =0 = A0 .
The θ in the last theorem can be made smaller by using the smoothing
properties of the semigroup, cf. §6.2.1. Moreover, the smoothing estimate
(5.11) can be used to show that solutions to the KPP equation, the Allen-
Cahn equation, and the GL equation become arbitrary smooth and even
analytic for t > 0. This is done for instance in §5.3.3 or §6.2.2.
Proof. Again the solutions are extended to odd 2π-spatially periodic solu-
tions. Then we have
2π
d 2π 2
(5.25) u dx = 2 −(∂x u)2 + αu2 − u4 dx
dt 0 0
2π
d 2π
(5.26) (∂x u)2 dx = 2 −(∂x2 u)2 + α(∂x u)2 − 3u2 (∂x u)2 dx.
dt 0 0
If α < 0 all terms on the right-hand side are negative and we have
lim sup u(t)H 1 = 0.
t→∞
In order to obtain estimates which are also good for small α ≥ 0 we split the
parameter regime α ≥ 0 in two parts. First let α ∈ [0, 1/2]. Adding (5.25)
5.3. The Chafee-Infante problem 169
Proof. For convenience we repeat the main steps from the proof of Theorem
2.4.4. The attractor is defined by
*
A= At
t≥0
with At = closH 1 (St (B)). Since B is positively invariant, the family (At )t≥0 ,
satisfies At1 ⊂ At2 for t1 > t2 . Hence, A ⊂ A0 is bounded. Since St is a
compact operator for t > 0, the set At is compact for t > 0. Since (At )t≥0
is a decreasing family of compact non-empty sets, the attractor A = ∩t≥0 At
is non-empty and compact.
We skip the proof of the time invariance and restrict ourselves to the
attractivity which is proved by contradiction. We assume that B is not
attracted by A. Then there exists a δ > 0, sequences tn → ∞ and un ∈ B,
such that dist(Stn (un ), A) > δ > 0 for all n ∈ N. For a small t > 0 the
sequence Stn −t (un ), (n ∈ N) is bounded. Since St is a compact operator
there exists a subsequence such that vj = Stnj (unj ) converges towards a w
for j → ∞. Therefore, w ∈ A which contradicts the above assumption that
the sequence is bounded away from A.
spaces. More or less all definitions in the theory of dynamical systems, such
as continuity of solutions w.r.t. time, stability of solutions, etc. depend on
the chosen norm. Therefore, we expect that the choice of a suitable phase
space in infinitely many dimensions in general plays a crucial role. It is the
purpose of this section to explain that for systems with smoothing proper-
ties this often is not the case. If there is a global bound in one H θ -space,
then it does not matter which H θ -space is chosen as long as these spaces
are connected with a smoothing estimate.
The estimate (5.27) can be generalized to
t
u(t)H θ+1 ≤T (t − τ )u(τ )H θ+1 + T (t − s)u3 (s)H θ+1 ds
τ
t
≤C(t − τ )−1/2 u(τ )H θ + C(t − s)−1/2 ds sup u(s)3H θ
τ s∈[τ,t]
−1/2
≤C(t − τ ) Cθ (τ ) + 2C(t − τ ) 1/2 3
Cθ (τ ) ,
where Cθ (τ ) = sups∈[τ,∞] u(s)H θ . Hence
with potential
π
1 α 1
V (u) = (∂x u(x))2 − u(x)2 + u(x)4 dx
0 2 2 4
and β a linear map defined below. In order to justify this formula, first
recall that for a function V : Rd → R we have
V (u + εv) = V (u) + εaT v + O(ε2 ) = V (u) + ε
a, v + O(ε2 ) for all v ∈ Rd ,
where
u, v = uT v is the scalar product between the vectors u and v, i.e.,
the derivative is defined as an element of the dual space of Rd . However, it
can be identified with Rd through the map
β : Lin(Rd , R) → Rd ,
a, · → a.
For a map V : X → R where the function space X is equipped with the
scalar product π
u, v = u(x)v(x) dx
0
we define the map
β : Lin(X, R) → X,
a, · → a.
This is well defined since in Hilbert spaces the dual space Lin(X, R) can be
identified with X by the Riesz representation theorem [Alt16, Satz 4.1].
Using the boundary conditions and integration by parts we find
V (u + εv) − V (u)
lim
ε π
ε→0
1 α 1
= lim ε−1 ( (∂x (u + εv))2 − (u + εv)2 + (u + εv)4 )
ε→0 0 2 2 4
1 α 1
− ( (∂x u)2 − u2 + u4 ) dx
2 2 4
π
=− (∂x2 u + αu − u3 )v dx
0
and so by comparison
β∂u V (u) = −(∂x2 u + αu − u3 ).
Therefore, (5.24) is a gradient system in H01 , and hence the function t →
d
V (u(t)) decreases along solutions u = u(t), i.e., dt V (u(t)) ≤ 0, where equal-
ity only holds in fixed points. Consequently, no non-trivial periodic solution
can occur. Moreover, V is bounded from below, since
π π 2
1 α 1 α πα2
V (u) = (∂x u(x)) − u(x) + u(x) dx ≥ −
2 2 4
dx = − .
0 2 2 4 0 4 4
Similar to the finite-dimensional situation, cf. Theorem 2.4.15, the attractor
consists of the fixed points and their unstable manifolds, cf. [Rob01, The-
orem 10.13]. In case that only finitely many fixed points exist, the attractor
5.3. The Chafee-Infante problem 173
consists of these fixed points and their heteroclinic connections. This can be
seen directly. In a gradient system every solution must end in a fixed point.
Solutions in the attractor must also start in one of the finitely many fixed
points. This follows from the fact that backwards in time the system in the
attractor is a gradient system, too. The potential is given by −V and it is
bounded on the attractor.
We compute the fixed points, or stationary solutions, of the PDE, which
satisfy
∂x2 u + αu − u3 = 0.
Due to the boundary conditions u(0) = u(π) = 0 in the (u, u )-plane we
have to find solutions which start from the v = u -axis, end on this axis,
and need for this part of the orbit the ’time’ x = π. For all α > 0 the phase
portrait looks qualitatively the same. The periodic orbits around the origin
have a periodicity which is i) minimal at the origin, namely the periodicity
√
of the linearization, 2π/ α, ii) infinity at the heteroclinic orbits, and iii)
which increases strictly monotonic with the distance from the origin. Thus,
non-trivial equilibria of (5.24) can only exist for α > 1 since for α ≤ 1 the
solutions are too slow to make half of the periodic orbit in a time π.
Using i)-iii) the complete bifurcation picture can be established in a
rigorous way. The number of solutions with u(0) = u(π) = 0 changes for
√
mπ/ α = π with m ∈ N, an integer multiple of half the minimal period.
As a consequence, for α ∈ (−∞, 1] we have one equilibrium, the origin; for
α ∈ (1, 4] we have 3 equilibria, the origin, and two equilibria called u±1 ;
for α ∈ (4, 9] we have 5 equilibria, . . . ; and for α ∈ (m2 , (m + 1)2 ] we have
2m + 1 equilibria, the origin, u±1 , . . ., and u±m .
Hence, for fixed α there are only finitely many fixed points which are
elements of the attractor. In order to understand the dynamics in the at-
tractor, i.e., to find the heteroclinic connections between the fixed points,
we analyze the linearization at the fixed point u ≡ 0, i.e.,
∂t u = ∂x2 u + αu,
with u(0, t) = u(π, t) = 0, or equivalently, with u(x, t) = n (t) sin(nx),
n∈N u
d
n = (α − n2 )
u un .
dt
Therefore, the linear operator
Λ· = ∂x2 · +α·
with Dirichlet boundary conditions has eigenvectors u(x) = sin mx with
associated eigenvalues λ = α − m2 for m ∈ N. Equivalently the infinite-
dimensional diagonal matrix
nm )n,m∈N = ((α − m2 )δnm )n,m∈N
(Λ
174 5. PDEs on an interval
-15 -8 -3 Re
1 4 9 α
Figure 5.5. Left: the spectrum of the operator Λ· = ∂x2 ·+α· under
Dirichlet boundary conditions for α = 1. Right: the bifurcation
diagram. At the parameter values α = n2 unstable equilibria bifur-
cate via a pitchfork bifurcation from the trivial branch u(α) ≡ 0.
Hence, for α ∈ (−∞, 1) the origin is asymptotically stable. For α > 1 the
origin is unstable, with a one-dimensional unstable manifold for α ∈ (1, 4],
with a two-dimensional unstable manifold for α ∈ (4, 9], and with an m-
dimensional unstable manifold for α ∈ (m2 , (m + 1)2 ].
For α ∈ (1, 4) the one-dimensional unstable manifold of the origin ends
in the stable equilibria u±1 . For α ∈ (4, 9) the equilibria u±1 , u±2 lie on
the two-dimensional unstable manifold of the origin. Since u±2 bifurcates
from the unstable origin, these fixed points are also unstable and their one-
dimensional unstable manifold ends in u±1 .
The reasons are as follows. Since for fixed α the fixed points uj are
isolated and since the linearization only has real eigenvalues due to the gra-
dient structure, no eigenvalue of the linearization around the equilibria uj
crosses the imaginary axis after the bifurcation when α is increased. There-
fore, the dimension of the unstable manifold of uj is the same as at their
bifurcation point from the trivial branch. The fixed points u±1 bifurcating
at α = 1 are always stable. The fixed points u±2 bifurcating at α = 4 have
a one-dimensional unstable manifold which ends in the fixed points u±1 .
The fixed points u±3 bifurcating at α = 9 have a two-dimensional unsta-
ble manifold and so heteroclinic connections to the fixed points to u±1 and
u±2 exist. Figure 5.6 sketches the dynamics in the attractor of (5.24). For
α ∈ (n2 , (n + 1)2 ) we have an attractor of dimension n consisting of finitely
many fixed points and heteroclinic orbits between these fixed points, in par-
ticular, it contains the n-dimensional unstable manifold of the origin. A local
bifurcation analysis via center manifold reduction can be found in §13.2.1.
Further Reading. Our point of view of PDEs over bounded sets as count-
ably many ODEs is similar to [Hal88, Rob01, KP13], while [Paz83] gives
5.3. The Chafee-Infante problem 175
Exercises
5.1. Prove that the space c00 = {u : Z → R : un = 0 for finitely many n} equipped
with the 1 -norm is not complete.
d2
5.2. Consider un = −ωn2 un , n ∈ N, with un (t) ∈ R, and ωn ∈ R. Write the
dt2
equation as first order system and find some phase space where the infinitely many
first order ODEs define a C0 -semigroup. Under which additional assumptions on
the ωn is the semigroup uniformly continuous, differentiable, or analytic?
5.12. Let d ∈ N and Ω = B1 (0) in Rd . For which α do we have (a) |x|α ∈ H 1 (Ω);
(b) (sin |x|)α ∈ H 1 (Ω); (c) (ln |x|)α ∈ H 1 (Ω) ?
1/2
5.13. For H = L2 (0, 1) define F : H → R by F (u) = 0 u(x) dx. Do we have
F ∈ H ? If so, find a representation
·, vL2 of F with v ∈ H.
5.3. The Chafee-Infante problem 177
5.14. For the following PDEs with x ∈ (0, 2π) and periodic boundary conditions
investigate whether the solution operator defines a C0 -semigroup in L2per ((0, 2π), R)
with smoothing properties
a) ∂t u = ∂x4 u, b) ∂t u = −∂x4 u, c) ∂t u = ∂x3 u.
5.15. Consider the complex GL equation
∂t u = (1 + iα)∂x2 u + Ru − (1 + iβ) |u|2 u
with 2π-periodic boundary conditions, u(x, t) ∈ C, and α, β ∈ R. In case |β| < 1/3
prove the global existence of solutions in H 1 for all R ∈ R.
5.16. Consider ∂t u = ∂x2 u + u3 for t ≥ 0, x ∈ (0, π) and u(x, t) ∈ R with boundary
condition u(0, t) = u(π, t) = 0 and initial condition u(x, 0) = φ(x), cf. [Hen81,
Page 49]. Prove that there are solutions which converge
π in finite time towards ∞.
Hint: Derive a differential inequality for s(t) = 0 sin(x)u(x, t)dx. With Hölder’s
π
inequality we obtain s(t) ≤ 22/3 ( 0 sin(x)u3 (x, t)dx)1/3 and so dtd
s ≥ −s + s3 /4.
5.17. Write ∂t u = −∂x4 u + sin(u), with u(x, t) ∈ R, x ∈ R, and t ≥ 0, subject to
periodic boundary conditions u(x, t) = u(x + 2π, t) as a gradient system.
5.18. Consider the Cahn-Hilliard equation ∂t u = ∂x2 (−γ∂x2 u−u+u3 ), with u(x, t) ∈
R, γ > 0, and 2π-periodic boundary conditions.
d
2π
a) Prove that dt C = 0, where C(t) = 0 u(x, t) dx.
2π 1 2
b) Let F (u) = 0 4 (u − 1)2 + γ2 |∂x u|2 dx and show that dt d
F = − |∇w|2 dx
with w = u3 − u − γ∂x2 u.
c) Find the possible ω-limit sets.
Chapter 6
The Navier-Stokes
equations
6.1. Introduction
In this chapter we give an introduction to the Euler and Navier-Stokes equa-
tions, which over unbounded domains will also play a role in subsequent
chapters. The global existence and uniqueness of solutions of the three-
dimensional (3D) Navier-Stokes equations is one of the seven so called ’one
million dollar’ or millennium problems in mathematics presented by the Clay
Mathematics Institute in the year 2000. There are a number of reasons for
this choice. On the one hand, the solution of this problem would allow us to
understand and simulate the motion of fluids more rigorously. On the other
hand, the 3D Navier-Stokes equations are interesting PDEs which resisted
so far all attempts to prove the global existence and uniqueness of solutions.
Their history goes back a long way. The equations describing the motion
of non-viscous fluids are called Euler equations and have been derived by
Leonhard Euler (1707–1783). The Navier-Stokes equations generalize the
Euler equations and include the case of viscous fluids. They have been
derived independently by a number of people, including Claude-Louis Navier
(1785–1839), George Stokes (1819–1903), Simeon Poisson (1781–1840) and
Jean Claude Saint-Venant (1797–1886).
First we recall the derivation of the Navier-Stokes equations, following
[Fow97, §6]. Then we focus on the analysis of the Navier-Stokes equa-
tions in Ω = Td = Rd /(2πZ)d , i.e., Ω = [0, 2π)d with periodic boundary
conditions. After the presentation of some local existence and uniqueness
results we formulate the global existence question. The local existence and
179
180 6. The Navier-Stokes equations
mẍj = Fj (x1 , . . . , xN )
where we used the Gauss integral theorem and where n(x) = (n1 , . . . , nd )(x)
is the outer unit normal in the point x at the boundary S. Since this relation
holds for all test volumes V the integrands must be equal, i.e.,
(6.1) ∂t ρ + div(ρu) = 0.
6.1. Introduction 181
For examples see below. Hence, we obtain for the change of the momentum
d d
d
ρui dV = − (ρui )uj nj dS + σij nj dS.
dt V S S
j=1 j=1
Application of the Gauss integral theorem and the above arguments yield
∂t (ρui ) + div(ρui u) = div(σi· ).
Using conservation of mass gives
∂t (ρui ) = ρ∂t ui + ui ∂t ρ = ρ∂t ui − ui ∂xj (ρuj ),
j=1,...,d
V
x2
x1
Figure 6.1. Shear flow: the internal friction is proportional to ∂x2 u1 .
d
∂xj uj = ∇ · u = divu = 0.
j=1
d d
From k=1 ε̇kk = k=1 ∂xk uk = 0 we obtain τij = 2με̇ij . Moreover, we find
d
d
d
d
d
∂xj τij = ∂xj (∂xj ui + ∂xi uj ) = ∂x2j ui + ∂xi ( ∂xj uj ) = ∂x2j ui ,
j=1 j=1 j=1 j=1 j=1
such that
1
∂t u + (u · ∇)u = − ∇p + νΔu,
ρ
where ν = μ/ρ is called the kinematic viscosity. The Navier-Stokes equations
are then given by
1
∂t u + (u · ∇)u = − ∇p + νΔu,
(6.6) ρ
∇ · u = 0.
see Figure 6.2c). A 3D flow generalizing the 2D strain flow is the irrotational
stationary jet
⎛ ⎞
−γ1 x1
1
u(x, t) = ⎝ −γ2 x2 ⎠ , p(x, t) = p0 − (γ12 x21 + γ22 x22 + (γ1 + γ2 )2 x23 ).
2
(γ1 + γ2 )x3
a) b) c)
x 2 ,u 2 x 2 ,u 2 x 2 ,u 2
x 1 ,u 1 x 1 ,u 1
x 1 ,u 1
Figure 6.2. Three planar exact solutions: Poiseuille flow, strain flow,
a vortex.
and that the variable p occurs without time derivative at all. This problem
is solved by prescribing the equation ∇ · u = 0 as additional condition in the
definition of the phase space. The term −∇p in the first equation will be
interpreted as projection P onto the divergence free vector fields such that
the Navier-Stokes equations can be written as
∂t u = P Δu − P (u · ∇)u.
For periodic boundary conditions we will have P Δ = ΔP such that we
finally have to consider ∂t u = Δu − P (u · ∇)u in the space of divergence free
vector fields {u : u = P u}.
The Navier-Stokes equations are semi-linear parabolic differential equa-
tions such that for the construction of local solutions in time we use again
the variation of constant formula
t
(6.11) u(t) = e u(0) −
tΔ
e(t−s)Δ (P (u · ∇)u)(s) ds
0
and the scheme introduced in §5.1.4. The semigroup etΔ generated by the
linear part is smoothing, i.e., u0 ∈ L2 implies that tm/2 ∂xm u(t) is bounded
in L2 (Td ) for every t > 0 and m ∈ N. Semi-linear means here that the
nonlinearity only contains terms with less derivatives than in the linear part,
i.e., for the Navier-Stokes equations first derivatives in (u · ∇)u compared
with the second order derivatives in Δu. More precisely, we will prove that
etΔ maps H m into H m+1 with a singularity t−1/2 and that P (u · ∇)u is a
bilinear map from H m+1 × H m+1 → H m . Then all assumptions following
(5.21) are satisfied and the local existence and uniqueness Theorem 5.2.22
will apply. Hence, the major step is to give a precise definition of P and to
investigate its analytic properties.
The projection on the divergence free vector fields. We define the
projection P via the solution v = P f of the system of PDEs
(6.12) v + ∇p = f, ∇ · v = 0,
with vj,k , fj,k , pk ∈ C for j = 1, 2. Plugging this into (6.12) yields
(6.13) v1,k + ik1 pk = f1,k , v2,k + ik2 pk = f2,k , ik1 v1,k + ik2 v2,k = 0.
188 6. The Navier-Stokes equations
where p(x, t) is 2π-periodic w.r.t. the xj . Then ∂xj p(x, t) = αj + ∂xj p(x, t)
and so vj,0 +αj = fj,0 . Thus, to a fj,0 we always find an αj such that vj,0 = 0.
Example 6.2.1. To illustrate the difference between Case i) and Case ii),
we consider the vector field
−2(x1 −π)2 −2(x2 −π)2 2 + tanh(x2 − π)
f (x1 , x2 ) = e .
0
In Figure 6.3 we show the different effects of choosing i) or ii).
Figure 6.3. Illustration of the difference between cases i) and ii), con-
cerning the boundary conditions for the pressure in the definition of the
projection u = P f for f from Remark 6.2.1, via a plot of the pressure
function. In the left panel we require a periodic pressure, giving a mean
flow in P f , while in the right panel we require zero mean flow, giving a
linear growth of the pressure p.
6.2. The equations on a torus 189
Proof. We find
Pf1,m = (Pk fk )k∈Z2 1,m ≤ sup Pk (fk )k∈Z2 1,m ≤ Cf1,m .
k∈Z2
The proof for 2,m works exactly the same. Using that Fourier transform is
an isomorphism between 2,m and Hperm , cf. Lemma 5.2.9, yields
The phase space and the fixed point argument. In the following we
will solve the Fourier transformed Navier-Stokes equations in two classes of
phase spaces, namely
div = {
1,m = Pu
u ∈ (1,m )d : u }, div = { = P u
u ∈ (2,m )d : u
2,m },
and the Navier-Stokes equations in physical space in the class of phase spaces
div,m
Hper = {u ∈ (Hper
m d
) : u = P u}.
In case of periodic boundary conditions we have P Δ = ΔP . Hence, for u
with u = P u the solution operator of the linearized Navier-Stokes equations
is given coordinate-wise by the solution operator etΔ of the linear diffusion
190 6. The Navier-Stokes equations
equation ∂t u = Δu. In one space dimension this operator has been discussed
a number of times, cf. Example 5.1.21 and Example 5.2.19. The statements
made above about this operator transfer line to line from R1 to Rd .
Theorem 6.2.4. The solution operator (e−|k| t )k∈Zd of the linearized Navier-
2
div,m
defines a C0 -semigroup in Hper . Moreover, for all m ∈ R and r ≥ 0 there
exists a C > 0, such that
(e−|k| t )k∈Zd u ≤C max(1, t−r/2 )
2
div udiv ,
1,m+r 1,m
−|k|2 t −r/2
(e div
)k∈Zd u ≤C max(1, t udiv ,
)
2,m+r 2,m
−r/2
e uH div,m+r ≤C max(1, t
tΔ
)uH div,m .
per per
as
d
d
d
d
uj ∂xj ul = uj ∂xj ul + ul ∂xj uj = ∂xj (uj ul ),
j=1 j=1 j=1 j=1
or equivalently
(6.16) uv T H m−1+δ ≤ CuH m vH m
for a δ > 0 and all m > m∗ for a m∗ ∈ R. To find the minimal m∗ we start
with the following lemma.
∈ R, r ≥ 0, and δ > 0 there exists a C > 0 such
Lemma 6.2.5. For all m
that
u ∗ v2,m ≤ C (
u2,m+r
v +
u
v 2,m+r
).
2, d
2 −r+δ 2, d
2 −r+δ
∈ ω1,β then u = F −1 u
Lemma 6.2.9. If u is analytic in a strip
Sβ = {z = (z1 , . . . , zd ) ∈ Cd : max |Imzj | < β}
j=1,...,d
k∈Zd s∈Rd
can be bounded independently of t.
6.2. The equations on a torus 193
Lemma 6.2.11. For every M > 0 there exists a T0 > 0 such that the map
t
−k2 t
e−k (t−τ ) P(k)ik(
2
F (
u)(k, t) = e (k, 0) +
u u∗u
T )(k, τ ) dτ
0
is a contraction in
u : Z × [0, T0 ] → C :
X = { u−u
lin X ≤ M },
lin (k, t) = e−k t u
2
uX = supt∈[0,T0 ]
where u(t)ω √ and where u (k, 0).
1, t
u ∗ vω √ ≤
Proof. It is easy to see, cf. Exercise 6.4, that uω √
v ω √
1, t 1, t 1, t
which implies
(6.18) u ∗ vX ≤ uX vX .
√ √ √
Using this estimate and t − s ≥ t − s shows that
t √
e−k (t−τ ) P(k)ik(
2
F (
u) − ulin X ≤ sup | T )(k, τ ) dτ |e t|k|
u∗u
t∈[0,T0 ] k∈Z 0
t √ √ √
|e−k P(k)ik( t|k| − τ |k| τ |k|
2 (t−τ )
≤ sup u∗u
T )(k, τ )e e e | dτ
t∈[0,T0 ] k∈Z 0
t √ √
|e−k P(k)ike t−τ |k| τ |k|
2 (t−τ )
≤ sup u∗u
( T )(k, τ )e | dτ
t∈[0,T0 ] 0 k∈Z
t √ √
sup |e−k t−τ |k| τ |k|
2 (t−τ )
≤C sup ike | dτ sup |(
u∗u
T )(k, τ )e |
t∈[0,T0 ] 0 k∈Z s∈[0,T0 ] k∈Z
1/2
≤CT0
u2X < ∞.
Thus, we proved that F maps the space X into itself if T0 > 0 is sufficiently
small. The proof of the contraction property works the same way.
Corollary 6.2.12. For all C > 0 there exists a T0 > 0 such that the solu-
tions u of the Navier-Stokes equations are analytic w.r.t. x in S√t ⊂ C for
all t ∈ [0, T0 ] if
u(0)1 ≤ C.
Remark 6.2.13. It is easy to see that Theorem 5.1.23 can be generalized to
ω1,β -spaces and the Fourier transformed Navier-Stokes equations such that
there is local existence and uniqueness in ω1,β , too.
Remark 6.2.14. In general it cannot be expected that in nonlinear prob-
lems the strip of analyticity is arbitrarily wide. An explicit, but typi-
cal, example is the z → tanh(z) equilibrium of the Allen-Cahn equation
194 6. The Navier-Stokes equations
6.2.3. Global existence in 2D. In order to prove the global existence and
uniqueness of solutions of the Navier-Stokes equations in a phase space X
we need a local existence and uniqueness result in X and a priori bounds for
the solutions in X. Then as explained already a number of times the local
existence and uniqueness result can be applied again and again to construct
a solution for all t ≥ 0.
Bounds for the L2 -norm. The L2 -norm of the solutions u of the Navier-
Stokes equations can be bound in every space dimension. By using integra-
tion by parts and the incompressibility dj=1 ∂j uj = 0 we obtain
d d
1 d
uj uj dx = uj ∂t uj dx
2 dt Td Td
j=1 j=1
d
d
d
= uj (∂x ∂x uj − ∂j p − u ∂x uj ) dx
Td j=1 =1 =1
= − (∂x uj )(∂x uj ) dx.
Td
Poincaré’s inequality, cf. Lemma 5.2.17, implies
Lemma 6.2.15. For all d ≥ 2 we have
1 d
u2L2 ≤ −u2L2 , and so u(t)L2 ≤ e−t u(0)L2 .
2 dt
Bounds for the H 1 -norm. In R2 with periodic boundary conditions also
the H 1 -norm
can be bound. By using integration by parts and the incom-
2
pressibility j=1 ∂j uj = 0, in R2 we find after some explicit calculation, cf.
Exercise 6.5, that
2 2 2
(6.19) (∂xj u )∂xj (um ∂xm u ) = 0.
T2 j=1 =1 m=1
6.2.4. The millennium problem. Figure 6.4 displays the Sobolev num-
bers for the local existence and uniqueness and the a priori estimates.
d=2 d=3
0 1 0 1/2
Figure 6.4. A priori estimates in L for d = 3 and H 1 for d = 2. Local
2
Partial results are already known, from which we list only the two abso-
lute basic ones.
• Jean Leray [Ler34] proved the global existence of so called weak so-
lutions, cf. §7.4.2, of the Navier-Stokes equations. These solutions
are very rough and they are not unique.
• For small initial conditions due to the linear stability of the ori-
gin one easily obtains an a priori estimate and thus it follows
limt→∞ u(t)H div,m (T3 ) = 0 with some exponential rate for every
per
m > m∗ = 1/2.
Remark 6.2.20. With a little more work it can be shown that A is a Hper 1
have
n −tA −A nt n C −a t n C
A e X→X ≤ Ae X→X ≤ e n ≤ n e−at ,
t t
such that e−tA : X → D(An ) for t > 0, and the semigroup is smoothing.
Remark 6.3.5. a) The concept of sectorial operators is very robust under
perturbations. Let A be a sectorial operator with A(λ−A)−1 ≤ C for all λ
in a chosen sector. Moreover, let B be a linear operator with D(B) ⊃ D(A)
satisfying
Bx ≤ εAx + Kx
with ε, K some constants. If εC < 1, then also A + B is a sectorial operator.
For a self-adjoint A it is sufficient that ε < 1. Hence, it is sufficient to check
the assumptions for the principal part of a given operator. Such estimates
can often be found in the existing literature.
b) The most essential remark is that, due to Parseval’s identity, in a
Hilbert space every self-adjoint operator which is bounded from below is a
sectorial operator.
6.3. Other boundary conditions and more general domains 199
Definition 6.3.7. We set Aα = (A−α )−1 for α > 0. The domain of defi-
nition is given by D(Aα ) = R(A−α ). We introduce X α = D(Aα ) equipped
with the norm
uX α = Aα uX .
4
1 C 1/2 3ε4/3
≤ uL2 ∇uL2 + Δu2L2
4 ε 4
≤ Δu2L2 + Cu2L2 ∇u4L2
using (6.23) with q = 4/3 and p = 4. Hence, we obtain
d
∇u2L2 ≤ a(t)∇u2L2 , where a(t) = 2Cu2L2 ∇u2L2 .
dt
From
1 d
u2L2 = −∇u2L2 ,
2 dt
cf. §6.2.3, follows u(t)L2 = e−t u0 L2 , and then
t
∇u(τ )2L2 dτ = u0 2L2 − u(t)2L2 .
0
Note, that this estimate does not imply a uniform bound for h(t) = ∇u(t)L2 ,
but its square-integrability. As a consequence,
t t
0≤ a(t) ≤ C sup u(τ )L22
∇u2L2 dτ < ∞
0 τ ∈(0,t) 0
Exercises
6.1. (a) Show that u : R2 → R2 with ∇ × u = 0 and ∇ · u = 0 is equivalent to the
Cauchy-Riemann differential equations for w(z) = u1 (x, y) − iu2 (x, y), z = x + iy.
(b) Sketch the flow belonging to w(z) = z 2 and calculate the associated pressure.
3
6.2. (d’Alembert Paradox) Let a > 0, U ∈ R3 , and φ(x) = 2x a
3 + 1
U, x
for x ∈ Ω = R3 \ Ba (0). Sketch u = ∇φ, show that div u = 0, and calculate the
drag f = − Ba (0) φn dS.
6.6. Prove the following weak form of the Gagliardo-Nirenberg inequality for (2π)d -
periodic functions. For d = 1, 2, 3 and δ > 0 we have
1−d/4−δ d/4+δ
uL4 (Td ) ≤ CuL2 (Td ) uH 1 (Td ) .
Hint: Use uLp ≤ C uLq for 1/p+1/q = 1 and q ∈ [1, 2] and the Hölder inequality
|ak bk ck | ≤ ( |ak |p1 )1/p1 ( |bk |p2 )1/p2 ( |ck |p3 )1/p3
k∈Z k∈Z k∈Z k∈Z
This is the first chapter of Part III of this book. Here and in the remainder
of this book we consider PDEs on unbounded domains. In order to avoid
dealing with far away boundaries, whose influence on the solutions in the
interior of the domain is small at least for a long time, we idealize the
large domain to an unbounded domain. For instance, instead of x from
some large interval (−L, L) we consider x ∈ R. From a didactic point
of view the consideration of unbounded domains has certain advantages.
Since we do not have to deal with boundary conditions which are often a
source of functional analytic difficulties, this idealization allows to explain
genuine PDE phenomena such as transport, diffusion and dispersion. Hence,
it allows us to keep the functional analytic tools at a minimum. Unbounded
domains are easy in this respect.
On the other hand, compared to PDEs over bounded domains there are
new fundamental and challenging open questions, mainly due to the fact that
PDEs on unbounded domains define dynamical systems with uncountably
many modes (degrees of freedom). In contrast to the situation of countably
many modes considered in Chapters 5 and 6, a separation of the uncountably
many modes into single modes is a highly singular action from a functional
analytic point of view, and therefore in general of little use. The recovery
of compactness by smoothing properties is no longer true, and therefore
finite-dimensional attractors in general cannot be expected.
To illustrate our point of view, the following example shows that also
for PDEs defined on a very large domain in space the interpretation as
countably many ODEs is no longer a big help.
205
206 7. Some dissipative PDE models
nonlinear reaction term u−u2 . Therefore, it brings together PDE with ODE
dynamics.
Inserting u(x, t) = v(t) into (7.1) gives the one-dimensional ODE
(7.2) v̇ = v − v 2 .
The 1D phase portrait shows that the fixed point v = 0 is unstable and
that the fixed point v = 1 is asymptotically stable. The term +v in the
KPP equation describes exponential growth for small v and the term −v 2
represents saturation. For instance, a population of animals or a chemical
reaction initially increases with some exponential rate until the growth is
saturated by the available food or the missing reactant. If v(0) > 0, then
limt→∞ v(t) = 1.
Before we combine the ODE dynamics coming from the reaction term
u − u2 with the dynamics coming from the diffusion term ∂x2 u we discuss the
modeling and the properties of linear diffusion in the next two subsections.
Under the assumption that all higher moments of ρ decay faster than τ for
τ → 0, in the limit τ → 0 we obtain the linear diffusion equation
(7.4) ∂t u = D∂x2 u.
Einstein derived the relation 2D = RT /(NA ν) where R is the gas constant,
T the temperature, NA the Avogadro number, and ν a friction coefficient.
A discrete random walk. We consider a two-dimensional rectangular
lattice, comprising the sites {(mδx, nδt) : m = 0, ±1, ±2, . . . ; n = 0, 1, 2, ...}.
A particle starting in x = 0 at a time t = 0 decides at each time nδt to
move an amount δx to the left or to move an amount δx to the right, both
possibilities with probability 1/2. Denote by p(m, n) the probability that
the particle is at the position mδx at the time nδt. Then p(0, 0) = 1 and
p(m, 0) = 0 for m = 0. Also, p(m, n + 1) = 12 (p(m − 1, n) + p(m + 1, n)),
hence
1
p(m, n + 1) − p(m, n) = (p(m − 1, n) − 2p(m, n) + p(m + 1, n)).
2
2
Now assume that (δx) /(δt) = 2D which corresponds to (7.3) above. Then
1 D
(p(m, n + 1) − p(m, n)) = (p(m − 1, n) − 2p(m, n) + p(m + 1, n)),
δt (δx)2
and sending δ → 0 again yields the linear diffusion equation (7.4).
Fourier’s law. Let u : R3 → R be the temperature inside a material body
and let V ⊂ R3 be a test volume with surface S. Then
d
u dV = − j · n dS = − div j dV ,
dt
V S V
where j = j(x, t) ∈ R3 is the heat flow. Since this is true for all test volumes
V we find
∂t u + div j = 0 .
It is reasonable to assume that the heat flow from warm to cold is propor-
tional to the negative temperature gradient, i.e., j = −D∇u. This so called
Fourier’s law again yields the heat equation
∂t u = D div ∇u = DΔu.
For t > 0 the general formula for the solutions of the linear diffusion
equation (7.5) is given by
∞
1 (x−y)2
(7.6) u(x, t) = √ e− 4t u(y, 0) dy.
4πt −∞
The derivation of this formula is given subsequently, but also by a different
method in §7.3. The existence of this integral is guaranteed for t > 0 if for
instance supy∈R |u(y, 0)| < ∞.
From (7.6) we immediately obtain the estimate
∞
1
sup |u(x, t)| ≤ √ |u(x, 0)| dx,
x∈R 4πt −∞
i.e., solutions to spatially localized initial conditions decay uniformly towards
zero with a rate t−1/2 . Since mass is conserved, i.e.,
∞ ∞
u(x, t) dx = u(x, 0) dx
−∞ −∞
for all t ≥ 0, this is how diffusion is expected to work. The conservation of
mass follows for instance with the use of the solution formula from
∞ ∞ ∞
1 − (x−y)2
u(x, t) dx = √ e 4t dx u(y, 0) dy
−∞ −∞ −∞ 4πt
∞
= 1 · u(y, 0) dy.
−∞
The decay happens in a universal manner. The initial condition
u(y, 0) = δ0 (y),
with δ0 the ”δ-distribution in x = 0”, cf. Example 5.2.2, leads to the self-
similar solution
1 − x2
(7.7) u(x, t) = √ e 4t .
4πt
The solution only exists for t > 0. This is a general rule. For an arbitrary
initial condition the diffusion equation cannot be solved backwards in time.
Since this solution is the starting point of the construction of the general
solution. In order to derive
solution formula it is also called fundamental
(7.7) we make the ansatz u(x, t) = v √xt and find v = − 12 ξv which is
ξ2
solved by v = ce− 4 with a constant c ∈ R. Since with u(x, t) = v √xt
also ∂x u = √1t v √xt is a solution of (7.5) we find (7.7).
In lowest order self-similar behavior is also observed for general spatially
localized initial conditions, namely,
A∗ x
u(x, t) = √ v √ + O(1/t) with v(ξ) = e−ξ /4
2
(7.8)
t t
210 7. Some dissipative PDE models
Finally for every t0 > 0 and x0 ∈ R the function (x, t) → u(x, t) can be
expanded in a convergent power series around (x0 , t0 ), i.e., u is an analytic
function and can be extended into the complex plane. See §6.2.2.
we estimate
u(t, u0 ) − u0 C 0 = sup |u(x, t) − u(x, 0)|
b
x∈R
∞ 1
x−y
= sup √ H √ (u(y, 0) − u(x, 0)) dy
x∈R −∞ t t
∞
√
= sup H(z) u(x − tz, 0) − u(x, 0) dz
x∈R −∞
≤ sup | . . . | dz + sup | . . . | dz = s1 + s2 .
x∈R |z|≤R x∈R |z|≥R
For a given ε > 0 we have to find a t0 > 0 such that for all t ∈ (0, t0 ) we
have s1 + s2 < ε. We can estimate
s2 ≤ 2 H(z) dz sup |u(x, 0)| < ε/2
|z|≥R x∈R
Since x → u(x, 0) is uniformly continuous for all ε > 0 there exists a δ > 0
such that for all y ∈ R with |y| < δ we have |u(x + y, 0) − u(x, 0)| < ε/2.
Choosing t0 > 0 so small that t0 R < δ we are done.
The deeper reason why X = Cb0 would not be a good choice is explained
in the following remark, but for a slightly simpler PDE.
u(x) = sin(x2 ), where the faster and faster oscillations for |x| → ∞ destroy
the uniform continuity w.r.t. x and t.
Remark 7.1.4. With (x, t) → u(x, t) a solution of (7.5) also (x, t) → u(x +
y, t) is a solution of (7.5), i.e., every solution shifted by y is a solution,
too. More abstractly, the solution operator and the translation operator
commute. With t → u(x, t) a solution, obviously every derivative t →
∂xn u(x, t) and every integral is a solution, too.
x2
Figure 7.2. The solutions u(x, t) = √ 1 e− 4t , ∂x u and its first integral
4πt
for t = 1.
where T (t) is the solution operator of the linear diffusion equation, i.e.,
∞
1 (x−y)2
(T (t)u0 )(x) = √ e− 4t u0 (y)dy,
4πt −∞
and where N (u)(τ ) represents the reaction term u(x, τ ) − u(x, τ )2 .
Proof. The proof of the local existence and uniqueness of solutions is based
on the contraction mapping principle and on the variation of constant for-
mula. Fix C1 > 0. For sufficiently small T0 > 0 the right-hand side of the
variation of constant formula (7.9) defines a contraction F in the complete
metric space
M = C 0 ([0, T0 ], {u(t) ∈ X : u(t) − T (t)u0 X ≤ C1 })
with the metric d(u, v) = u − vM induced by the norm
uM = sup u(t)X .
t∈[0,T0 ]
t
≤ sup dτ (N (u) − N (v))M
t∈[0,T0 ] 0
1
≤ T0 (1 + 2C2 )u − vM ≤ u − vM
2
for T0 (1 + 2C2 ) ≤ 1/2. Thus, for T0 > 0 sufficiently small there is a fixed
point u = F (u), which is a mild solution of the KPP equation.
Remark 7.1.8. The KPP equation is a semi-linear parabolic equation
[Hen81], i.e., the semigroup T (t) generated by the operator Λ = ∂x2 is
smoothing and the nonlinearity N contains only derivatives of strictly lower
order than in Λ. Like for the diffusion equation the solutions of the KPP
equation or more general of a semi-linear parabolic equation on the real line
are analytic for all t > 0. The proof is similar to the one given in §6.2.2.
7.1.4. The maximum principle. Second order scalar parabolic and ellip-
tic PDEs have a special property which helps a lot in their analysis, namely
the maximum principle.
Theorem 7.1.9. Let u1 , u2 be bounded solutions of the KPP equation (7.1)
in X for t ∈ [t0 , t1 ]. From u1 (x, t0 ) ≤ u2 (x, t0 ) for all x ∈ R it follows that
u1 (x, t) ≤ u2 (x, t) for all x ∈ R and all t ∈ [t0 , t1 ].
Idea of the proof. We show a slightly modified statement, namely that
u1 (x, t0 ) < u2 (x, t0 ) for all x ∈ R implies u1 (x, t) < u2 (x, t) for all x ∈ R
and all t ∈ [t0 , t1 ].
If this is wrong, then there exists a t∗ ∈ (t0 , t1 ) and a x0 ∈ R such that
u1 (x0 , t∗ ) = u2 (x0 , t∗ ). Since the solutions change smoothly in x and t in
this point we necessarily have ∂x u1 (x0 , t∗ ) = ∂x u2 (x0 , t∗ ) and generically
∂x2 u1 (x0 , t∗ ) < ∂x2 u2 (x0 , t∗ ). Then it follows
∂t (u2 (x0 , t∗ )−u1 (x0 , t∗ )) = ∂x2 (u2 (x0 , t∗ ) − u1 (x0 , t∗ ))
+ (u2 (x0 , t∗ )−u1 (x0 , t∗ ))−(u22 (x0 , t∗ )−u21 (x0 , t∗ ))
= ∂x2 (u2 (x0 , t∗ ) − u1 (x0 , t∗ )).
In the generic case ∂x2 (u2 (x0 , t∗ ) − u1 (x0 , t∗ )) > 0, we have again u2 (x0 , t) >
u1 (x0 , t) for t ∈ [t0 , t1 ], i.e., u1 can never be larger than u2 . For a complete
proof see for instance [RR04, Theorem 4.26].
Lemma 7.1.10. Solutions that start at t0 with values in [0, 1] will stay with
their values in [0, 1] for t ≥ t0 , i.e., 0 ≤ u(x, t0 ) ≤ 1 for all x ∈ R implies
0 ≤ u(x, t) ≤ 1 for all x ∈ R and all t ≥ t0 .
Proof. For all initial conditions with values in [0, 1] the local existence and
uniqueness theorem 7.1.7 guarantees the existence for all t ∈ [0, T0 ] with for
instance T0 = 1/6. By the last lemma we have 0 ≤ u(x, T0 ) ≤ 1. Then
we start the KPP equation again, but now with initial condition u(T0 ).
The local existence and uniqueness theorem guarantees for the same reason
solutions for all t ∈ [T0 , 2T0 ]. Repeating the argument shows the assertion.
Theorem 7.1.12. For u0 ∈ X with 0 ≤ u0 (x) ≤ 1 there exists a unique
solution u ∈ C([0, ∞), X) of the KPP equation with u|t=0 = u0 and X =
0
Cb,unif (R, R).
Proof. For all initial conditions with values in [0, 1] the local existence and
uniqueness theorem 7.1.7 guarantees the existence for all t ∈ [0, T0 ] with for
instance T0 = 1/10. By the last lemma we have 0 ≤ u(x, T0 ) ≤ 1. Then
we start the KPP equation again, but now with initial condition u(T0 ).
The local existence and uniqueness theorem guarantees for the same reason
solutions for all t ∈ [T0 , 2T0 ]. Repeating the argument shows the assertion.
By having a look at the phase portrait in Figure 7.4 we see that the only
equilibria having only values in [0, 1] are the fixed points u = 0 and u = 1.
u’
0
−1
−1 0 1 2
u
Figure 7.4. The phase portrait for the stationary solutions of the
KPP equation.
Definition 7.1.13. A fixed point u∗ is called stable for the KPP equation
in X if for any ε > 0 there is a δ > 0 such that u0 − u∗ X < δ implies
u(t, u0 ) − u∗ X < ε for all t ≥ 0. Otherwise, it is called unstable. A stable
fixed point is called asymptotically stable in X if additional limt→∞ u(t, u0 ) =
u∗ holds.
For ODEs, in most cases the eigenvalues of the linearized system deter-
mine stability or instability, cf. Theorem 2.3.4. For u∗ = 0 we find the linear
operator (∂x2 + 1)· which has the eigenfunctions eikx for k ∈ R and the spec-
trum (−∞, 1]. For u∗ = 1 we find the linear operator (∂x2 − 1)· which has the
eigenfunctions eikx for k ∈ R and the spectrum (−∞, −1]. Hence, we expect
that u∗ = 0 is unstable, whereas u∗ = 1 is expected to be asymptotically
stable. However, instead of using these spectra, here we use the maximum
principle to study stability and instability.
Proof. Choose δ = min{ε, 1/10} and let u0 ∈ X with u0 − u∗ X < δ. Let
v− = 1 − δ and v + = 1 + δ such that v− ≤ u0 (x) ≤ v+ for all x ∈ R. By the
maximum principle we have for the associated solutions
v− (t) ≤ u(x, t) ≤ v+ (t)
for all x ∈ R and t ≥ 0. Since limt→∞ v− (t) = limt→∞ v+ (t) = 1 we can
conclude
lim u(t, u0 ) − u∗ X = 0,
t→∞
i.e., the asymptotic stability of u∗ .
2
w 1
w
0
−1
−2
1
1
0.5
v
0.5
v
0
0 5 10 0
x−t 0 5 10 15
x−2t
Figure 7.6. The phase portraits for the front solutions for the KPP
equation for c = 1 (left) and c = 2 (right), with the associated
fronts.
Proof. Let uf be a front solution with velocity c, and u0 (x) = min(uf (x), δ/2).
Then uf − u0 X < δ with u(t, uf )[x] = uf (x − ct), but limt→∞ u(t, u0 ) = 1
due to the maximum principle, since δ/2 ≤ u0 (x) ≤ 1, and hence
lim u(t, uf ) − u(t, u0 )X = 1 > ε
t→∞
independent of how small δ > 0 is chosen. See Figure 7.8.
uf
δ/2
0
0
Figure 7.8. Instability of front solutions w.r.t. perturbations in Cb,unif .
7.1. The KPP equation 221
Remark 7.1.18. On the other hand, by the maximum principle the front
solutions are stable w.r.t. smaller sets of perturbations. If the small pertur-
bation v of the front solution uf is contained between two translates of the
same front then by the maximum principle it will stay there for all times.
In detail, if
uf (x − x− ) ≤ uf (x) + v(x, 0) ≤ uf (x + x+ )
for all x ∈ R, then
uf (x − ct − x− ) ≤ uf (x − ct) + v(x, t) ≤ uf (x − ct + x+ )
for all t ≥ 0 and all x ∈ R. See Figure 7.9.
Definition 7.1.19 makes sense due to the fact that in infinite dimensions
there are infinitely many non-equivalent norms. For instance, the fronts are
2
0
(X1 , Cb,unif ) stable, where uX1 = supx∈R |u(x)ex |, see, e.g., [BK00, §9.3].
Note that e−x decays much faster than the difference of two fronts which
2
0
in Cb,unif as t → ∞. In summary, we have asymptotic (X, X) stability of
u = 0, where uX = supx |u(x)e−βx |, i.e., u0 X ≤ δ ⇒ u(t)X → 0 as
t → ∞. Note that this does not mean that u stays bounded. What happens
is that for c > 2 and hence β < 0 the (growing) mass is transported to
−∞ where it vanishes in the X-norm. This is sometimes called convective
stability.
Remark 7.1.21. The idea of weighted variables from Example 7.1.20 can
be used to prove the nonlinear stability of the front solutions uf for |c| > 2.
In a co-moving frame ξ = x − ct the deviation v = u − uf from the front
satisfies
∂t v(ξ, t) = ∂ξ2 v(ξ, t) + c∂ξ v(ξ, t) + (1 − 2uf (ξ))v(ξ, t) − v(ξ, t)2 .
The weighted variable w(ξ, t) = v(ξ, t)e−βξ satisfies
∂t w(ξ, t) = Lw w(ξ, t) − eβξ w(ξ, t)2 ,
with
Lw w(ξ, t) = ∂ξ2 w(ξ, t) + (2β + c)∂ξ w(ξ, t) + (β 2 + cβ + 1 − 2uf (ξ))w(ξ, t).
Since uf ≥ 0 we can find a β < 0 such that w = 0 is linearly stable, e.g., in
H 1 . However, the nonlinearity is not well defined in H 1 due to the factor
eβξ . Therefore, the equation for v and w have to be combined, namely
∂t v(ξ, t) =Lv v(ξ, t) + (2 − 2uf (ξ))eβξ w(ξ, t) − v(ξ, t)2 ,
∂t w(ξ, t) =Lw w(ξ, t) − v(ξ, t)w(ξ, t),
with
Lv v(ξ, t) = ∂ξ2 v(ξ, t) + c∂ξ v(ξ, t) − v(ξ, t).
Since the semigroups generated by Lv and Lw decay with some exponential
rate and since (2−2uf (ξ))eβξ is bounded, the asymptotic stability of v=w=0
follows, i.e., uf is asymptotically stable w.r.t. perturbations v which are
small w.r.t. the norm vH 1 + veβx H 1 . This idea goes back to [Sat76].
of two parts, namely the diffusion term ∂x2 u and the nonlinear reaction term
u − u3 . Inserting u(x, t) = v(t) into (7.13) gives the one-dimensional ODE
(7.14) v̇ = v − v 3 .
The phase portrait shows that the fixed point v = 0 is unstable and that the
two other fixed points v = ±1 are asymptotically stable. The last two fixed
points can be interpreted as two stable phases. Hence, in the Allen-Cahn
equation both phases have the tendency to grow. At their interfaces both
phases try to move the interfaces towards the other phase. The Allen-Cahn
equation therefore describes interface motion between two stable phases with
applications in material science, image processing, biology and geology.
The local existence and uniqueness of solutions follows exactly as for
the KPP equation. As a consequence of the dynamics of (7.14) and the
maximum principle we have the following global existence und uniqueness
result for mild solutions.
Theorem 7.2.1. For u0 ∈ Cb,unif 0 there exists a unique mild solution u ∈
C([0, ∞), Cb,unif ) of the Allen-Cahn equation with u|t=0 = u0 .
0
-1
Solutions of the Allen-Cahn equation with more than one interface are
the next non-trivial set of solutions which can be understood completely.
These solutions consist of almost flat regions, where they are close to the
exponentially stable states ±1. In between there are the interface regions,
where the solutions connect the almost flat regions.
We define classes of functions with m interfaces
Im = {u ∈ Cb0 is an interface function with m zeros, lim u(x) = 1 or − 1}.
x→±∞
Then from the analysis in [CP89, CP90] it is known that the interfaces
annihilate. Moreover, it is known that this happens in a very robust way,
i.e., small perturbations do not change the dynamics. With the help of the
maximum principle the following global result can be obtained
lim u(·, t) ∈ I0 , if u(·, 0) ∈ I2m , lim u(·, t) ∈ I1 , if u(·, 0) ∈ I2m+1 ,
t→∞ t→∞
with I0 = {−1, 1}. The time for the annihilation of the interfaces goes
exponentially (O(eCL )) with the distance L of the interfaces. It has been
shown that by separating the interfaces in the right way, i.e., making the
distances bigger and bigger for x → ±∞ it is possible to have an annihilation
of interfaces for all times, i.e., infinitely many annihilations.
Comparing to the Allen-Cahn equation on a bounded domain where it
has a finite dimensional attractor, cf. §5.3, we see that large time asymptotics
over unbounded domains can be fundamentally different from the bounded
domain case.
is finite.
Proof. It will turn out in §7.3.4 that the Fourier transform of the
δ-distribution is the constant function with value 1. With this knowledge
the formal argument is as follows.
−1 1 −ik·y
(F F u)(x) = e ik·x
e u(y) dy dk
Rd (2π)d Rd
1 ik·(x−y)
= e dk u(y) dy
Rd (2π)d Rd
= δ(x − y)u(y) dy = u(x).
Rd
In order to make this formal argument rigorous we use an approximation
of the δ-distribution and the roll-off formula. Hence, we set v(k) = e−|k| /2
2
and let vr (k) = v(k/r) for r > 0. For r → ∞ and fixed k ∈ Rd we have
vr (k) → 1 monotonically from below.
For u ∈ S we set ux (k) = u(k − x) and find
ik·x
e u (k)vr (k) dk = u/
−x (k)vr (k) dk = u−x (k)
vr (k) dk
Rd Rd Rd
= u−x (k)rd v(rk) dk = u−x (m/r)v (m) dm
Rd Rd
= (2π)−d/2 u−x (y/r)v(y) dy
Rd
230 7. Some dissipative PDE models
by Lemma 7.3.10 and since F (e−|x| /2 )(k) = (2π)−d/2 e−|k| /2 , cf. Exercise
2 2
7.12. Now
ik·x
e u (k)vr (k) dk → (k) dk as r → ∞
eik·x u
Rd Rd
by the Lebesgue dominated convergence theorem 7.3.3 since the integrand
is dominated by |u| ∈ L1 . Similarly,
−d/2 −d/2
(2π) u−x (y/r)v(y) dy → (2π) u(x) v(y) dy = u(x)
Rd Rd
=(2π)d (x)g(x) dx = (2π)d
u u, vL2 .
Rd
Since S is dense in
L2 we find that F can be extended to an isomorphism
in L which is isometric up to a factor (2π)d .
2
which can be obtained in the same way. We have the following estimate
Lemma 7.3.19. (Young’s inequality for convolutions) For 1 ≤ p ≤ ∞
and f ∈ Lp (Rd ), g ∈ L1 (Rd ) we have f ∗g ∈ Lp and f ∗gLp ≤ f Lp gL1 .
Proof. For p = ∞ we have Rd |f (x − y)g(y)| dy ≤ f L∞ gL1 < ∞ a.e..
Thus let 1 ≤ p < ∞. The map (x, y) → f (x − y)g(y) is measurable as a
product of measurable functions. By Hölder’s inequality we obtain
|f (x − y)g(y)| dy = |f (x − y)||g(y)|1/p|g(y)|1/p dy
Rd Rd
1/p 1/p
≤ |f (x − y)| |g(y)| dy
p
|g(y)| dy
Rd Rd
see Exercise 7.12, which yields the assertion. We have the following com-
mutative diagram
G(t)
u0 −→ u(·, t)
−1
↓F ↑F
G(t)
0
u −→ u
(·, t)
for all φ ∈ D and t → 0. To show this let ε > 0 and φ ∈ D. Then there
exists a δ > 0 with |φ(x) − φ(0)| < ε/2 for |x| < δ, and thus
G(x, t)φ(x) dx − φ(0) = G(x, t)(φ(x) − φ(0)) dx
R
R
≤ . . . dx + . . . dx < ε/2 + ε/2
|x|<δ |x|>δ
for t > 0 sufficiently small, where we used in the first integral that
R G(x, t) dx = 1.
7.3.5. Sobolev spaces. For the handling of nonlinear PDEs, Lp -spaces are
of no use since they are not closed under multiplication. In §5.2.1 we defined
the Sobolev spaces H m (Ω) over some bounded domain Ω, and in §5.2.2 we
showed that for suitable boundary conditions there is an easy alternative
characterization using Fourier series, see in particular Lemma 5.2.9. Here
we use Fourier transform to study the Sobolev spaces over Rd . Again we
usually do not distinguish real-valued from complex-valued or vector-valued
functions.
Definition 7.3.26. For m ∈ N define the Sobolev space
W m,p (Rd ) := {u ∈ Lp (Rd ) : ∂xα u ∈ Lp (Rd ) for |α| ≤ m},
where ∂xα with multi-index α denotes the distributional derivative of order
α. It is equipped with the norm
⎛ ⎞1/p
uW m,p = ⎝ ∂xα upLp ⎠ .
|α|≤m
For m ∈ N0 this definition coincides with the previous one. Next, sim-
ilar to Lemmas 5.2.3 and 5.2.24 we obtain the following embedding and
multiplication results.
236 7. Some dissipative PDE models
Proof. For ρ(k) = (1 + |k|2 )1/2 we have ρ−m ∈ L2 if m > d/2, and thus
which implies
u∞ ≤ C
uL1 ≤ C
uL2m ≤ CuH m .
uvH m ≤C
u ∗ vL2m ≤ C(ρm u
L2
v L1 +
uL1 ρm vL2 )
uL2m
≤2C v L2m ≤ 2CuH m vH m ,
If viscous effects are taken into account, one obtains the viscous Burgers
equation
1
(7.26) ∂t u = ν∂x2 u − ∂x (u2 ),
2
where ν > 0 is the viscosity parameter. This equation also occurs as a
modulation equation describing the behavior of, e.g., the free surface in
viscous flows [Uec07], or the evolution of wave numbers in stable periodic
pattern [DSSS09]. On the real line the parameter ν > 0 can be scaled to
1, hence we consider ν = 1 in the following.
Remark 7.4.2. a) Like for the KPP equation the maximum principle, cf.
Theorem 7.1.9, holds for the viscous Burgers equation, too. With the nota-
tion from page 215 we have
0 1
∂t u2 (x0 , t1 ) − u1 (x0 , t1 )
0 1
= ∂x2 u2 (x0 , t1 ) − u1 (x0 , t1 )
0 1
− ∂x u2 (x0 , t1 ) − u1 (x0 , t1 ) (u2 (x0 , t1 ) + u1 (x0 , t1 ))
0 1
− u2 (x0 , t1 ) − u1 (x0 , t1 ) ∂x (u2 (x0 , t1 ) + u1 (x0 , t1 ))
0 1
= ∂x2 u2 (x0 , t1 ) − u1 (x0 , t1 ) ,
and so the same argument applies.
b) The Burgers equation possesses the so called boost invariance u → u + c
and x → x − ct, i.e., if we add a constant to a solution and go into a co-
moving frame, then we again obtain a solution. Therefore, for a number
of
purposes it is sufficient to restrict to, e.g., solutions with zero mean, i.e.
R u(x, t) dx = 0.
Theorem 7.4.4. For all C0 > 0 there exists a T0 > 0 such that the following
holds. For u0 ∈ X with u0 X ≤ C0 . Then there exists a unique solution
u ∈ C([0, T0 ], X) of the Burgers equation (7.26) with initial datum u0 .
t
≤ sup 1 + (t − τ )−1/2 (N (u) − N (v))(τ )X dτ
t∈[0,T0 ] 0
t
≤ sup 1 + (t − τ )−1/2 dτ (N (u) − N (v))M
t∈[0,T0 ] 0
1
1/2
≤ T0 + 2T0 C2 u − vM ≤ u − vM
2
for T0 > 0 sufficiently small. Hence, there is a unique fixed point u of the
map F which is a mild solution of the Burgers equation.
Therefore, due to Lemma 5.2.10 this bounded linear operator for t > 0 can
be extended to a bounded linear operator from the dense subspace to the
full space by
T (t)∂x u = lim ∂x T (t)un
n→∞
for u ∈ Cb,unif
0 , and un ∈ Cb,unif
1 with un → u in Cb,unif
0 for n → ∞.
Remark 7.4.6. Like the KPP equation, the Burgers equation is a semi-
linear parabolic equation [Hen81], i.e., the semigroup T (t) generated by
the operator −A = ∂x2 is smoothing and the nonlinearity N contains only
derivatives of lower order than A. As already explained in §5.2.3 the scheme
and the estimates of the last proof work for equations of the form
with f a smooth function. It is also the same scheme and the same estimates
as for the Navier-Stokes equations in §6.2.1. It does not work for
1, for x ≥ t/2,
u(x, t) =
0, for x < t/2,
is a weak solution of the inviscid Burgers equation (7.23) since
∞ ∞ ∞
1 2 1
u∂t φ + u ∂x φ dx dt = ∂t φ + ∂x φ dx dt
0 R 2 0 t/2 2
∞ 2x ∞
1
= ∂t φ(x, t) dt dx − φ(t/2, t) dt = 0
0 0 2 0
for all φ ∈ C0∞ (R × R+ ).
As the following example together with Exercise 7.4.9 shows, weak so-
lutions are in general not unique.
1, for x ≥ 0,
Example 7.4.10. Consider u0 (x) = For this initial con-
0, for x ≤ 0.
dition the characteristics leave a complete region in the (x, t)-plane empty.
See Figure 7.13. There are many possibilities to fill this region. A physically
realistic solution will be a so called rarefaction wave
⎧
⎨ 1, for x ≥ t,
u(x, t) = x/t, for x ∈ [0, t],
⎩
0 for x ≤ 0.
We remark that u(x, t) = limε→0 uε (x, t), where uε solves ∂t uε = ε∂x2 uε −
uε ∂x uε . Thus, u is called viscosity solution.
Figure 7.13. The characteristics do not enter parts of the (x, t)-
plane and create a rarefaction wave.
7.4. The Burgers equation 243
0, for x ≥ 0,
lim vh,ν (ξ) =
ν→0 1, for x < 0,
i.e., the weak shock from Example 7.4.9 occurs as a vanishing viscosity limit.
With the help of the maximum principle a number of stability results
0
can be established. Recall that we have chosen X = Cb,unif (R, R) as phase
space.
Theorem 7.4.11. The fixed points u∗ ∈ X defined by u(x, t) = u∗ ∈ R are
stable, but not asymptotically stable.
Proof. For given ε > 0 choose δ = ε > 0. Then by the maximum principle
from u0 − u∗ X < δ, i.e., u∗ − δ < u0 (x) < u∗ + δ for all x ∈ R, we have
that u∗ − δ < u(x, t) < u∗ + δ for all x ∈ R and t > 0, i.e., u(t, u0 ) − u∗ X <
ε which shows the stability. Since there are infinitely many fixed points
{u∗ + δ : |δ| < δ0 } in every δ0 -neighborhood of u∗ in X, the fixed point u∗
cannot be asymptotically stable.
244 7. Some dissipative PDE models
Remark 7.4.12. Like for the KPP equation again the maximum principle
can be used to prove the stability of the traveling waves under small localized
perturbations by confining the perturbation between two translates of the
0
traveling wave. The instability in Cb,unif comes from the fact that there are
nearby traveling waves with a different velocity.
Figure 7.14. Shocks are robust against small spatially localized per-
turbations. For the shock from Example 7.4.9 these perturbations are
transported by the characteristics into the shock with relative velocities
±1/2, where they finally vanish. This idea can also be used for analytic
stability proofs of viscous shocks, cf. [Kap94b].
1
(7.28) ∂t φ = ν∂x2 φ − (∂x φ)2 .
2
Next we let φ = g(ψ) with g to be chosen such that the quadratic term in
(7.28) vanishes. We obtain
1
g (ψ)∂t ψ = νg (ψ)(∂x ψ)2 + νg (ψ)∂x2 ψ − (g (ψ)∂x ψ)2
2
which we rewrite as
1
g (ψ)(∂t ψ − ν∂x2 ψ) = (νg (ψ) − g (ψ)2 )(∂x ψ)2 .
2
The condition νg (ψ) − 12 g (ψ)2 = 0 leads to g(ψ) = −2ν log ψ. Thus,
x
1 ∂x ψ(x, t)
ψ(x, t) = exp − u(y, t) dy and u(x, t) = −2ν ,
2ν a ψ(x, t)
where a ∈ R is arbitrary. Moreover, we have by construction ∂t ψ = ν∂x2 ψ.
Denoting the Cole-Hopf transformation by T we thus have the commutative
7.4. The Burgers equation 245
diagram
u0 nonlin. evolution in the Burgers eq. u(·, t)
−−−−−−−−−−−−−−−−−−−−−−−−−−−→ −1
↓T ↑T
ψ0 linear evolution in the heat equation ψ(·, t)
−−−−−−−−−−−−−−−−−−−−−−−−−−−→
Example 7.4.13. We consider the explicit solution
x
√ 1
e−y /4 dy,
2
(7.29) ψ(x, t) = 1 + zErf(x/ νt) where Erf(x) = √
4π −∞
which describes the diffusive mixing of the asymptotic states 1 for x < 0
and 1 + z for x ≥ 0. Here Erf is related to the more standard definition of
the so called error function
x
2 1
e−t dt via Erf(x) = (1 + erf(x/2)).
2
erf(x) = √
π 0 2
We obtain
√
e−x /4νt
2
∂x ψ νz
(7.30) u(x, t) = −2ν = −√ √ .
ψ πt 1 + zerf(x/ νt)
This is illustrated in Figure 7.15 with the small diffusion coefficient ν = 0.01.
Initially the wave steepening by nonlinear transport dominates. However, for
t > 0 the solution is always smooth, and as t → ∞ the solution decays to zero
in L∞ with a rate O(t−1/2 ). Figure 7.16 shows the so called renormalized
asymptotic profiles
√
√ √ νz e−x /4ν
2
50
40
50
1
1 40 30
30
20
0
20 t
0
10
10
-3 0
-3 0 0 3 6 x
3 9
1
√
ν=0.01 tu
ν=0.2
ν=1
0.5
0 √
−5 0 5 tx
Exercises
7.1. Derive the two-dimensional diffusion equation ∂t u = Δu, u = u(x, y, t), from
a two-dimensional random walk.
7.2. Compute u = u(z, t) with z = xr + ixi using the solution formula for linear
diffusion. Show the convergence of the integral for every t > 0 and z ∈ C.
7.3. Prove that the solutions u(t, u0 ) of the linear diffusion equation satisfy the
fundamental property of a dynamical system u(t + s, u0 ) = u(t, u(s, u0 )).
7.4. Use the maximum principle to give an alternative proof of (E1) on page 213.
7.5. Prove that if u ∈ C([0, T0 ], Cb,unif
2
) ∩ C 1 ([0, T0 ], Cb,unif
0
) solves the variation of
constant formula, then u is also a classical solution of the KPP equation.
7.6. Let R = {(v, w) : 0 < v < 1, −kv < w < 0}. Show that for suitable k the
vector field f = (w, −cw − v + v 2 ) for (7.10) points inwards on ∂R. Use this and
the Poincaré-Bendixon theorem to prove that for c = 2 there exists a monotonic
front connecting u = 1 and u = 0 for the KPP equation.
7.7. a) Discuss for which α, β there exist similarity solutions u(x, t) = tα v(x/tβ ) of
∂t u = ∂x2 u + f (x), x ∈ R, where (i) f (x) = 0, (ii) f (x) = 1, (iii) f (x) = x.
b) Find α, β such that ∂t u + up ∂x u + μ∂x3 u = 0 is invariant under u → au, x → aα x,
t → aβ t.
7.8. a) Solve explicitly ∂t u = ∂x2 u with IC u0 (x) = x2 .
b) Use a) to calculate explicitly s2 e−s ds.
2
∈ Cb0 .
7.11. Show that u ∈ L1 implies u
7.12. a) For u(x) = e−x √1 e−k /2
2 2
/2
(k) =
show u 2π
by solving the Fourier trans-
(k) = e−k t , t > 0.
2
formed ODE which is satisfied by u. Calculate u(x) for u
7.13. Consider the linear heat equation with initial conditions
a, for x < 0,
u0 (x) =
b, for x ≥ 0.
√ x
Prove that u(x, t) = a+(b−a)erf(x/ t), where erf(x) = √14π −∞ e−y /4 dy.
2
r
7.14. Prove that Fourier transform is an isomorphism between Hm and Hrm .
7.15. Let g(x) = e−x
2
/2
. Find g ∗ g.
7.16. For f ∈S, prove the Heisenberg uncertainty principle xf L2 kfL2 ≥
2 f L2 . Hint. Use Parseval, the relation between differentiation and Fourier trans-
1 2
7.17. The scaling of a distribution is, as usual, defined by applying the scaling to
the test function. I.e., for k = 0, T ∈ D and φ ∈ D we set Sk1 T (φ) = T (φ(k·)),
Sk2 T (φ) = k1 T (φ(·/k)). Show that Sk1 δ = Sk2 δ.
D
7.18. Prove that (a) If un ∈ C 0 (R) and un → u uniformly in x, then un → u for
n → ∞.
D
(b) If un ∈ L2 (R) and un − uL2 → 0 , then un → u for n → ∞.
D D
(c) sin nx → 0 and n2 sin nx → 0 for n → ∞.
2 dn −x2
7.19. The Hermite functions are ψn (x) = (−1)n ex /2
e , n ∈ N0 .
dxn
i) Sketch ψ0 , ψ1 and ψ2 .
ii) Show the recursions ψn (x) = xψn (x) − ψn+1 (x) and F(ψn ) (k) = kψn (k) −
iψn+1 (k).
iii) Use i) to inductively show ψn = λn ψn with λn = (−i)n .
Remark. iii) means that all the ψn√ are eigenfunctions of F. They form a orthogonal
system in L2 (R),
ψn , ψm = 2n n! πδnm , and the normalized hn = ψn /ψn form
a complete orthonormal system in L2 (R), which is quite useful in various fields,
e.g., quantum mechanics.
7.20. Consider the scalar first order PDE
a(x, t, u)∂t u + b(x, t, u)∂x u = c(x, t, u).
Make the ansatz v(s) = u(x(s), t(s)) and find a system for (x(s), t(s), v(s)).
7.21. Show that u(x, t) = u0 (x0 ) = u0 (x − tu(x, t)) is an implicit representation
of the solutions of ∂t u + u∂x u = 0. Show by differentiation of this representation
w.r.t. x that for t = t∗ = −1/(inf x0 ∈R u0 (x0 )) a shock occurs, i.e., that u = u(x, t)
is no longer differentiable w.r.t. x.
7.22. Given ul > ur and u0 (x) = ul for x ≤ 0
and u(x) = ur for x > 0 calculate the
ul , for x ≤ αt,
shock speed α of the weak solution u(x, t) = of the inviscid
ur , for x > αt
Burgers equation.
7.23. Consider for a general scalar conservation law ∂t u = −∂x (f (u)) the weak
solution from Exercise 7.22. Derive the so called Rankine-Hugoniot jump condition
α = [f ]/[u] := (f (ur ) − f (ul ))/(ur − ul ) for the velocity of the shock.
7.24. Discuss the phase portrait of −cv = v − vv in the (v, v )-plane. For v− = 1
and v+ = 0 show that c = 1/2 and that v(ξ) = (1 + eξ )−1 is the explicit solution.
7.25. Consider the scalar conservation
law ∂t u + u3 ∂x u = 0, x ∈ R, t > 0. Find
1, 0 < x < 1,
the solution for u(x, 0) = g(x) =
0, else.
the viscous Burgers equation ∂t u = ν∂x u − 2 ∂x (u ) with u(x, 0) =
2 1 2
7.26. Solve
−1, for x < 0,
H(x) = Give a sketch of the solution.
1, for x ≥ 0.
Chapter 8
Three canonical
modulation equations
In this chapter we consider another three scalar equations on the real line,
namely the Nonlinear Schrödinger (NLS) equation, the Korteveg-de Vries
(KdV) equation, and the Ginzburg-Landau (GL) equation. These equations
are the three most important so called modulation equations, sometimes
also called amplitude or envelope equations. By multiple scaling analysis
the NLS equation can be derived in order to describe the evolution of an
envelope of a spatially and temporarily oscillating wave packet. This will
be discussed in Chapter 11. Similarly, the KdV equation can be derived
for the description of long waves in dispersive media, see Chapter 12, and
the GL equation can be derived to describe the evolution of the envelope of
spatio-temporal pattern in dissipative systems, see Chapters 10.
Besides the basic theory and the occurring phenomena for each equation,
we introduce some general concepts, which are useful for the analysis of these
equations but also in other circumstances. These are for instance the method
of stationary phase in §8.1.4, the use of uniformly local Sobolev spaces in
§8.3.1, and a concept of attractors on unbounded domains in §8.3.4.
As already said, each of the equations considered in this Chapter plays a
very important role in the description of various phenomena in more compli-
cated systems, and thus this chapter is partly a preparation for the remain-
der of this book. On the other hand, as for the equations from Chapter 7,
for each of these equations there exists many papers, and often whole books
entirely devoted to the particular equation. Again we keep the exposition
rather brief, and give a list of further reading at the end of the chapter.
249
250 8. Three canonical modulation equations
a) α = 1, k = 1 b) α = −1, k = 1 c) α = −1, k = 2
1 2 1 2 1 2
0 0 0
1 1 1
-1 -1 -1 t
-4 -2 0 -4 -2 0 0 -4 -2 0 0
0 2 4 2 4 2 4 x
Figure 8.1. Nonlinear oscillations (real part) for the NLS equation,
with r = 1. a) shows the defocusing case, where all waves travel
left, while b), c) show the focusing case, where waves can travel
left or right.
Proof. Proceeding
ik2 t ikxexactly as in Example 7.3.20 it remains to calculate
G(x, t) = R e e dk. We have
2 ix
∂x G(x, t) = ikeik t eikx dk = − G(x, t).
R 2t
This differential equation is solved by G(x, t) = Ce−ix /4t . The constant C
2
Remark 8.1.2. For the transport equation and the linear wave equation we
obtain ω(k) = ck with a constant c ∈ R and hence the group velocity ω (k) =
c is bounded and the transport equation and the linear wave equation can
0
be solved in Cb,unif (R, R).
It turns out that X = L2 (R, C) is a good choice for the handling of the
linear Schrödinger equation as a dynamical system.
U (t + s, u0 ) =F −1 (k → eik
2 (t+s)
0 (k))
u
=F −1 (k → e ik2 t
(e ik2 s
0 (k))) = U (t, U (s, u0 ))
u
As already said, we come back to the NLS equation in Part IV, while
here we close with some bounds for oscillatory integrals which are used in
the derivation of dispersive estimates.
Remark 8.1.9. Since no compactness argument has been used in the proof
of Lemma 8.1.8 values a = −∞ and b = ∞ are allowed if the integrals exist.
We use Lemma 8.1.8 and Remark 8.1.9 to confirm the dispersive estimate
(8.7).
such that φ(k) = −k 2 + kx/t. The integral is estimated for every fixed
ξ = x/t. Since φ (k) = −2 we thus have supξ∈R |G(ξt, t)| ≤ Ct−1/2 . From
supξ∈R |G(ξt, t)| = supx∈R |G(x, t)| it follows
For the existence of solutions un+1 of the linear equation (8.14) for a given
function un , see, e.g., [Paz83, Chapter 5]. Next the convergence of the
sequence (vn )n∈N , with vn = un+1 − un , towards zero and the boundedness
of the sequence (un )n∈N is shown. Then we have, cf. [Kat81, Paz83,
KPV91],
Theorem 8.2.1. Let u0 ∈ X = H s (R, C) with s > 3/2. There exist a
T0 > 0 and a unique local solution u ∈ C([0, T0 ], X) of the KdV equation
(8.13) with u|t=0 = u0 .
8.2.1. The solitary wave. The starting point of the KdV history is the
observation of a solitary wave in 1834 by the engineer Scott Russell in a
canal between Edinburgh and Glasgow. The solitary wave was created by a
stopping boat and was of 40 cm height and 10 m length. It traveled with a
speed of 16 km/h for many kilometers through the canal without changing
significantly its shape. Russell was the first to realize the importance of this
phenomenon. Motivated by his observation, in the following years he made
a number of experiments. After the publication of his results [Rus44], a
vigorous scientific debate followed on whether or not such a wave of perma-
nent form could exist. Airy [Air45] argued that even if dissipation, i.e., the
loss of energy, is neglected, dispersion, i.e., the spreading of energy, which is
concentrated in the middle of the solitary wave, in the linear problem will
destroy the solitary wave. It was finally accepted that such waves exist when
Boussinesq [Bou77] and Rayleigh [Ray76] found approximations to such a
wave by deriving the stationary KdV equation (8.15) with some perturba-
tion analysis which takes into account the nonlinear nature of the problem.
In 1895 Korteweg and de Vries [KdV95] derived the time-dependent KdV
equation. A rigorous proof that the full water wave problem possesses such
solitary waves remained open for another few decades and was given by
Friedrichs and Hyers [FH54] in 1954.
What already Boussinesq [Bou77] observed is that in the KdV equation
there is a balance between linear dispersion and nonlinear transport. This
balance creates a wave of permanent form, i.e., u(x, t) = v(x − ct) = v(ξ).
Inserting this ansatz into the KdV equation yields
−cv = −v + 6vv .
Integration w.r.t. ξ yields
−cv + v − 3v 2 = D
8.2. The KdV equation 261
0.5
−0.5
−0.6 −0.4 −0.2 0 0.2
Figure 8.3. The phase portrait for the stationary KdV equation
with D = 0, and the solitary wave.
= (−∂x2 u(x, t) + 3u(x, t)2 )v(x, t) dx,
R
Remark 8.2.2. For the linearized system we have that Tn+2 = (∂xn u)2 since
∂t (∂xn u)2 + ∂x ∂xn+2 u∂xn u − (∂xn+1 u)2 /2 .
8.2. The KdV equation 263
3
In Fourier space we find the solutions u (k, t) = eik t u
(k, 0). Thus, by intro-
ducing polar coordinates u (k, t) = r(k, t)e iφ(k,t) we find that r(k, t) = r(k, 0)
and φ(k, t) = φ(k, 0) + k 3 t mod 2π or, equivalently,
∂t r(k, t) = 0 and ∂t φ(k, t) = k 3 = const.
Thus, for the linearized system we found uncountably many action variables
r(k, ·) and uncountably many angle variables φ(k, ·), and consequently tori
of any dimension.
In the following we describe how a transformation discovered by Miura
[Miu68] and then generalized in [GGKM74] leads very easily to the con-
clusion that there are infinitely many conserved quantities for the KdV equa-
tion. The basic idea is that, given a transformation which maps solutions
of one equation to solutions of a second one, the existence of simple con-
served quantities for the first equation leads, via the transformation, to more
complicated conserved quantities for the second equation.
To a given u = u(x, t) we implicitly define w(x, t) via the formula
(8.17) u(x, t) = w(x, t) + ε∂x w(x, t) + ε2 (w(x, t))2 .
If w is smooth enough and ε is small, then we can invert this relation recur-
sively to obtain w in terms of u via the formula
(8.18) w =u − ε∂x u − ε2 (u2 + ∂x2 u) + ε3 (∂x3 u + 4u∂x2 u)
+ ε4 2u3 + 5(∂x u)2 + 6u∂x2 u + ∂x4 u + O(ε5 ).
Now we compute
∂t u + ∂x3 u − 6u∂x u =(∂t w − 6w∂x w − 6ε2 w2 ∂x w + ∂x3 w)
(8.19) + 2ε2 w(∂t w − 6w∂x w − 6ε2 w2 ∂x w + ∂x3 w)
+ ε∂x (∂t w − 6w∂x w − 6ε2 w2 ∂x w + ∂x3 w).
From this we immediately see that, if w satisfies the modified KdV equation
(8.20) ∂t w − 6w∂x w + 6ε2 w2 ∂x w − ∂x3 w = 0,
then u, defined by (8.17), satisfies the KdV equation. However, one also sees
immediately that the integral of w is a conserved quantity of (8.20) for all
values of ε, i.e., if we define Iε (t) = R w(x, t) dx, then Iε is a constant for
all values of ε. (We assume here that w is defined on the real line, and that
w and its derivatives go to zero as |x| tends to infinity. Similar results hold
for x running over a finite interval with periodic boundary conditions.) But
this immediately implies that, if we use (8.18) to expand Iε in powers of ε,
then the coefficients in this expansion must also be constant in time. Since
these coefficients will be expressed as integrals of u and its derivatives they
will give us (infinitely many) conserved quantities for the KdV equation.
For the first few of these we find:
264 8. Three canonical modulation equations
• K0 = R u(x, t) dx. The conservation of this quantity follows im-
mediately from the form of the KdV equation.
• K1 = R ∂x u(x, t) dx = 0, if we assume that u and its derivatives
tend to zero as |x| tend to infinity. Thus, we gain no new informa-
tion from this quantity and in fact, all the integrals coming from
the odd powers of ε turn out to be “trivial”. Therefore, we ignore
them and focus just on the even powers of ε.
• K2 = R (u2 − ∂x2 u) dx = R u2 dx. That this is a conserved quan-
tity is again easy to see directly from the KdV equation, just by
multiplying the equation by u and integrating w.r.t. x.
• K4 = R ∂x (2u∂x − ∂x (∂x2 u − u2 )) + 2u(∂x2 u − u2 ) + (∂x u)2 dx =
(3u2 + (∂x u)2 ) dx.
In [ZF71] it has been proved that the KdV equation is a completely inte-
grable Hamiltonian system. In particular, there exists a canonical transfor-
mation such that, w.r.t. the new coordinates, the Hamiltonian is a function
only of the action variables and hence the action variables remain constant
in time.
Hence, with L(t) from the example the left-hand side of the KdV equa-
tion and the left-hand side of the Lax pair formulation coincide. Before
we explain how to find operators M (t) such that also the right-hand sides
coincide we explain first consequences of a Lax pair representation.
Lemma 8.2.4. If the self-adjoint operator L=L(t) satisfies ∂t L(t)=[M, L](t),
with M (t) = −M ∗ (t) a skew-adjoint operator, then the eigenvalues of L(t)
are independent of t.
It turns out that for the operator L(t) from Example 8.2.6 an operator
M (t) can be found such that L and M give a Lax pair formulation of the
KdV equation. In order to derive the KdV equation we make the ansatz
M (t)· = −α∂x3 · +U ∂x · +∂x (U ·) + A
with α ∈ R, U = U (x, t) and A = A(x, t). We find
(M (t)L(t) − L(t)M (t))
=(−α∂x3 + ∂x3 U + ∂x2 A + 2(∂x u)U ) − (3α∂x2 u + 4∂x2 U + 2∂x A)∂x
− (−3α∂x u + 4∂x U )∂x2 .
In order to have a multiplication operator the coefficients in front of ∂x and
∂x2 in the second and third line have to vanish identically. From the third
line we obtain U = 3αu/4 and then from the second line that A = A(t) is
independent of x, such that finally
α
M (t)L(t) − L(t)M (t) = (−∂x3 u(x, t) + 6u(x, t)∂x u(x, t)).
4
Choosing α = 4 gives the right-hand side of our KdV equation.
Summary. If u = u(x, t) solves the KdV equation (8.13), then the eigen-
values λ = λ(t) of the eigenvalue problem
−∂x2 ψ(x, t) + u(x, t)ψ(x, t) = λ(t)ψ(x, t)
are independent of time and the associated eigenfunctions evolve according
to
∂t ψ(x, t) = −4∂x3 ψ(x, t)+3u(x, t)∂x ψ(x, t)+3∂x (u(x, t)ψ(x, t))+A(t)ψ(x, t).
268 8. Three canonical modulation equations
For the same L(t), beside the two M (t)s from above there are infinitely
many skew symmetric operators M (t) which lead to a multiplication oper-
ator for M (t)L(t) − L(t)M (t) giving a “hierarchy of KdV equations”. The
next one would be
∂t u = −∂x5 u + 10u∂x3 u + 20∂x u∂x2 u − 30u2 ∂x u.
The Hamiltonian of the nth equation in KdV hierarchy is given by the nth
conserved quantity of the KdV equation obtained via the Miura transfor-
mation, cf. [New85].
The evolutionary problem ∂t ψ(t) = M (t)ψ(t). As the next step on
our way to solve the KdV equation with the Lax pair formulation we have
to solve the evolutionary problem ∂t ψ(t) = M (t)ψ(t) for the eigenfunctions
ψ(t). It is given by
∂t ψ(x, t) = − 4∂x3 ψ(x, t)+3u(x, t)∂x ψ(x, t)+3∂x (u(x, t)ψ(x, t))+A(t)ψ(x, t)
(8.23) =2(u(x, t)+2λ)∂x ψ(x, t) − ∂x u(x, t)ψ(x, t)+A(t)ψ(x, t),
where we used the eigenvalue problem to simplify the system slightly. At
first sight it seems that we need to know the solution u = u(x, t) of the KdV
equation to solve this problem which would make the approach pretty use-
less. However, it turns out that ∂t ψ(t) = M (t)ψ(t) can be solved explicitly
without the knowledge of u. In order to explain why this is the case we
have to look more closely at the eigenfunctions ψ(t) of the self-adjoint linear
operator L(t) defined by
L(t)ψ(x, t) = −∂x2 ψ(x, t) + u(x, t)ψ(x, t),
where u = u(x, t) is a solution of the KdV equation. There are finitely many,
say N , discrete eigenvalues, 0 > λ1 > λ2 > . . . > λN and a continuum [0, ∞)
of spectral values which are denoted by λ(k) = k 2 for k ∈ R. Corresponding
to the eigenvalues there are eigenfunctions ψk ∈ L2 for k = 1, . . . , N and
ψ(k) ∈ L∞ \ L2 for k ∈ R.
Example 8.2.7. We compute the eigenvalues and eigenfunctions of Lψ =
ψ − u(x)ψ in case u(x) = −U0 δ(x), with U0 a positive constant and δ
the Dirac delta-distribution, cf. Example 5.2.2. Integration of −∂x2 ψ −
U0 δ(x)ψ = λψ gives
−∂x ψ|x=− − U0 δ(x)ψ(x) dx = λψ(x) dx,
− −
Since the eigenfunctions satisfy a second order scalar ODE they are
completely determined
by their asymptotic behavior for x → ±∞. For ψk
normalized by |ψk (x)|2 dx = 1 we have
−κ x
e k x → ∞,
ψk ∼ ck
eκk x x → −∞,
where κ2k = −λk and κk > 0. For ψ(k) we have
−ikx
e + b(k)eikx x → ∞,
ψ(k) ∼ −ikx
a(k)e x → −∞.
The question is if it possible to derive some ODEs for the so called scattering
data (λk (t), ck (t)) for k = 1, . . . , N and (λ(k, t), a(k, t), b(k, t)) for k ∈ R
from the PDEs ∂t ψ(t) = M (t)ψ(t) for the eigenfunctions ψ(t). It turns out
that this is possible, and that these ODEs can be solved explicitly without
knowing the evolution of the KdV solution u = u(x, t).
i) We start with the discrete spectrum. Multiplying (8.23) by ψk (x, t)
gives that A(t)ψk (x, t)2 equals
1
∂t (ψk (x, t)2 ) + ∂x (u(x, t)ψk (x, t)2 − 2(∂x ψk (x, t))2 − 4λk ψk (x, t)2 ),
2
270 8. Three canonical modulation equations
by straightforward calculation.
Performing the integration R . . . dx and
2
using the normalization R ψk (x, t) dx = 1 yields
1 d
ψk (x, t) dx = 0 = A(t) ψk (x, t)2 dx = A(t),
2
2 dt R R
i.e., A(t) = 0, such that ψk (t) satisfies
(8.24) ∂t ψk (x, t) = 2(u(x, t) + 2λk )∂x ψk (x, t) − (∂x u(x, t))ψk (x, t).
As already said, ψk (x, t) is fully determined by its asymptotics ck (t)e−κk x
for x → ∞. Since u(x, t) → 0 for x → ∞ we find by multiplying (8.24) with
eκk x in the limit x → ∞ that
ċk (t) = −4λk κk ck (t) = 4κ3k ck (t).
This equation is independent of the evolution of u = u(x, t) and can be
solved explicitly, namely
3
ck (t) = e4κk t ck (0),
where ck (0) has to be computed from the KdV initial condition u = u(x, 0).
The asymptotics
ψk (x, t) ∼ ck (t)e−κk x = ck (0)e4κk t−κk x = ck (0)e−κk (x−4κk t)
3 2
u0 nonlinear
−−−−−−−−− evolution
−−−−−−−in−−KdV
−−→ u(·, t)
−1
↓T ↑T
scattering data scattering data
(λk (0), ck (0))k=1,...,N linear evolution of scattering data (λk (t), ck (t))k=1,...,N
(λ(k, 0), b(k, 0))k∈R −−−−−−−−−−−−−−−−−−−−−−−−−→ (λ(k, t), b(k, t))
k∈R
272 8. Three canonical modulation equations
That this procedure is really useful is explained in the next example where
we compute 2-soliton solutions of the KdV equation.
Example 8.2.9. We consider the situation of two negative eigenvalues λ1 =
−1 and λ2 = −4, such that κ1 = 1 and κ2 = 2. Next we choose the initial
scattering data
√ √
c1 (0) = 6, c2 (0) = 2 3, and b(k, 0) = 0
for all k ∈ R. Such a potential is called reflectionless. We find
√ √
c1 (t) = 6e4t , c2 = 2 3e32t , and b(k, t) = 0
for all k ∈ R and t ∈ R. The function F in the Marchenko or Gelfand-
Levitan equation is then given by
F (x) = 6e8t−x + 12e64t−2x .
In order to solve this equation we make the ansatz
K(x, z) = 1 (x)e−z + 2 (x)e−2z .
With Fj (x) = c2j e−κj x we obtain
0 =1 e−z + 2 e−2z + F1 (x)e−z + F2 (x)e−2z
∞
+ 1 (x)e−y + 2 (x)e−2y )(F1 (z)e−y + F2 (z)e−2y dy.
x
i.e., the solution u(·, t) stays close to a translate of uc for all times.
The argument is similar to the one in Theorem 4.1.6 where the stability
of a fixed point has been shown using the positive definiteness of the second
derivative of the Hamiltonian at the fixed point. See Figure 8.5.
In case of orbital stability the argument is a little bit more complicated.
First of all the solitary waves uc are only constrained minima, i.e., they
minimize H in the invariant subspaces of the functions with the same L2 -
norm. Moreover, in each of these subspaces there is a one-dimensional family
of minima due to the translation invariance of the problem. In the following
we sketch the idea to overcome these difficulties.
274 8. Three canonical modulation equations
o * +
Since uc satisfies the traveling wave equation u c − cuc − 3u2c = 0 for the KdV
equation we see that ∂u M (uc ) = 0 and that uc is a stationary point of M .
Next we restrict u0 to the energy surface of the traveling wave uc , i.e., we
restrict ourselves to the consideration of initial conditions u0 with E(u0 ) =
E(uc ) =: Ec . It can then be shown that uc is a minimum of M in this
surface, and moreover that there exist C1 , C2 > 0 such that
(8.28) 0 ≤ C1 inf u(· + x0 ) − uc (·)2H 1 ≤ M (u) − M (uc ) ≤ C2 u − uc 2H 1
x0 ∈R
for all sufficiently small u − uc H 1 , see [Ben72]. Next we have the positive
definiteness of the second variation
1 2 c
2
∂u M (uc )[v, v] = v + 3u c v 2 + v 2 dx
R 2 2
of M (uc ), cf.(8.27), in the subspace of constant L2 -norm. From (8.28) we
get orbital stability of uc in the energy surface E(u0 ) = Ec . This is called
in the following conditional orbital stability.
8.3. The GL equation 275
From E(uc ) = 8c3/2 /3 it follows that for |E(u0 ) − Ec | < δ there exists a
c∗ with |c − c∗ | ≤ Cδ and E(u0 ) = Ec∗ . Hence, uc − uc∗ H 1 < ε/2 if δ > 0
is sufficiently small. Using this and the conditional orbital stability of uc∗
yields
inf u(·, t)−uc (·+x0 , t)H 1
x0 ∈R
≤ inf u(·, t) − uc∗ (· + x0 , t)H 1 + inf uc∗ (·, t) − uc (· + x0 , t)H 1
x0 ∈R x0 ∈R
≤ε/2 + ε/2
for δ > 0 sufficiently small, i.e., the orbital stability of uc . The orbital
stability of N -solitons of the KdV equation has been established in [MS93].
θ is rather independent of
definition of the uniformly local Sobolev spaces Hul
the particular choice of ρ. Therefore, we may fix ρ once and for all to
ρ(x) = 1/ cosh(x) or ρ(x) = 2/(2 + x2 ),
which both satisfy the conditions on ρ. We let
L2
2 (R) = { u ∈ L2 (R) : u 2 < ∞ },
ul loc L ul
where
u2L2 = sup ρ(y + x)u(x)2 dx.
ul y∈R R
The statement that different weight functions lead to the same uniform space
with equivalent norms is made rigorous with the following lemma.
Lemma 8.3.2. The norm · Lul is equivalent to the norm · ∗ defined by
1/2
y+1/2
u∗ = sup |u(x)|2 dx .
y∈R y−1/2
1/2 ⎛ ⎞1/2
y+1/2
≤ sup |u(x)|2 dx ⎝ sup ρ(y + x)⎠
y∈R y−1/2 y∈Z x∈[−1/2,1/2]
1/2
y+1/2
≤C sup |u(x)| dx
2
y∈R y−1/2
since C = ( y∈Z supx∈[−1/2,1/2] ρ(y + x))1/2 < ∞ due to the assumptions on
ρ. This shows uLul ≤ αu∗ .
278 8. Three canonical modulation equations
In the rest of this section we omit the domain R as all function spaces
considered here are defined over the real line. With the translation operator
Ty : L2
2 → L
ul
22 , (T u)(·) → u( · + y), our final space of uniformly local L2
ul y
functions is given as
L2ul = { u ∈ L2
2 : T u − u 2 → 0 as y → 0 }.
ul y L ul
Proof. a) To show that L2ul is a closed subspace let (un )n∈N be a Cauchy
sequence in L2 with limit u ∈ L2
ul
2 . Then
ul
Ty u−uL2 ≤ Ty u−Ty un L2 +Ty un −un L2 +un −uL2 = s1 +s2 +s3 .
ul ul ul ul
Since un approximates u there exists an n such that s1 = s3 < ε/3. For the
corresponding un ∈ L2ul we find a δ > 0 such that for all y ∈ (−δ, δ) we have
s2 < ε/3.
3
b) Let u ∈ H 1 . With part a) we conclude
ul
Ty u−u2L2 = sup{ ρ(x)(u(x+y+z)−u(x+z))2 dx : z ∈ R } ≤ C|y|u2H 1 .
ul
R ul
3 3
ul ⊂ Lul and hence Hul ⊂ Hul ⊂ Lul for θ ∈ N. To see that Hul
Thus, H 1 2 θ θ 2 1
Example 8.3.4. The embedding H 1 (R) → L2 (R) is not compact, since the
sequence (un )n∈N , with un (x) = u0 (x + n) and u0 (x) = e−x is bounded in
2
Proof. For every ε > 0 we have to show that B admits in Hρ1 a finite covering
by balls of radius less than ε. We decompose every u ∈ B into u = v + w
with v = uχβ and w = u(1 − χβ ), where the smooth cut-off function χβ
vanishes for |x| ≤ β and equals 1 for |x| ≥ β + 1. Then vHρ1 < ε for β
sufficiently large, as ρ decays for |x| → ∞ and u varies in a bounded set.
Moreover, w ∈ H θ ([−β − 1, β + 1]) which can be embedded compactly into
H 1 ([−β − 1, β + 1]). Since for functions with support [−β − 1, β + 1] the
norms in H θ and Hρθ are equivalent, there is a finite covering of this set by
balls BHρ1 (wi , ε) with i = 1, . . . , m < ∞. Thus, ∪i=1,...,m {BHρ1 (wi , 2ε)} is a
finite covering of B, since u − wi Hρ1 ≤ vHρ1 + w − wi Hρ1 .
Multiplier theory in Hul θ -spaces. An important tool for studying trans-
θ is multiplier theory which uses
lational invariant operators on the spaces Hul
q
Fourier transform methods. An operator M : Hul → Hulθ is called multiplier
4 : R → C such that M u = F −1 (M
if there exists a function M 4F u), i.e., if the
associated operator in Fourier space is a multiplication operator. Taking the
Fourier transform of a function u ∈ L2l,u gives a tempered distribution. This
allows easily to define operators via their action in Fourier space. However,
it is rather complicated to estimate the norm of the operators in physical
space. In order to do so we proceed as follows. We will make use of the
fact that Fourier transform is an isomorphism between Hnm and Hm n , cf.
Definition 7.3.30 and Lemma 7.3.31. This is the case for almost all opera-
tors considered so far in this book. We have the following classical result of
multiplier theory.
4(k) ∈ C 0 (R, C).
Lemma 8.3.6. Let q, θ ≥ 0 and wθ−q (k) = (1 + k 2 )(θ−q)/2 M b
4F u) is well defined with the estimate
Then M : H q → H θ , u → F −1 (M
M uH θ ≤ C(q, θ)wθ−q C 0 (R,C) uH q ,
b
4.
where C(q, θ) does not depend on M
280 8. Three canonical modulation equations
4.
where C(q, θ) does not depend on M
Proof.
We choose χ ∈ C0∞ such that its support is contained in [−1, 1] and
j∈Z χ(x + j) ≡ 1. The operator Mul then can be defined as follows. For
q
u ∈ Hul we set uj (x) = u(x − j)χ(x). Since uj ∈ H2q we find vj = M uj ∈ H2θ
according to
M uj H θ ≤ CM 4uj H 2 ≤ Cwθ−q C 2
uj Hq2 ≤ uj H2q .
2 θ b
θ . But,
Now let M u = j∈Z Tj vj . Clearly this sum does not converge in Hul
since Tj vj is concentrated around x = j and decays like 1/(1 + (x − j)2 ) it is
θ with norm
easy to see that j∈Z Tj vj converges locally to a function in Hul
≤ C(q, θ)wθ−q C 2 (R,C) uH q .
b ul
T (t)uH θ = T(t)
uL2 ≤ T(t)C 0
uL2 ≤
uL2 = uH θ .
θ b θ θ
uL1−r
+ v L2 w
L1−r +
uL1−r
v L1−r w
L2 ).
θ+2r θ+r
uL1−r ≤ C
Since uL2 for every δ > 0 and since Fourier transform is
1/2−r+δ
an isomorphism between L2s and H s we can choose θ = −1 − 2δ, r = 1/2 + δ
to establish the local Lipschitz-continuity of the nonlinear term N (u) =
u − (1 + iβ)|u|2 u from L2 into H −1−2δ for every δ > 0. Since the semigroup
T (t) is smoothing with a singularity t−1/2−δ from H −1−2δ to L2 the right-
hand side of the variation of constant formula is a contraction in the space
C([0, T0 ], L2 ), too, if T0 > 0 is sufficiently small. Hence, Theorem 8.3.8 is
already true for every θ ≥ 0.
≤9uH θ vH θ
ul ul
for θ > 1/2 and as a consequence uvH θ ≤ 9uH θ vH θ . Similarly, the
ul ul ul
embedding follows.
Then we have the following local existence and uniqueness result.
Theorem 8.3.12. Let θ > 1/2 and u0 ∈ Hul θ . Then there exists a T =
0
T0 (u0 H θ ) > 0 and unique mild solution u ∈ C([0, T0 ], Hul
θ ) of the GL
ul
equation with u|t=0 = u0 .
to get a bound for the semigroup T (t) = F T(t)F −1 in Hulθ we use Lemma
k∈R
such that T(t)C 2 is bounded for finite t. Hence, the right-hand side of the
b
θ ) if
variation of constant formula is a contraction in the space C([0, T0 ], Hul
T0 > 0 is sufficiently small. Therefore, we are done.
The local existence and uniqueness theorem can be improved from θ >
1/2 to θ ≥ 0. In order to do so we combine Remark 8.3.9 and Lemma 8.3.11.
Using the inequality of Remark 8.3.9 yields
1
1 1
uvwH θ [k,k+1] = ( uk+j )( vk+j )( wk+j )
j=−1 j=−1 j=−1 θ
H [k,k+1]
1
1
1
≤ uk+j1 vk+j2 wk+j3 H θ [k,k+1]
j1 =−1 j2 =−1 j3 =−1
1
1
1
≤ uk+j1 vk+j2 wk+j3 H θ
j1 =−1 j2 =−1 j3 =−1
1
1
1
≤C uk+j1 L2 vk+j2 L2 wk+j3 L2
j1 =−1 j2 =−1 j3 =−1
≤ 9CuL2 vL2 wL2 ,
ul ul ul
∂x r
(8.31) ∂t φ = ∂x2 φ + 2 + ∂x φ, ∂t r = ∂x2 r + r − r3 − r(∂x φ)2 .
r
Since supx∈R |u(x, t)| = supx∈R r(x, t) = R(t) it is sufficient to get a bound
for r. Since (∂x φ)2 ≥ 0 we obtain
∂t r ≤ ∂x2 r + r − r3 .
The real case in L2ul . Let α = β = 0 and assume that u0 ∈ L2ul . In order
to find the bound in L2ul we consider weighted energy estimates for Ey (t) =
2 −1 and ρ (x) = ρ(x + y). We
R u(x, t)u(x, t)ρy (x) dx with ρ(x) = 2(2 + x ) y
284 8. Three canonical modulation equations
R R
Using the variation of constant formula and the smoothing properties of the
semigroup we find that
Remark 8.3.16. First a priori estimates for the size of the solutions have
been established in [BCD+ 90] leading to the terminology of soft and hard
turbulence regimes in the (α, β)-plane. Later on the estimates have been
improved in [Mie98] where also estimates for higher space dimensions can
be found.
8.3. The GL equation 285
Proof. Let (un )n∈N ∈ B be a sequence with limit u in Zρ . Since un < r
and since Ty is continuous in Zρ we obtain for fixed y
Ty uρ ≤ Ty un ρ + Ty un − Ty uρ ≤ r + εn
with εn → 0 for n → ∞. Thus, u ≤ r which is the desired result.
0
Example 8.3.20. Consider Z = Cb,unif and let Zρ be equipped with the
norm
uρ = sup |u(x)/(1 + x2 )|.
x∈R
8.3. The GL equation 287
These two norms allow us to define two different distances between sets,
distZρ and distZu .
Definition 8.3.21. We define
distZρ (b, A) = inf a∈A b − aρ ,
distZu (b, A) = inf a∈A b − a = inf a∈A supy∈R Ty b − Ty aρ .
For both distances we let dist(B, A) = supb∈B dist(b, A) for B ⊂ Z.
dist∗Zρ (St (B), A) = sup sup inf Ty St (b) − Ty aρ → 0 for t → ∞.
b∈B y∈R a∈A
We further remark that if Zρ1 and Zρ2 define equivalent norms in Zu via
(A1) then for the attractors constructed via Zρi we have A1 = A2 .
Due to the smoothing and global existence results for the GL equation
(8.29) we then have
Theorem 8.3.25. For every θ ≥ 0 the GL equation (8.29) has a global
θ , H θ )-attractor A which satisfies attractivity in dist∗ , is translational
(Hul ρ G Hθ ρ
invariant, and invariant under the rotations Rφ : A → eiφ A.
−2 0
−2
−1
−3 −3 −2
−2 −1 0 1 2
−2 −1 0 1 2 −2 −1 0 1 2
t
for a constant φ∗ ∈ R, such that φ ∼ t−1/2 , ∂t φ ∼ t−3/2 , and ∂x2 φ ∼ t−3/2
for t → ∞. It turns out that s is slaved by φ and that in lowest order s =
−(∂x φ)2 /2 ∼ t−2 for t → ∞. Therefore, (8.42) and (8.43) asymptotically
behave like
(8.44) ∂t φ = ∂x2 φ + O(t−7/2 ),
(8.45) O(t−1 ) = −2s − (∂x φ)2 + O(t−1 )
for t → ∞. Hence, all other terms vanish faster than those that we claimed
to describe the asymptotic behavior. This argument can be made rigorous
292 8. Three canonical modulation equations
Further Reading. Besides its importance for applications, the NLS equa-
tion also plays a big role inside mathematics. Bourgain (in 1994) and Tao (in
2006) received Fields medals partly for work about the local and global exis-
tence and uniqueness of solutions in spaces of functions with low regularity,
see, e.g., [Bou99, Tao06]. There are numerous books and review papers
about the NLS equation, e.g. [AS81, DJ89, SS99b, BK00, Fib15]. A
major result is that the NLS equation is a completely integrable Hamil-
tonian system. There exist so called inverse scattering schemes, namely
the Ablowitz-Kaup-Newell-Segur (AKNS) scheme and the Zakharov-Shabat
(ZS) scheme which allow to solve the NLS equation explicitly, cf. [AS81,
DJ89]. The so called Birkhoff normal form for the NLS equation in case of
periodic boundary conditions is discussed in [GK14], which also contains a
modern account of the Hamiltonian formalism, of the Poisson bracket, and
of the Lax pair formalism for the NLS equation in terms of the ZS operator.
Besides the motivation by the water wave problem, cf. §8.2.1, the KdV
equation can be derived as an approximation to various other physical sys-
tems, see also [Cri95] and Chapter 12. The question occurs which phenom-
ena of the KdV equation are robust under perturbations. Viewing the KdV
equation as a completely integrable Hamiltonian system, this is analogous
to the questions studied by the Kolmogorov-Arnold-Moser (KAM) theory,
cf. §4.4, and has led to a development of KAM-like results for a number of
different PDEs like the KdV equation. If one considers the KdV equation
with periodic boundary conditions, temporally periodic or quasi-periodic
solutions will persist under small perturbations, cf. [KP03]. The situation
is more complicated and less well-understood for the equation on the whole
line. The local and global existence of low regularity solutions of the KdV
equation and related equations is a very active field of research. As already
said, see, e.g., [Tao06, §4], [LP09, Ch. 7], or [Koc15, Ch. 6] for modern
accounts.
The GL equation as a universal modulation equation occurs at the
end of the 1960s in a number of papers, cf. [NW69, Seg69, dES71],
cf. Chapter 10. There exist many further explicit solutions of (8.29), cf.
[AK02, Mie02]. A famous example are the so called Bekki-Nozaki holes,
see [BN85, vSH92]. There also exist stability results for such rather com-
plicated solutions of the GL equation, cf. [BNSZ14]. Definition 8.3.18 about
8.3. The GL equation 293
8.11. Prove directly that the energy R
u2 (x, t) dx is a conserved quantity for the
KdV equation.
√
8.12. Find α ∈ R such that u(x, t) = αc sech2 ( c(x − 4ct)) , with arbitrary c > 0
is an exact solution of KdV.
8.13. Let L(t)=L(t)T ∈Rd×d satisfy L(0) = U (t)T L(t)U (t) with U (t)U (t)T = I.
Show that L = L(t) satisfies
∂t L(t)=M (t)L(t)−L(t)M (t)
with
M (t)= − (∂t U (t))U (t)T =U (t)∂t U (t)T = − M (t)T ,
and that as a consequence the eigenvalues of L = L(t) are independent of t.
3 + 4 cosh(2ξ + 24t) + cosh(4ξ)
8.14. Consider the solution u(x, t) = −12 of the
(3 cosh(ξ − 12t) + cosh(3ξ + 12t))2
KdV equation, where ξ = x − 16t. Prove that for t → ±∞ this solution
separates
1 1
into two single waves, i.e., u(x, t) ∼ −8 sech 2ξ ∓ log 3 −2 sech η ± log 3
2 2
2 2
with η = x − 4t, where ∼ means asymptotically equal.
8.15. Show that xu + 3t2 u2 dx is conserved for the KdV equation.
8.16. Show that the Kadomtsev-Petviashvili (KP) equation ∂x (∂t u−6u∂x u+∂x3 u)+
3∂y2 u = 0 follows from the Lax pair L = −∂x2 + ∂y + u and M = −4∂x3 + 6u∂x +
x
3ux + 3 ∂y u dξ.
8.17. Let u = ur + iui and rewrite the GL equation (8.29) as a real system for
(ur , ui ).
8.18. Show that the GL equation ∂t u = ν2 ∂x2 u + ν0 u − ν3 |u|2 u, with νj ∈ C,
Reν0 , Reν2 , Reν3 > 0, and x ∈ R, can be rescaled to the standard form
∂t u = (1 + iα)∂x2 u + u − (1 + iβ)|u|2 u, α, β ∈ R.
8.19. a) Show local existence of solutions of the GL equation (8.29) for u0 ∈ H 1 (R).
b) Derive a priori estimates to show global existence. Do solutions stay bounded
in H 1 (R)?
8.20. Show that the space of uniformly continuous functions can equivalently be
0
characterized by Cb,unif (R, R) = { u ∈ Cb0 (R, R) : Ty u − uCb0 → 0 as y → 0 }.
8.21. Use the variation of constant formula and smoothing properties of the GL
semigroup to establish a bound lim supt→∞ u(t)Hul
1 = O(|β|) for |β| → ∞ from
√
lim supt→∞ u(t)L2ul ≤ 4π 2 for the solutions of the GL equation.
8.22. Derive the expansion (8.37).
8.23. (Benjamin–Feir instability for the NLS equation) Discuss the sta-
bility of the solution a(t) = a0 e−ia0 t of the NLS i∂t u − ∂x2 u − |u|2 u = 0. For
2
the linearization around a, make an ansatz u(x, t) = a(t)(1 + b(x, t)) with b(x, t) =
b1 ei(ωt+kx) +b2 ei(ωt−kx) , and derive an algebraic system for the coefficients b1,2 ∈ C.
Show that this system possesses non-trivial solutions if ω 2 = 2k2 (a20 − k2 ), and
from this derive that a is unstable w.r.t. perturbations with wave numbers k with
k2 < a20 .
Chapter 9
Reaction-Diffusion
systems
In §7.1 we considered with the KPP and the Allen-Cahn equation the sim-
plest examples of scalar nonlinear reaction-diffusion equations. In general,
chemical reactions involve a number of different species, and so we consider
in this section reaction-diffusion systems of the form
(9.1) ∂t u = DΔu + f (u),
with t ≥ 0, x ∈ Rn , u(x, t) ∈ RN , where f : RN → RN a smooth func-
tion, and D ∈ RN ×N is a positive (semi-)definite diffusion matrix. Of-
ten D is diagonal with strictly positive diagonal elements dj > 0, i.e.,
D = diag(d1 , . . . , dN ).
Famous examples of two-component reaction-diffusion systems, where
a, b, d, . . . denote parameters, are:
The Schnakenberg model [Sch79]
(9.2) ∂t u = Δu − u + u2 v, ∂t v = dΔv + b − u2 v.
The Brusselator [PL68]
(9.3) ∂t u = Δu + a − (b + 1)u + u2 v, ∂t v = dΔv + bu − u2 v.
The Lengyel-Epstein model [LE91]
4uv uv
(9.4) ∂t u = Δu + a − u − , ∂t v = dΔv + b(u − ).
1 + u2 1 + u2
The Gray-Scott model [GS83]
(9.5) ∂t u = Δu − uv 2 + f (1 − u), ∂t v = dΔv + uv 2 − (f + k)v.
295
296 9. Reaction-Diffusion systems
While models (9.2)-(9.6) essentially are all from chemistry, (9.7) models
the transmission of nerve impulses in some giant squid axon. Moreover,
the Kolmogorov models from ecology, i.e., the 2-species interactions from
Example 2.6.5, reappear as reaction-diffusion systems when combined with
diffusion of one or both of the species.
Typical questions about reaction-diffusion systems are again the local
and global existence of solutions in a suitable phase space, and the existence
and stability of special solutions such as pulses and fronts. Concerning
global existence, the crucial difference compared to scalar equations is that
the maximum principle is no longer valid for systems. However, it can often
be replaced by the method of invariant regions. Similarly, the existence of
special solutions becomes more complicated due to the fact that the traveling
wave system in general can no longer be written as a two-dimensional first
order system, such that phase plane methods are no longer available but
have to be replaced by more sophisticated methods, which often use so
called trapping regions.
In §9.1 we start with modeling aspects and the local existence and
uniqueness theory for reaction-diffusion systems, and discuss quadratic au-
tocatalysis as an example of a two-dimensional system of the form (9.1). It
turns out that when d1 = d2 such systems can be rescaled to the KPP equa-
tion from §7.1. We explain the concept of invariant regions which allows
to prove global existence and uniqueness of solutions for many reaction-
diffusion systems.
The FitzHugh-Nagumo system (9.7) and the Gray-Scott system (9.5)
are used as examples in §9.2 to discuss, on a heuristic level, some typical
phenomena of travelling pulses and pattern formation in reaction-diffusion
systems. However, our main objective is the Turing instability of spatially
homogeneous states, also called diffusion driven instability [Tur52, Mur89],
which we consider in §9.3. This is a famous explanation for morphogenesis,
which refers to pattern formation during growth in biological system. The
bifurcating pattern very often can be described by the GL approximation,
which will be studied in Part IV.
9.1. Modeling, and existence and uniqueness 297
Proof. The proof is very similar to the proof of Theorem 7.1.7. First we fix
a C1 > 0 and define the semigroup etDΔ by
2 2
etDΔ u = diag(etd1 ∂x u1 , . . . , etdN ∂x ud ),
298 9. Reaction-Diffusion systems
− x−y
where etdj ∂x uj = √
2 1 2
e 4dj t uj (y) dy if dj > 0, cf. (7.6), and etdj ∂x uj =
4πdj t R
uj if dj = 0. Then etDΔ is a C0 -semigroup in X and etDΔ uX ≤ uX .
Next we use the variation of constant formula and show that
t0
F (u) = etDΔ u0 + e(t−s)DΔ f (u(s)) ds
0
maps
M = C 0 ([0, t0 ], {u ∈ X : u − etDΔ u0 X ≤ C1 }),
equipped with the norm
uM = sup u(t)X ,
0≤t≤t0
into itself and is a contraction in M if t0 > 0 is sufficiently small.
Remark 9.1.2. If dj > 0 for j = 1, . . . , N , then the solution u(·, t) becomes
analytic for t > 0 and hence a classical solution. If one or more of the dj
are zero, then in general the solutions only stay as smooth as the initial
conditions.
left, one propagating to the right. Numerical simulations of (9.10) show that
this is indeed the case, see Figure 9.1. Note that the critical front speed only
depends on the diffusion coefficient of the autocatalyst B, which moreover
determines the steepness of the front. We remark that the shape of the front
vhet (ξ) can be obtained by considering asymptotic limits, in particular the
limit 0 < d 1, but for this and more analysis we refer to the literature,
e.g., [BK00].
Figure 9.1. Fronts (b-component) for (9.10) with d=1 (left) and
d=5 (right) for the initial condition (u0 , v0 )(x) = (1, 0.2 sech(x/2)).
The idea of the proof is as for the maximum principle in §7.1.4. For a
d
solution u coming to the boundary for (x0 , t0 ) we have dt gi (u(x0 , t0 )) < 0
such that the solution goes back to the interior of Σ.
Remark 9.1.6. Since D is diagonal, any left eigenvector ∂u g of D is also
a right eigenvector. If D = diag(d1 , . . . , dN ) with di = dj for i = j,
then the eigenvectors are obviously of the form e1 := (1, 0, . . . , 0), e2 :=
(0, 1, 0, . . . , 0), . . ., eN := (0, 0, . . . , 0, 1), which shows that each gi actually is
a function of only one variable. Essentially, this means that Σ is a rectangle
and that we consider the system in a “decoupled way”. If, e.g., di = dj for
exactly two i = j, then we have a two-dimensional eigenspace span{ei , ej }
and hence also more freedom for gi and gj .
Remark 9.1.8. Often, only rectangles Σ of the form (9.14) can be con-
structed such that f does not point outward on ∂Σ, i.e., f may vanish on
∂Σ or may be tangential to ∂Σ. Then, instead of (3) in Theorem 9.1.5 we
only have ∂u gi (u0 ) · f (u0 ) ≤ 0. However, under mild technical assumptions
302 9. Reaction-Diffusion systems
on f , which are fulfilled for smooth f , the assertion of Theorem 9.1.5 remains
true, i.e., Σ is positively invariant, see [Smo94, Thm 14.11].
a 1 u
f(u)−v=0
so called excitable case with exactly one stable fixed point. In the so called
oscillatory case the unique fixed point is unstable, while in the so called
bistable case there are three fixed points, two stable and one unstable (see
below).
In any case, there are large contracting sets, as illustrated by the rec-
tangle in Figure 9.2. Thus, using Theorem 9.1.1, and Corollary 9.1.7 we
conclude that for arbitrary (u0 , v0 ) ∈ X = [Cb,unif
0 (R)]2 we have a global
solution to (9.16).
From the modeling point of view, in suitable parameter regimes, (9.16)
should possess pulse solutions, i.e., solutions of the form
u, v)(x − ct)
(u, v)(x, t) = (
u, v)∗ as ξ → ±∞, where (
u, v)(ξ) → (
with ( u, v)∗ is a fixed point of (9.17).
These pulses are found in the regime b, c and d small. Then (9.17) is
often rescaled to the form
(9.18) εu̇ = f (u) − v, v̇ = u − γv,
where 0 < ε 1 is a small parameter. Since u̇ = 1ε (f (u) − v) in (9.18) is
large except if f (u) − v = 0, the u-dynamics is called fast in contrast to the
slow dynamics in v.
Excitable dynamics. If 0 < a < 1/2 and γ < 4/(a − 1)2 , then (u, v) =
(0, 0) is the only fixed point of (9.18). It is asymptotically stable, with
eigenvalues λ1 = −a/ε and λ2 = −1. However, as is evident from the phase
portrait in Figure 9.3, a finite perturbation with, e.g., u > 0 and v < 0 may
lead to a large excursion with u > 1 before the orbit returns to (0, 0) from
the left. This is called excitable behavior. In particular, this happens for
ε > 0 sufficiently small.
To understand this behavior we follow [Kee88, §12]. We first set ε = 0,
which formally yields
(9.19) v = f (u), v̇ = u − γv.
This is an implicit ODE for v. The algebraic equation v = f (u) has three
branches of solutions, namely
u = g− (v), u = g0 (v) and u = g+ (v),
see Figure 9.3. The lower/upper branches g−√(v) and g+ (v) exist only for
u < u∗− and u > u∗+ , where u∗± = 12 (a + 1 ± a2 − a + 1), and the middle
branch g0 (v) exists for u∗− ≤ u ≤ u∗+ . Equivalently, g− (v) and g+ (v) exist
for v > v− ∗ := f (u∗ ) and v < v ∗ := f (u∗ ), respectively. Thus, depending
− + +
on the branch, (9.19) has three different meanings, namely v̇ = g± (v) − v or
v̇ = g0 (v)−v. On the middle and upper branches, v increases monotonically,
while g− (v) − v changes sign at v = 0.
304 9. Reaction-Diffusion systems
0.2
f1=0
0.15 f =0
2
v bu−cv=0
g−(v) 0.1
0.05
g0(v) 0
a 1 u
f(u)−v=0 −0.05
−0.1
g+(v) 0 0.5 1
(9.20) U̇ = f (U ) − V, V̇ = ε(U − V ).
Given (u0 , v0 ) we can now (heuristically) piece together the solution of (9.18)
using the above two different scaling limits. Assume that u > 0 and v < v− ∗.
0.05 0.1
0.05 0
0
0
−0.05
−0.05 −0.5
−0.1 −0.1 0 1 2 3 4 5
0 0.5 1 −0.5 0 0.5 1
Figure 9.4. The bistable (a, γ, ε) = (0.25, 8, 0.01) and the oscilla-
tory (a, γ, ε) = (−0.25, 1, 0.01) case: phase portrait and relaxation
oscillations.
Travelling pulses for the FHN system. In the excitable regime we may
expect pulse solutions for the FHN system. The heuristic reason is as fol-
lows. Assume that we start with a (suitable, sufficiently large) perturbation
of (u, v) ≡ (0, 0), localized near some x0 . Then, locally, the ODE dynamics
wants to run the excitation. Meanwhile, diffusion pulls the solution beyond
the threshold for excitation also in a neighborhood of the primary perturba-
tion, and the process repeats. Thus, we may expect one or more (depending
on the details of the IC) excitation pulses to emerge, see Figure 9.5 for two
examples. Moreover, the pulses are asymptotically stable in the numerics,
with a large basin of attraction as we do not need any fine tuning of the
initial data.
306 9. Reaction-Diffusion systems
0.8
0.6
1200
1500
0.4
1
1000 0.5 800
1
0.2 0
0.5
0 500 400
0
-300
-200
-100
0
-300 -200 -100 0 -0.2 100
0 100 200 200 0
300 -300 -200 -100 0 100 200 300 300
The analysis for the existence and in particular for the stability of the
FHN pulses is rather difficult, and we refer to the literature, see, e.g.,
[Jon84, Kue15].
9.2.2. The Gray Scott model. The Gray Scott model [GS83] is a famous
model for cubic autocatalysis of two species according to A + 2B → 3B, rate
k1 ab2 , where A is fed into the system at rate kf and B decays with rate k2 b,
and where the diffusion coefficients are DA and DB . After suitable non-
dimensionalization the system reads
∂t u = ∂x2 u − uv 2 + f (1 − u),
(9.22)
∂t v = d∂x2 v + uv 2 − (f + k)v,
with typically 0 < d < 1 in applications.
The nullclines of the ODE for (9.22) are given by
g1 = 0 : u = f /(f + v 2 ), g2 = 0 : u = (f + k)/v or v = 0.
We always have the stable fixed point P = (1, 0). For f > 4(f + k)2 there
occurs a saddle-node bifurcation with saddle Q = ((f + k)/v− , v− ) and
unstable node R = ((f + k)/v+ , v+ ) with
f f2
v± = ± − f.
2(f + k) 4(f + k)2
The radicand is positive for f > 4(f + k)2 , or equivalently k < −f +
f /4 =: fsn (f ). However, for k < kHopf (f ), R regains stability by a Hopf
9.3. The Turing instability 307
v 0.5
0.5 f1=0
f1=0
f =0
f =0 0.4 2
0.4 2
f(u,v)=0
0.3 0.3
R
f<0 0.2 0.2
7500
3000 1500
0.5 0.4
0.25 5000
2000 0.2 1000
0 0.5
0
0.25
2500
1000 500 0
-200
-100 -300
0 -200
0 -100
100 0 0
100
200 0 300 0 200
assume that at some fixed μ the system has a stable spatially homogeneous
equilibrium u∗ , i.e., u∗ is a stable fixed point of the ODE u̇ = f (u).
Intuitively one would expect that diffusion cannot destabilize a homoge-
nous state u∗ . However, Turing [Tur52] recognized that this intuition is
wrong if the diffusion matrix D is not a scalar multiple of the identity, and
it turns out that u0 may be unstable w.r.t. harmonic waves with some wave-
number kc = 0. An attempt for an intuitive explanation using fire as the
activator with fast diffusion and “sweating grasshoppers” as inhibitor with
slow diffusion is made in [Mur89].
The general situation is as follows. The linearization
∂t v = DΔv + fμ (u∗ )v
of (9.1) around u∗ has solutions of the form
v(x, t) = eikx+λ(k)t v(k) + c.c.,
where the eigenvalue λ(k) ∈ C and the vector v(k) ∈ CN are determined
from the N × N eigenvalue problem
(fμ (u0 ) − k 2 D)
v (k) = λ(k)
v(k),
i.e., there exist N curves k → λj (k, μ) with j = 1, . . . , N , which we order
such that Reλj (k, μ) ≥ Reλj+1 (k, μ).
The homogeneous state u∗ is unstable if there exists a k0 ∈ R with
Reλ1 (k0 ) > 0. The instability appears as some parameter μ of the system
is varied. For instance Reλ1 (k) < 0 for all k ∈ R if μ < μ0 , while at μ = μ0
we have Reλ1 (k0 ) = 0 for some k0 > 0 and Reλ1 (k) < 0 for k = ±k0 .
Finally, for μ > μ0 , Reλ1 (k) > 0 for all k in two unstable bands, i.e.,
±k ∈ (k− (ε), k+ (ε)), with ε2 = μ − μc .
9.3. The Turing instability 309
Im λ 1
Im λ 1
k=k c
k=k c
Re λ 1 Re λ 1
Re λ 1
Re λ1,2
Im λ 2
Figure 9.8. The four different instability scenarios
with
2
u ∂x u+u + b2 v 1
L(∂x ) = , u, v) = (2u0 u
g( v + u
v0 + u
2
v)
2
.
v d∂x2 v − 2
u − b2 v −1
Plugging the ansatz ( u, v)(x, t) = eikx+λ(k)t ϕ with ϕ = ϕ(k) ∈ C2 into the
u, v) = L(∂x )(
linearization ∂t ( u, v) yields the eigenvalue problem
2
−k + 1 b2
(9.25) ϕ = λϕ,
−2 −dk 2 − b2
with wave number k ∈ R as parameter. For notational convenience in the
following we drop the dependence on (b, d). The two curves of eigenvalues
λ1,2 (k) are determined by
(9.26) λ2 − λ[−(1 + d)k 2 + 1 − b2 ] + [dk 4 + (b2 − d)k 2 + b2 ] = 0.
The stability of the stationary point for the reaction system is determined
by λ1,2 (0). We obtain
λ1,2 (0) = (1 − b2 )/2 ± (1 − b2 )2 /4 − b2
such that (b, 1/b) is unstable for b ∈ (0, 1). We have Im(λ1,2 (0, b)) = 0 for
65 5 7
√ √
b∈ (3 − 5)/2, (3 + 5)/2 .
1
4 Reλ1
Imλ1 (k)
Imλ1
0
2
Reλ
2
Imλ2 (k)-1 0
-2 Reλ1,2 (k) −2
-3 −4
-2 0 2 −2 −1 0 1 2
Figure 9.9. The curves of eigenvalues in the Hopf and in the Turing
case. Here (b, d) = (0.7, 1) and (b, d) = (3, 60), respectively.
Reλ1
0.2 Reλ2
0
-0.2
-0.4
-0.6
-0.8
-1
-1.5 -1 -0.5 0 0.5 1 1.5
satisfy
λ2 − (fu + gv )λ + (fu gv − fv gu ) = 0.
Hence, the steady state (u∗ , v ∗ ) is stable for the ODE (9.28) if
(9.29) tr A = fu + gv < 0 and det A = fu gv − fv gu > 0.
The eigenvalues λ of
∂x2 0
+A
0 d∂x2
satisfy
λ2 + λ(k 2 (1 + d) − (fu + gv )) + h(k 2 ) = 0,
with h(k 2 ) = dk 4 −(dfu +gv )k 2 +det A. Since already fu +gv < 0, a necessary
condition for Turing instability of the steady state (u∗ , v ∗ ) is h(k 2 ) < 0 for
some k, which requires dfu + gv > 0, hence, by (9.29) d = 1 and fu gv < 0.
Ultimately, we need hmin < 0, where, by calculus,
h(k 2 ) = hmin = det A − (dfu + gv )2 /(4d) at k 2 = km
2
= (dfu + gv )/(2d).
At the bifurcation point the condition hmin = 0 defines via |A| = (dfu +
gv )2 /4d a critical diffusion ratio as a root of d2c fu2 +2(2fv gu −fu gv )dc +gv2 = 0.
The critical wave number kc is given by
(9.30) kc2 = (dfu + gv )/(2d) = (fu gv − fv gu )/dc .
312 9. Reaction-Diffusion systems
Exercises
9.1. Show that any quadratic autocatalytic system (i.e., arbitrary k, DA , DB > 0)
9.2. Show that the vector field f defined by (9.13) points inward on ∂R. Use this to
prove that there exists a heteroclinic orbit vhet (ξ) connecting (0, 1, 0) and (1, 0, 0).
9.3. Consider the reaction–diffusion–systems
∂t u = ∂x2 u + uM (u, v), ∂t v = d∂x2 v + vN (u, v),
associated to the Kolmogorov form of the equations for 2–species interaction in
mathematical ecology, see Example 2.6.5. For each type, (PP), (C) and (S), con-
struct (weakly) invariant regions, either abstractly, or for the concrete examples
given in Example 2.6.5.
9.4. Show that (9.18) is equivalent to (9.17) with ε = b, γ = c/ε and τ = εt.
9.5. Let the linearization of a system (9.27) around (u, v)∗ be given by
1 + ∂x2 4
wt = Aw with A = .
−1 −3 + d∂x2
Use Lemma 9.3.2 to find dc such that (u, v)∗ is Turing unstable for d>dc , and
calulate kc .
9.6. Sketch the Turing space for the unique fixed point of (simplified) Gierer–
Meinhard system
∂t u = ∂x2 u + a − u + u2 /v, ∂t v = d∂x2 v + u2 − v, a, d > 0.
9.7. In general, reaction diffusion systems may include so called cross-diffusion,
where the diffusion of one species directly influences some other species. For sim-
plicity we restrict to
u 1 0
(9.31) ∂t U = DΔU + F (U ), where U = , D= .
v d3 d4
Thus, if d3 > 0, then “cross–diffusion” in ∂t U = DΔU yields a decrease/increase
of v at the maxima/minima of u. Assume that (9.31) has a homogeneous steady
state U∗ . Show that U∗ is Turing unstable if
fu + gv < 0, fu gv − fv gu > 0, d4 fu − d3 fv + gv > 0,
(d4 fu − d3 fv + gv ) − 4d4 (fu gv − fv gu ) > 0,
2
d4 fu − d3 fv + gv
with critical wave number kc given by kc2 = .
2d4
Chapter 10
Dynamics of pattern
and the GL equation
This is the first chapter of Part IV of this book. Part IV is about modula-
tion theory which is applied to physically realistic systems over unbounded
domains. Like for systems over bounded domains, which very often can be
reduced to finite-dimensional ODEs, also for problems over unbounded do-
mains often only a subset of the uncountable many degrees of freedom plays
a role. These still uncountable many degrees of freedom are described by
simpler PDEs, which in this book are called modulation equations. They
can be derived via perturbation theory. In particular, the Korteveg-de Vries
(KdV), the Nonlinear Schrödinger (NLS), and the Ginzburg-Landau (GL)
equations from Chapter 8, play an important role as modulation equations,
and one unifying theme of the remaining chapters of this book is what can
(and what can not) be deduced about the original full system from these
reduced PDEs.
Part IV consists of five chapters. The first three chapters are about the
derivation and justification of such approximations. Here, in Chapter 10 we
discuss the GL approximation, in Chapter 11 the NLS approximation, and
in Chapter 12 the KdV approximation. The last two chapters handle aspects
of the existence and stability theory of special solutions of nonlinear PDEs
on unbounded domains which are related to modulation theory. We start in
Chapter 13 with the construction of solutions via the spatial dynamics and
invariant manifold approach. The final Chapter 14 is about the stability of
solutions via the diffusive or dispersive stability approach.
315
316 10. Dynamics of pattern and the GL equation
10.1. Introduction
In §9.3 we already discussed, on a linear level, Alan Turing’s remarkable
observation [Tur52] that diffusion can destabilize ODE fixed points in the
associated PDE, i.e., destabilize spatially homogeneous fixed points u∗ of
two-component reaction-diffusion systems of the form
∂t u = DΔu + f (u),
with u(x, t) ∈ R2 , x ∈ Rd , a reaction f : R2 → R2 , Δ = dj=1 ∂x2j , and
D ∈ R2×2 a diagonal matrix with entries d1 , d2 > 0. Until the work of Turing
it was believed that whatever the reaction is, diffusion will homogenize the
reactants in space. Turing found that this not true and that even simple
reaction-diffusion systems can lead to a spatial structure of the solutions
similar to the pattern on animals such as tigers, zebras, leopards and many
others.
As already explained in §9.3 the idea behind this observation is surpris-
ingly simple. The linearization
(10.1) ∂t v = DΔv + f (u∗ )v
around u∗ is solved in case x ∈ R by
(10.2) v(x, t) = eikx+λt v,
with k ∈ R, λ ∈ C, and v ∈ C2 , where λ and v are determined by the
eigenvalue problem
v + f (u∗ )
−|k|2 D v = λ
v.
10.1. Introduction 317
The first line of the residual vanishes by choosing the envelope A of the
bifurcating spatially periodic pattern eix to satisfy the GL equation
(10.7) 2
∂T A = 4∂X A + A − 3A|A|3 .
320 10. Dynamics of pattern and the GL equation
x
x
1/ε
In the phase portrait for (10.10) we also find periodic solutions (A, B) =
(A, B)per (X) surrounding the origin (A, B) = (0, 0). For the SH equation
we find formally the equilibria
u(x) = 2εAper (εx) cos x + h.o.t..
322 10. Dynamics of pattern and the GL equation
The envelope A = Aper (X) is spatially periodic with some period of order
O(1/ε) w.r.t. the fast spatial variable x and modulates the underlying pat-
tern cos x which has some period of order O(1). These solutions are sketched
in Figure 10.3.
A 1/ε
x
u
1/ε
Figure 10.4. The front solution in the GL equation and the modulating
front for the SH equation.
40
0.5
0 20
-0.5
-100
-50
0
Figure 10.5. Comparison of the true 50
(numerical)
100
0solution of the
SH equation with ε=0.5 and initial condition u0 (x)=A0 (εx) cos(x),
A0 (X) = 1/ cosh(X), with the (numerical) solution A (dashed line)
of the GL equation with initial condition A0 (X), indicating that
ψGL (x, t) = εA(εx, ε2 t) cos(x) gives a good approximation for all
times considered.
and by choosing A3 = − 64
1 3
A we obtain (10.14). In order to obtain
(10.15) Res(εψ5 ) = O(ε5 )
we define
εψ5 = εψ4 + (ε2 A12 (εx, ε2 t)E + ε4 A32 (εx, ε2 t)E3 + c.c.)
where A12 and A32 are new functions which are chosen below. We find
Res(εψ5 ) = ε4 E(−∂T A12 + 4∂X
2
A12 + A12 − 3A2 A12 − 6|A|2 A12 − 4i∂X
3
A)
+ε4 E3 (−64A32 − 3A2 A12 − 96∂X A3 ) + O(ε5 ).
By choosing A12 to satisfy the inhomogeneous linearized GL equation
(10.16) 2
∂T A12 = 4∂X A12 + A12 − 3A2 A12 − 6|A|2 A12 − 4i∂X
3
A,
and A32 to satisfy −64A32 − 3A2 A12 − 96∂X A3 = 0, we obtain (10.15). In
order to achieve Res(εψn ) = O(εn ) we choose
(m)
α
(10.17) εψn = εα(m)+j Amj (X, T )Em
m=−N,...,N j=1
|
u(k)|
O(1)
O(ε)
O(ε2 )
−1 |O(ε)|
k
As before A11 satisfies the GL equation. The evolution of the A1j for
j ≥ 2 are determined by inhomogeneous linearized GL equations, and the
Amj for m = ±1 are determined by algebraic equations. The solutions Amj
of the resulting system exist as long as the solution A11 of the GL equation
326 10. Dynamics of pattern and the GL equation
exists. In case of an odd nonlinearity such as for (10.4) we can set Amj = 0
for m odd.
Estimates for the residual in Cb0 . The formal orders of the residual
can be turned into estimates in norms. We find for instance
Res(εψGL )C 0 ≤ s1 + s2 + s3 ,
b
where
s1 = 3ε3 E 3 A(ε·)3 C 0 ≤ 3ε3 A3C 0 ,
b b
s2 = ε4 E(4i∂X
3
A(ε·))C 0 ≤ 4ε4 AC 3 ,
b b
s3 = ε5 E(∂X
4
A(ε·))C 0 ≤ ε5 AC 4 ,
b b
and
sup εψGL (t) − εψn (t)C 0 ≤ Cε3 .
b
t∈[0,T0 /ε2 ]
properties of the linear part in (10.16), then we need A(·, T ) ∈ Cb7 . Since the
structure of the approximation equations is the same for the next orders, we
lose three derivatives in every step, such that the estimates are possible with
θA = 3(n − 3) + 1. However, by using the smoothing properties of the linear
part in (10.16) and by constructing the approximation in a more clever way,
θm can be chosen much smaller, see below.
trivial than before due to the scaling properties of the L2 -norm. For εψGL
we find for θ > 1/2 similar as before
Res(εψGL )H θ
≤ C ε3 A(ε·)2C θ A(ε·)H θ + ε4 A(ε·)H θ+3 + ε5 A(ε·)H θ+4 .
b
and
sup εψGL (t) − εψn (t)C 0 ≤ Cε3 .
b
t∈[0,T0 /ε2 ]
spaces to use the advantages of both spaces Cb and H , namely that Cb0
0 θ
contains non-decaying functions for |x| → ∞ and the fact that Fourier trans-
form is a suitable tool in H θ . The residual is estimated in Hul
θ as follows.
and
sup εψGL (t) − εψn (t)C 0 ≤ Cε3 .
b
t∈[0,T0 /ε2 ]
The equations for the error. Estimates for the residual, even in norms,
are only a necessary condition for the GL equation to correctly predict the
behavior of the original systems. Such estimates are not sufficient. The
errors can accumulate in time and there are a number of counter-examples,
cf. [Sch95b, Sch05, SSZ15], §10.6.3, and §11.5.3, showing that modu-
lation equations, although derived in a formally correct way, make wrong
predictions about the behavior of the original system.
The error
εβ R = u − εψ,
i.e., the difference between the solution u and the approximation εψ = εψn ,
with β and n suitably chosen, is estimated with the help of Gronwall’s
inequality. It satisfies
∂t R = −(1 + ∂x2 )2 R + ε2 R − 3ε2 ψ 2 R − 3εβ+1 ψR2 − ε2β R3 + ε−β Res(εψ).
We use the variation of constant formula
t
tΛ 2
(10.19) R(t) = e R(0) + ε e(t−τ )Λ F (τ ) dτ,
0
we have to estimate the other terms in F , and thirdly the linear semigroup
(eΛt )t≥0 .
In order to make clear that estimates for F are easy to obtain indepen-
dently of the chosen norm we will give these estimates in all three spaces.
θA
Therefore, we assume that A ∈ C([0, T0 ], Cb,unif ), or A ∈ C([0, T0 ], H θA ), or
θA
A ∈ C([0, T0 ], Hul ), is a solution of the GL equation with θA ≥ 0 sufficiently
large. Then we obtain
Lemma 10.2.6. For every θ ≥ 1 there is a C > 0 such that for all ε ∈ (0, 1]
we have
ε2 F X θ ≤ C(ε2 RX θ + εβ+1 R2X θ + ε2β R3X θ + ε2 ),
where X θ stands either for Cb0 , H θ , or Hul
θ depending on the chosen phase
space for A.
Proof. All spaces X θ are closed under multiplication if θ > 1/2. Since the
H θ -norm of ψ is of order O(ε−1/2 ) we have to estimate ψ in the Cbθ -norm,
where θ = 0 in case X θ = Cb0 and θ = θ in the other cases. Then the
estimate follows from
ψ 2 RX θ ≤ Cψ2 θRX θ ,
Cb
R X θ ≤
3
CR3X θ ,
−β
ε Res(εψ)X θ ≤ Cε , 2
and similar estimates for the other terms in ψ. Sobolev’s embedding theorem
gives the additional bounds
AC θ ≤ CAH θ+1
and AC θ ≤ CA
θ+1 .
b b Hul
The associated estimate in 0
is less trivial and can be found in
Cb,unif
θ.
[KSM92, Lemma 2.3]. Our favorite choice is the space Hul
Lemma 10.2.8. For every θ ≥ 0 there exists a C > 0 such that for all
ε ∈ (0, 1] the semigroup (etΛ )t≥0 generated by Λ = −(1 + ∂x2 )2 satisfies
sup etΛ H θ →H θ ≤ C.
ul ul
t∈[0,T0 /ε2 ]
and thus
where εψn is the full ansatz (10.17), and εψGL is the lowest order approxi-
mation (10.6).
Proof. We use Lemma 10.2.6 and Lemma 10.2.7, respectively Lemma 10.2.8,
to bound the mild solutions of the equations for the error in the variation
of constant formula (10.19). By Theorem 5.2.22 we have local existence and
uniqueness of the solutions in Cb0 , H θ , or Hul
θ for θ > 1/2. The solutions
exist until the norm of the solutions becomes infinite, and thus we only have
to derive a priori estimates to guarantee that the solutions exist and stay
10.2. The Swift-Hohenberg equation 331
≤Cεβ + Cε3/2
Remark 10.2.10. For clarity we formulate the last step of Theorem 10.2.9
in a different way. We define t = sup{t : R(t)X θ ≤ M }, with M defined
as above. We are done, if we show t ≥ T0 /ε2 . Choosing ε0 > 0 in the above
way guarantees the validity of (10.22) for all t ≤
t and so the validity of
t
R(t)X θ ≤ 2CT0 + Cε2 R(τ )X θ dτ
0
for all t ≤
t. As above, Gronwall’s inequality implies supt∈[0,T0 ε2 ] R(t)X θ ≤
M , and so t ≥ T0 /ε2 .
Remark 10.2.11. The estimate (10.21) shows that already the original
ansatz yields a suitable approximation, i.e., for ε > 0 sufficiently small we
see the dynamics predicted by the GL equation in the original system. For
times t ∈ [0, T0 /ε2 ] the error of order O(εmin(3/2,β) ) in X θ and by Sobolev’s
embedding theorem also in Cb0 is much smaller than the solution and ap-
proximation which are both of order O(ε) in Cb0 .
332 10. Dynamics of pattern and the GL equation
k kc k
Figure 10.7. The curves of eigenvalues k → Reλn (k) in the stable case
are shown in the left panel. In the right panel instability occurs due to
one curve taking positive values at the wave numbers ±kc .
with complex-valued kernels s2,n,n1 ,n2 (k, k − l, l), s3,n,n1 ,n2 ,n3 (k, k − l, l −
m, m), etc. Like the eigenvalues and eigenvectors, the kernels additionally
depend on ε2 , e.g. s2,n,n1 ,n2 (k, k − l, l; ε2 ). We have explicit representations
such as
9 :
s2,n,n1 ,n2 (k, k − l, l) = ϕ∗n (k), e−ikx N2 (ϕn1 (k − l)ei(k−l)x , ϕn2 (l)eilx )
L2 (Σ)
where ϕ∗n (k) is the associated adjoint eigenfunction w.r.t. the scalar product
·, ·L2 (Σ) , with
ϕ∗1 , ϕ1 L2 (Σ) = 1.
The modes u of the Fourier transformed SH equation solely correspond
to the modes c1 . For general systems additionally to these modes we have
infinitely many modes c2 , c3 , . . . which are damped with some exponential
rate. Since the Fourier transform of x → εA(εx, ε2 t)eikc x is given by k →
336 10. Dynamics of pattern and the GL equation
∞ ∞
(10.34) +2 s211n (kc , −kc , 2kc ; 0)
A1,−1 (K−κ, T )An,2 (κ, T ) dκ
n=1 −∞
∞ ∞
+3s31111 (kc , kc , kc , −kc ; 0) 1,1 (K−κ1 , T )
A
−∞ −∞
1,1 (κ1 −κ2 , T )A
×A 1,−1 (κ2 , T ) dκ2 dκ1 ,
where we used the symmetry of the bilinear terms in their arguments. With
k−jkc = εK we find at ε in the n-th equation for the modes concentrated
at jkc for j = 0, 2 and n ≥ 1 that
0 =λn (0, 0)An,0 (K, T )
∞
+2s2n1−1 (0, −kc , kc ; 0) A 1,1 (κ, T ) dκ,
1,−1 (K−κ, T )A
−∞
n,2 (K, T )
0 =λn (2kc , 0)A
∞
+s2n11 (2kc , kc , kc ; 0) A 1,1 (κ, T ) dκ.
1,1 (K − κ, T )A
−∞
Since Reλn (0, 0) < 0 and Reλn (2kc , 0) < 0 these algebraic relations deter-
mine A n,0 and An,2 in terms of A 1,1 and A 1,−1 such that (10.34) becomes
the GL equation in Fourier space, namely
1,1 (K, T ) =ν0 A
∂T A 1,1 (K, T ) − ν2 K 2 A
1,1 (K, T )
∞ ∞
(10.35) +ν3 1,1 (K − κ, T )A
A 1,1 (κ − κ 1,−1 (
, T )A κ, T ) d
κ dκ,
−∞ −∞
10.4. An abstract approximation result 337
with coefficients
1
(10.36) ν2 = ∂k2 λ1 (kc , 0), ν0 = ∂α λ1 (kc , 0),
2
and
∞
ν3 = − 4 s211n (kc , kc , 0; 0)s2n1−1(0, −kc , kc ; 0)/λn (0, 0)
n=1
∞
(10.37) −2 s211n (kc , −kc , 2kc ; 0)s2n11 (2kc , kc , kc ; 0)/λn (2kc , 0)
n=1
+3s31111 (kc , kc , kc , −kc ; 0).
This derivation of the Fourier transformed GL equation (10.35) explains the
universality of the GL equation (10.3) and the formulas (10.36) for ν2 and
ν0 . The universal behavior can also be seen when computing the higher
order approximations. We find linearized GL equations at c1 and k = kc
and algebraic equations for all other modes, cf. Exercise 10.2. We leave the
case of a complex critical curve of eigenvalues λ1 to Exercise 10.3.
Remark 10.3.2. The formula (10.37), using the complete expansion in
eigenfunctions, is conceptually useful but of little practical use. Instead, the
exponentially damped modes can be collected in one big vector, and the
value of ν3 can be found by restricting to 2π/kc -spatially periodic functions,
see for instance §10.5.1, and also Remark 10.5.2. This often allows to ob-
tain this coefficient from the literature about the associated center manifold
reduction.
Clearly we have εEc Ψ = O(ε) and εEs Ψ = O(ε2 ). The idea is to split
the error into a critical and a non-critical part which are scaled differently,
i.e.,
εβ R = εβ Rc + εβ R−c + εβ+1 Rs ,
where, depending on the chosen space, the support of R±c in Fourier space
is contained in a set equal to or slightly larger than the support of E±c ,
and where the support of Rs in Fourier space is contained in a set equal to
or slightly larger than the support of Es . The parts of the error satisfy a
system of the form
∂t Rc =Lε (∂x )Rc + O(ε2 |Rc | + ε2 |Rs |) + O(ε2 ),
∂t Rs =Lε (∂x )Rs + O(|Rc | + ε|Rs |) + O(1).
This is because
Ec applied to the quadratic interaction of critical modes vanishes,
i.e., Ec (ε∂x2 ((Ec Ψ)Rc )) = 0. Since Rs is exponentially damped by the
semigroup generated by Lε (∂x ), it is easy to obtain an estimate Rs =
O(|Rc |) + O(1). Inserting this into the first equation yields
∂t Rc = Lε (∂x )Rc + O(ε2 |Rc |) + O(ε2 ).
10.4. An abstract approximation result 339
A direct estimate for all t ∈ [0, T0 /ε2 ] with the help of Gronwall’s inequality
is now possible and so we obtain Rc = O(1) and Rs = O(1) for all t ∈
[0, T0 /ε2 ].
In the following we will extract assumptions which will allow us to handle
not only the KS equation, but very general pattern forming systems.
(10.40) ∂t u = Λu + N (u)
Then we introduce mode filters Ec and Es for extracting the critical and
stable modes. For these we assume
(A1) There exist linear operators Ec and Es which are continuous both
in X and Y. Moreover, they commute with the semigroup (etΛ )t≥0 , i.e.,
Using these mode filters we split the error R in two parts which we scale
differently, i.e., we set R = Rc +εRs and define Rc and Rs to be the solutions
340 10. Dynamics of pattern and the GL equation
of
It turns out that this system is a good interface for coupling the systems
which we have in mind to the abstract set-up. For the formulation of the
other assumptions we need additional mode filters Ech and Esh . This is due
to the fact that for functional analytic reasons Ec and Es cannot be chosen
as projections in Hulθ . We assume
(A2) There exist linear operators Ech and Esh which are continuous both
in X and Y and satisfy Ech Ec = Ec Ech = Ec and Esh Es = Es Esh = Es . The
mode filters applied to the semigroup give the following estimates. There
exist CΛ , σc ≥ 0, α ∈ [0, 1), and a σs > 0 such that for all t ≥ 0 and ε ≥ 0
we have
2
etΛ Ech Y→Y ≤CΛ eσc ε t ,
2
etΛ Ech X →Y ≤CΛ eσc ε t ,
etΛ Esh Y→Y ≤CΛ e−σs t ,
etΛ Esh X →Y ≤CΛ max(1, t−α )e−σs t .
(A3) The nonlinear terms obey the following estimates. There exist con-
stants C1,c and C1,s , monotonically growing functions C2,c (Mc , Ms ) and
C2,s (Mc , Ms ) and an ε0 > 0 such that for all ε ∈ (0, ε0 ) we have
(A4) The residual terms obey the following estimates. There exist con-
stants Cres and ε0 > 0 such that for all ε ∈ (0, ε0 ) we have
sup Ec Res(εΨ(τ ))Y ≤Cres εβ+2 ,
τ ∈[0,T0 /ε2 ]
2
≤ CΛ eσc ε t Rc (0)Y
t
2
+ CΛ eσc ε (t−τ ) (C1,c ε2 (Rc (τ )Y + Rs (τ )Y )
0
+ C2,c (Mc , Ms ) min(ε3 , εβ )(Rc (τ )Y + Rs (τ )Y )2 ) dτ
t
+ ε−β
2
CΛ eσc ε (t−τ ) Cres εβ+2 dτ
0
and
Rs (t)Y ≤ etΛ Esh Y→Y Rs (0)Y
t
+ e(t−τ )Λ Esh X →Y ε−(β+1) Es (N (εΨ(τ ) + εβ R(τ )) − N (εΨ(τ )))X dτ
0
t
−(β+1)
+ε e(t−τ )Λ Esh Y→Y Es Res(εΨ(τ ))Y dτ
0
≤ CΛ e−σs t Rs (0)Y
t
+ CΛ max(1, (t − τ )−α )e−σs (t−τ ) (C1,s Rc (τ )Y + C1,s εRs (τ )Y
0
+ C2,c (Mc , Ms ) min(ε, εβ−1 )(Rc (τ )Y + Rs (τ )Y )2 ) dτ
t
−(β+1)
+ε CΛ e−σs (t−τ ) Cres εβ+1 dτ,
0
ε ∈ (0, ε1 ) we have
(10.44) Ss (t) ≤ (CΛ Ss (0) + CΛ Cσ Cres + 1) + 2CΛ Cσ C1,s Sc (t),
as long as Sc (t) ≤ Mc and Ss (t) ≤ Ms , if 0 < ε1 1 is chosen so small that
where
β0 =CΛ eσc T0 Sc (0)
+ T0 CΛ eσc T0 Cres + T0 CΛ eσc T0 C1,c (CΛ Ss (0) + CΛ Cσ Cres + 1) + 1,
β1 =CΛ eσc T0 C1,c (1 + 2CΛ Cσ C1,s ),
if we choose 0 < ε2 1 so small that for given Mc and Ms we have
(10.46) CΛ eσc T0 T0 C2,c (Mc , Ms ) min(ε3 , εβ )(Mc + M2 )2 ) ≤ ε2
for all ε ∈ (0, ε2 ). Using Gronwall’s inequality immediately yields
Sc (t) ≤ β0 eβ1 T0 =: Mc ,
and due to (10.44) we define
Ms := (CΛ Ss (0) + CΛ Cσ Cres + 1) + 2CΛ Cσ C1,s Mc .
To this Mc and Ms we define ε0 = min(ε1 , ε2 ) where ε1 > 0 is chosen so
small that for all ε ∈ (0, ε1 ) condition (10.45) is satisfied and ε2 > 0 so small
that for all ε ∈ (0, ε2 ) condition (10.46) is satisfied.
With the help of the mode filters we are now able to give the
Proof of Lemma 10.2.8. For the proof of the estimate we will use Lemma
8.3.7. Unfortunately, a direct estimate of k → e−(1−k ) t C 2 would lead to
2 2
b
some unwanted growth rates in time. Hence, we introduce two mode filters
Ec and Es satisfying Ec + Es = 1 where Ec is defined by E c = χ1 + χ−1 ,
∞
with χ ∈ C0 (R, [0, 1]) an even cut-off function with χ(k) = 1 for |k| ≤ 1/10
and χ(k) = 0 for |k| ≥ 1/5, and where χj (k) = χ(k − j). Then we find
and immediately
s (k)e−(1−k
Es etΛ H θ →H θ ≤ Ck → E
2 )2 t
C 2 ≤ Ce−σt
ul ul b
344 10. Dynamics of pattern and the GL equation
for a σ > 0 since the semigroup decays with some exponential rate for the
wave numbers in the support of Es . The estimate for the Ec part uses that
for θ > 1/2 we have
Ec uH θ ≤CEc uL2 ≤ C
χ1 uL2 ≤ C χ(ue−ik·x )L∞
χ1 uL∞ ≤ C
ul ul ul
−ik·x
≤CSε (
χ(ue χ(ue−ik·x ))H θ
))L∞ ≤ CSε (
ul
−ik·x
≤C
χ(ue )H θ ≤ CEc uH θ ≤ CEc uL2
ul ul ul
where the scaling operator Sε is defined by (Sε u)(x) = u(εx). Hence, for
functions with a fixed compact support in Fourier space we have the equiva-
χ(ue−ik·x ))H θ .
lence of the norms Ec uH θ , respectively Ec uL2 , and Sε (
ul ul ul
W.r.t. the last norm Lemma 8.3.7 requires us to estimate k → eλε (k)t C 2
b
with
λε (k) = −(1 − (εk − 1))2 )2 .
Since λε (k) ∼ −(εk)2 for small k the term eλε (k)t behaves as e−k T with
2
T = ε2 t and so its Cb2 -norm is uniformly bounded for t on the O(1/ε2 ) time
scale.
Since the nonlinear terms are given by N (u) = ∂x (u2 ) we make the
choice
(A0) We set X = H θ−1 and Y = H θ .
The semigroup (etΛ )t≥0 is defined as a multiplier in Fourier space via
tΛ F where e4
etΛ = F −1 e4 tΛ (k) = e(−(1−k2 )2 +ε2 )t . Therefore, we also introduce
= F −1 χj F F −1 e4
tΛ F = E etΛ
j
and
(e4 = (χj e4
2 )2 +ε2 )t 2 )2 +ε2 )t
tΛ χ )(k) = e(−(1−k χj (k) = χj (k)e(−(1−k tΛ )(k).
j
10.4. An abstract approximation result 345
Es (A(ε·)E)C θ ≤CEs (A(ε·)E) −1
b
L1 ≤ C(1 − χ1 )A(ε (· − 1))L1
θ θ
+ε β−1
Rc 2H θ +εβ Rc H θ Rs H θ +εβ+1 Rs 2H θ ).
As for the SH equation, due to Ψc H θ = O(ε−1/2 ) and Ψs H θ = O(ε−1/2 )
it is essential at this point to estimate Ψc and Ψs in Cbθ and not in H θ .
The residual terms are of the form
N
Res(εΨ) = εβj sj (εx)eijx ,
j=−N
with coefficients βj satisfying β±1 = 4 and βj ≥ 3 for all other j ∈ Z\{−1, 1}.
Hence, it remains to prove
Ec εβj sj (ε·x )eij·x H θ ≤ Cε7/2 ,
but this follows exactly as the estimate for Es (A(ε·)E)C θ in the proof
b
of Lemma 10.4.4. With this remark it is an easy exercise to check that
the approximation from Example 10.3.1 satisfies the Assumption (A4) with
β = 3/2 and εΨ from (10.26).
(A4) The residual terms obey the following estimates. There exist con-
stants Cres and ε0 > 0 such that for all ε ∈ (0, ε0 ) we have
sup Ec Res(εΨ(τ ))H θ ≤Cres ε7/2 ,
τ ∈[0,T0 /ε2 ]
We delegate the missing details to the exercises. Here, we only remark that
due to the fact that we have to expand the curves of eigenvalues at the wave
number k = kc , cf. (10.29) and Exercise 10.6, we need that the solutions of
10.5. Reaction-Diffusion systems 347
the GL equation have to be three spatial derivatives more regular than the
solutions of the underlying pattern forming system.
Hence, the KS equation with the previous choices for the mode filters
and H θ as phase space satisfies the assumptions (A0)-(A4) and so we have
the following approximation theorem.
Theorem 10.4.5. Fix θ ≥ 1. Let A ∈ C([0, T0 ], H θ+3 ) be a solution of the
GL equation (10.27). Then there exist a C > 0 and an ε0 > 0 such that for
all ε ∈ (0, ε0 ) we have solutions u of the KS equation (10.38) with
sup u − εΨGL H θ ≤ Cε3/2
t∈[0,T0 /ε2 ]
Hence, Reλj (k, μ) < −σ < 0 for all eigenvalues and all k ∈ R except of
a small neighborhood of kc , and we have Reλ1 (kc , μ) = O(ε2 ) > 0.
Plugging (10.49) into (10.48) and sorting the terms w.r.t. coefficients of
εm En , where En = einkc x , gives
εE : 0 = 0,
ε2 E2 : μc (2kc )Φ2 + A2 B(φ1 , φ1 ),
0=L
ε2 E0 : μc (0)Φ0 + 2|A|2 B(φ1 , φ1 ),
0=L
10.5. Reaction-Diffusion systems 349
and at ε3 E1 we find
0 1
∂T A = ∂μ λ1 (kc , μc ) − 12 ∂k2 λ1 (kc , μc )∂X
2 A
(10.51) & '
+ 2B(Aφ1 , Φ0 ) + 2B(Aφ1 , Φ2 ) + 3C(Aφ1 , Aφ1 , Aφ1 ), φ∗1 .
real-valued.
Thus, under rather generic assumptions on (10.48) we find that the
envelope A of the critical modes satisfy a GL equation. For a given problem,
of course it remains to determine the coefficients ν2 , ν0 , ν3 .
Remark 10.5.2. In the literature, e.g., [Man92, CH93b, Mie02, Pis06,
Mer15] the following derivation of the GL equation can be found. Plugging
the extended ansatz
(10.55) u(x, t) = εΨ(x, t) + ε3 Φ3 (εx, εt)eikc x + c.c.,
into (10.48) yields at ε3 E1
0 1
c )Φ3 = − ∂T Aφ1 + ν0 + ν2 ∂ 2 Aφ1
L(k X
c ) which is precisely
orthogonal to the zero eigenvector of the adjoint of L(k
(10.53).
0.97
φ1 (kc ) ≈ . Moreover,
−0.24
1 1
B(φ1 , φ1 ) = B(φ1 , φ1 ) = q , with q = 2bφ11 φ12 + φ211 ≈ −0.0563,
−1 b
−1.2 0 1 1.03
Φ2 = q , Φ0 = q , C(φ1 , φ1 , φ1 ) = φ21 φ2 , φ∗1 ≈ ,
0.13 1 −1 0.28
which fit well with the numerical simulations for the Schnakenberg model
(9.23) shown in Figure 10.8.
but this follows exactly as the estimate for the KS equation (10.38). There-
fore, (A4) is also valid for reaction-diffusion systems.
Theorem 10.5.4. Fix θ ≥ 1, assume that the reaction-diffusion system
(10.48) satisfies (SpecRD ), and let A ∈ C([0, T0 ], Hul
θ+3
) be a solution of the
GL equation (10.53). Then there exist a C > 0 and an ε0 > 0 such that for
all ε ∈ (0, ε0 ) we have solutions v of the reaction-diffusion system (10.48)
with
sup v − εΨGL H θ ≤ Cε2 ,
ul
t∈[0,T0 /ε2 ]
10.5.3. The Hopf bifurcation case. The GL equation can also be de-
rived in case when the bifurcating pattern is oscillatory in time without
oscillations in space, i.e., when a Hopf bifurcation occurs at the wave num-
ber k = 0, see Figure 9.8 d). In this case, the number of unbounded spatial
directions plays no role, and thus we consider x ∈ Rd . The most famous
example is the so called Brusselator, cf. (9.3) and [Kur84],
∂t v1 = d1 Δv1 + a − (b + 1)v1 + v12 v2 , ∂t v2 = d2 Δv2 + bv1 − v12 v2 ,
where a, b, d1 , d2 are non-negative constants, and v1 , v2 are functions of t ≥ 0
and x ∈ Rd . For this reaction-diffusion system there exists a unique uniform
steady state (v1∗ , v2∗ ) = (a, b/a). Introducing new coordinates (v1 , v2 ) =
(v1∗ + u1 , v2∗ + u2 ) gives the system
∂t u1 = d1 Δu1 + (b − 1)u1 + a2 u2 + f (u1 , u2 ),
(10.57)
∂t u2 = d2 Δu2 − bu1 − a2 u2 − f (u1 , u2 ),
where
f (u1 , u2 ) = (b/a)u21 + 2au1 u2 + u21 u2 .
We fix a and take b as a control parameter. The stability of u1 =u2 = 0
is determined by the linearization of (10.57). We find the eigenfunctions
(u1 (k), u2 (k))eik·x , where the associated eigenvalues λ satisfy
λ2 + α(|k|)λ + β(|k|) = 0,
d
with |k|2 = 2
j=1 kj and k = (k1 , . . . , kd ) ∈ Rd . The constants are given by
α(q) =1 + a2 − b + (d1 + d2 )q 2 ,
β(q) =a2 + (a2 d1 + (1 − b)d2 )q 2 + d1 d2 q 4 .
10.5. Reaction-Diffusion systems 353
possess two surfaces of eigenvalues λ± ∈ C ∞ (Uρ , C), with associated eigen-
0
± ∈ C ∞ (Uρ0 , Cq ), such that
functions ϕ
± (k, ε2 ) = ±iω + ε2 (λ0 ± iν0 ) + |k|2 (λ2 ± iν2 ) + O(ε3 + |k|3 ),
λ
with constants λ0 > 0, λ2 < 0, and ω > 0. Denote by Σ− (k, ε2 ) the set of all
ε2 ). Then, there exists an ε-independent constant
other eigenvalues of Λ(k,
0 > 0 such that
σ
;
−
sup Re Σ (k, ε ) < −
2
σ0 .
k
Moreover, λn (k) → −∞, for |k| → ∞ and fixed n, or for n → ∞ and fixed
k.
The trivial solution (Θ, ψ) = (0, 0) is stable if λ1 (k) < 0 for all k ∈ R.
Instability occurs when the curve λ1 touches the axis λ = 0 at a wave number
k = kc ∈ R for a parameter value R = Rc . This leads to the conditions
Rk 2 !
λ1 = 2 2
− (π 2 + k 2 ) = 0
π +k
and
R Rk 2 π2R !
∂k2 λ1 = − − 1 = − 1 = 0.
π 2 + k 2 (π 2 + k 2 )2 (π 2 + k 2 )2
From this we find πR1/2 = π 2 + k 2 and λ1 = R − 2πR1/2 . This shows that
λ1 = 0 for R = Rc = 4π 2 ≈ 39.48 at the critical wave number k = kc = π,
see Figure 10.9.
0
λ1(k)
−30
−60 λ2(k)
−6 −4 −2 0 2 4 k 6
Remark 10.6.2. The value Rc for spectral instability coincides with the
critical value Rc,e for which energetic stability of the trivial solution can be
shown. For this we multiply (10.65) with ψ and integrate over Ω to obtain
− |∇ψ|2 dx dy = −R Θ∂x ψ dx dy,
Ω Ω
we find
ik
ψ(x, y) = R
Θ(k, n, t)eikx dk sin(nπy),
k2 + n2 π 2
n∈N R
with eigenvalues
Rk 2
λn (k, ε2 ) = − (n2 π 2 + k 2 )
n2 π 2 + k 2
and symmetrized complex-valued kernels
1 i(k − l)
s2,n,n1 ,n2 (k, k − l, l) = Rπ (−n1 il + i(k − l)n2 )δn−n1 −n2
8 (k − l)2 + n21 π 2
1 il
+ Rπ 2 (−n2 i(k − l) + iln1 )δn−n1 −n2
8 l + n22 π 2
1
since 0 sin(nπy) cos(n1 πy) sin(n2 πy)dy = δn−n1 −n2 /4. Following the calcu-
lations of §10.5 we have to compute
ν2 =∂k2 λ1 (π; 0)/2 = 2,
1
ν0 =∂ε2 λ1 (π; 0) = ,
2
∞
ν3 = − 4 s211n (kc , kc , 0; 0)s2n1−1 (0, −kc , kc ; 0)/λn (0; 0)
n=1
∞
−2 s211n (kc , −kc , 2kc ; 0)s2n11 (2kc , kc , kc ; 0)/λn (2kc ; 0)
n=1
= − 4s2112 (kc , kc , 0; 0)s221−1 (0, −kc , kc ; 0)/λ2 (0; 0)
− 2s2112 (kc , −kc , 2kc ; 0)s2211 (2kc , kc , kc ; 0)/λ2 (2kc ; 0) = −4π 2 .
Thus, we finally find the GL equation
1
(10.69) 2
∂T A = 2∂X A + A − 4π 4 |A|2 A,
2
cf. Exercise 10.8. Via the solutions of the GL equation we describe slow
spatial and temporal modulations of small amplitude of the underlying con-
vection pattern via the approximation
ψ
(x, y, t) = εΨGL (x, y, t) + O(ε2 )
Θ
(10.70)
2πi iπx
= εA(X, T ) e sin πy + c.c. + O(ε2 ),
1
where X = εx and T = ε2 t are the long spatial and long temporal scale.
An immediate observation is that the GL equation (10.69) possesses stable
steady states A = 2√12π2 eiφ which are constant in space, where φ ∈ [0, 2π) is
a free phase. For φ = 0 this formally yields the steady convection rolls
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
u1 0 4π 2 cos πy sin πx
⎝u2 ⎠ = ⎝ 0 ⎠ + √ ε ⎝−4π 2 sin πy cos πx⎠ + O(ε2 ).
(10.71)
2 2π 2
T 1−y sin πy cos πx
10.6. Convection problems 361
√
(±1, 0) and (±1/2, ± 3/2). Modulations of hexagons are described via the
ansatz
√
1 3
u(x, y, t) =εA1 (εx, εy)eix + εA2 (εx, εy)ei( 2 x+ 2
y)
√
1 3
+ εA3 (εx, εy)ei( 2 x− 2
y)
+ c.c. + O(ε2 )
leading to the system
∂T A1 =4∂X2
A1 + A1 − 3A1 |A1 |2 − 6A1 (|A2 |2 + |A3 |2 ),
√
3 1
∂T A2 =4( ∂X + ∂Y )2 A2 + A2 − 3A2 |A2 |2 − 6A2 (|A1 |2 + |A3 |2 ),
√2 2
3 1
∂T A3 =4( ∂X − ∂Y )2 A3 + A3 − 3A3 |A3 |2 − 6A3 (|A1 |2 + |A2 |2 ).
2 2
The degeneracy w.r.t. diffusion again reflects the fact that a big part of the
ring of unstable modes is not covered by the ansatz.
iii) For rolls there are two possibilities. If the Fourier modes are concen-
trated in O(ε)-neighborhoods of the two wave vectors (±1, 0), then the above
system for the squares, but with A01 = 0, is obtained. The second possibil-
ity is an O(ε)-concentration in k-direction and an O(ε1/2 )-concentration in
l-direction near (±1, 0). The ansatz is then given
H s -spaces. In Hul
s -spaces new functional difficulties occur due to the projec-
tion on the divergence-free vector fields in case of more than one unbounded
space directions, cf. [CZ15]. For other approaches to describe planar pat-
tern via amplitude equations see also [Mel00, EINP03], and the further
reading at the end of this Chapter.
The anisotropic situation. Next we consider the case α > 0, i.e., we
break the rotational symmetry in (10.77). Then unstable modes occur in
O(ε)-neighborhoods of the wave vectors (±1, 0). With the ansatz
u(x, y, t) = εA(εx, εy)eix + εA(εx, εy)e−ix + O(ε2 )
a two-dimensional GL equation can be derived
(10.79) 2
∂T A = 4∂X A + α∂Y2 A + A − 3A|A|2 .
Our abstract approximation result, Theorem 10.4.3, applies in this situation
since now the quadratic interaction of the critical modes gives non-critical
modes.
The prototype of an anisotropic pattern forming system is electro-convec-
tion in nematic liquid crystals [OD08]. In this problem the pattern forma-
tion is due to some external time-periodic forcing. It can be brought back
with Floquet theory to a problem with an autonomous linear part. The rele-
vant nonlinear terms can be made autonomous with the help of normal form
transformations. Hence, the above theory applies to time-periodic systems
as well [SU07]. See §10.6.5.
the underlying ideas. The precise conditions on the underlying systems are
similar to the previous ones and can be found in [SU07].
We consider a translationally invariant system
(10.82) ∂t V = M (t)V + N (t, V )
which depend on some parameter α ∈ R, on a cylindrical domain R × Σ,
where M (t)V stands for the linear and N (t, V ) for the nonlinear terms
satisfying M (t) = M (t + 2π/ω0 ) and N (t, V ) = N (t + 2π/ω0 , V ) for an
ω0 > 0. As explained above, electro-convection is an anisotropic problem.
For such problems the subsequent ideas also apply in domains R2 × Σ, and
yield systems of more complicated amplitude equations.
In order to analyze the stability of the trivial solution V =0 in (10.82),
we consider the linearized system
(10.83) ∂t V = M (t)V.
Due to the translational invariance of the problem the solutions are given
by Floquet-Fourier modes
(10.84) V = ϕm (k, z, t)eikx eλm (k)t ,
with k ∈ R, m ∈ N, z ∈ Σ, and ϕm periodic in t, i.e.
ϕm (·, ·, t) = ϕm (·, ·, t + 2π/ω0 ).
If V = 0 is asymptotically stable, then we have that Re λm (k) < 0 for all
m ∈ N and k ∈ R. We assume that V = 0 becomes unstable at α = αc ,
i.e., at α = αc there exists one curve of Floquet exponents λ1 satisfying
Reλ1 (kc ) = 0 for some kc > 0. We set ε2 = α−αc > 0, write λm = λm (k, ε2 ),
and assume that all Floquet exponents possess a real part strictly less than
−σ0 for a σ0 > 0, except of λ1 (k) for k in small neighborhoods of ±kc . Hence
we have the same assumptions as in the autonomous case, except that the
curve of critical eigenvalues has to be replaced by a curve of critical Floquet
exponents.
Like for the autonomous case, it turns out that for the mathematical
analysis, it is essential to consider the Fourier transformed system w.r.t. the
unbounded spatial variable. In Fourier space (10.82) is given by
(10.85) ∂t V (k, t) = M
4(k, t)V (k, t) + N
(V )(k, t),
for every k ∈ R and t ∈ [0, 2π/ω0 ) a Floquet Schauder basis (ϕj (k, t))j∈N of
L2 (Σ), with ϕj (k, t) = ϕj (k, t + 2π/ω0 ) satisfying
4(k, t)ϕj (k, t) − λj (k)ϕj (k, t).
∂t ϕj (k, t) = M
The functions ϕj are normalized by ϕj (k, 0)L2 =1. For defining pro-
jections onto the ϕj (k, t) we consider the adjoint problem −∂t V (k, t) =
4∗ (k, t)V (k, t). Consequently, also this problem has for every k ∈ R a Flo-
M
quet Schauder basis (ϕ∗j (k, t))j∈N of L2 (Σ), with ϕ∗j (k, t) = ϕ∗j (k, t + 2π/ω0 ),
solving
4∗ (k, t)ϕ∗ (k, t) − λj (k)ϕ∗ (k, t),
−∂t ϕ∗j (k, t) = M j j
and satisfying the orthogonality
(10.86)
ϕ∗i , ϕj L2 (Σ) = δij .
A solution V (k, t) of (10.85) is expanded in terms of the Floquet functions
ϕj (k, t), i.e.
(10.87) V (k, t) =
aj (k, t)ϕj (k, t), with aj (k, t) ∈ C,
j∈N
such that
⎛ ⎞
∂t ⎝ aj (k, t)ϕj (k, t)⎠ =
((∂t
aj (k, t))ϕj (k, t) +
aj (k, t)∂t ϕj (k, t))
j∈N j∈N
= 4 t)ϕj (k, t) + N
aj (k, t)M(k, (V )(k, t).
j∈N
with bilinear and trilinear symmetric terms B and C and introduce the
abbreviations
kc , kc , kc , −kc )A1 |A1 |2 ,
+ 3
ϕ∗1 , C(t,
with ν0 = ∂ε2 λ1 (kc , 0) and ν2 = − 12 ∂k2 λ1 (kc , 0). In (10.92) we replace A0,j
by (10.90), A2,j by (10.91), and obtain the GL equation
(10.93) 2
∂T A1 = ν0 A1 + ν2 ∂X A1 + ν3 (t, ε)A1 |A1 |2 ,
(10.94) 2
∂T A1 = ν0 A1 + ν2 ∂X A1 + ν3 A1 |A1 |2 .
r2
ω2
ω1 r1
10.7.1. The symmetries. Before we start with the linear stability analy-
sis we discuss the symmetries of the underlying PDEs. In cylindrical coor-
dinates (r, φ, z) the velocity field v = (vr , vφ , vz ) and the pressure p satisfy
vφ vφ2 1 2 vr
∂t vr +vr ∂r vr + ∂φ vr +vz ∂z vr − = − ∂r p+ν(Δvr − 2 ∂φ vφ − 2 ),
r r ρ r r
vφ vφ vr 1 2 vφ
∂t vφ +vr ∂r vφ + ∂φ vφ +vz ∂z vφ − = − ∂φ p+ν(Δvφ + 2 ∂φ vr − 2 ),
r r ρr r r
vφ 1
∂t vz +vr ∂r vz + ∂φ vz +vz ∂z vz = − ∂z p+νΔvz ,
r ρ
and
1 vr
∂r vr + ∂φ vφ + ∂z vz + = 0,
r r
where Δ = ∂r2 + 1r ∂r + r12 ∂φ2 + ∂z2 . The equations are invariant under the
transformations of the velocity field defined by
(10.95) (τa v)(r, φ, z) =(vr , vφ , vz )(r, φ, z + a),
(10.96) (Sv)(r, φ, z) =(vr , vφ , −vz )(r, φ, −z),
(10.97) (Rϕ v)(r, φ, z) =(vr , vφ , vz )(r, φ + ϕ, z).
10.7. The Couette-Taylor problem 371
PR II PR I 2 real eigenvalues
4 imaginary (considered parameter region)
Rc
eigenvalues
Re λ
ε2
kc k
λ1
−σ
k
− kc
− kc kc k
such that
1 (y)eikc x ) + c.c. + O(ε2 )
S(εψ) =εA(−X, T ) S(U kc ,0
With the help of the projection Π0 we define the unbounded linear opera-
tor ΛR and the nonlinearity N (R, ·), depending smoothly on the parameters
R, η, and ω, by
ΛR U =Π0 (ΔU − R[(UCou · ∇)U + (U · ∇)UCou ]),
N (R, U ) = − RΠ0 (U · ∇)U,
such that (10.99) writes as
(10.104) ∂t U = ΛR U + N (R, U ).
For the analysis of (10.104) we need the two spaces
Z = {U ∈ H : U ∈ (Hul
2
(Ω))3 , U |R×∂Σ = 0}
and
Z ∗ = {U ∈ H : U ∈ (Hul
1
(Ω))3 }.
It turns out that
Lemma 10.7.3. The operator ΛR is the generator of an analytic semigroup
eΛR t in H with domain of definition Z satisfying
eΛR t Z ∗ →Z ≤ Ct−3/4
for t ∈ (0, 1].
10.7. The Couette-Taylor problem 375
to show that in the associated amplitude equation only four and not eight
different coefficients occur. See Exercise 10.14.
A simple example of an original system with these properties is a system
of coupled KS equations
∂t u = −(1 + ∂x2 )2 u − ∂x u + ε2 u + ∂x (u2 + uv + v 2 ),
(10.105)
∂t v = −(1 + ∂x2 )2 v + ∂x v + ε2 v + ∂x (u2 + uv + v 2 ),
where t ≥ 0, x ∈ R, and (u(x, t), v(x, t)) ∈ R2 . As before 0 < ε2 1 is
used as a small bifurcation parameter. The system is invariant under the
transformation (x, u, v) → (−x, −v, −u).
The linearization at (u, v) ≡ (0, 0) possesses eigenfunctions of the form
(u, v)(x, t) = (ak eikx , bk eikx ), with coefficients ak , bk ∈ C, wave number k ∈
R, and associated eigenvalues
1,2 (k, ε2 ) = −(1 − k 2 )2 ∓ ik + ε2 .
λ
We observe that Reλ 1,2 (k, ε2 ) is positive for wave numbers k close to the
values ±kc = ±1. As above we expect that the bifurcating solutions are
slow modulations in time and in space of the bifurcating pattern eix due to
1,2 . Therefore, we make the ansatz
the form of λ
u = εA(ε(x − cg t), ε2 t)ei(x−ω0 t) + ε2 A2 (ε(x − cg t), ε2 t)e2i(x−ω0 t)
1
+ ε2 A00 (ε(x − cg t), ε2 t) + c.c.,
2
v = εB(ε(x + cg t), ε2 t)ei(x+ω0 t) + ε2 B2 (ε(x + cg t), ε2 t)e2i(x+ω0 t)
1
+ ε2 B00 (ε(x + cg t), ε2 t) + c.c.,
2
with cg , ω0 ∈ R. Inserting the ansatz into the original system and equating
the coefficients in front of εn eimx eijω0 t to zero gives cg = −1 and ω0 = 1.
Setting X1 = ε(x − cg t), X2 = ε(x + cg t), T = ε2 t and eliminating the
algebraic equations for m = 0, ±2 we finally obtain
2 A(X , T ) + A(X , T )
∂T A(X1 , T ) = 4∂X 1 1 1
− 9 (A(X1 , T )|A(X1 , T )|2 + a1 + a2 ),
6
(10.106) 2 B(X , T ) + B(X , T )
∂T B(X2 , T ) = 4∂X 2 2 2
− 9 (B(X2 , T )|B(X2 , T )|2 + b1 + b2 ),
6
where
a1 =A(X1 , T )B(X2 , T )B(X2 , T ),
< =
a2 =A(X1 , T ) B(X2 , T )A(X1 , T )e−2iαT /ε +A(X1 , T )B(X2 , T )e2iαT /ε
2 2
< 2 2
=
+B(X2 , T ) A(X1 , T )B(X2 , T )e4iαT /ε +B(X2 , T )B(X2 , T )e2iαT /ε ,
10.7. The Couette-Taylor problem 377
< =
+A(X1 , T ) B(X2 , T )B(X2 , T )e2iαT /ε +A(X1 , T )A(X1 , T )e−2iαT /ε .
2 2
where L
1
F L (Y, T ) = F (Y + X, T ) dX
2L −L
for a fixed L > 0. In [Sch97, Theorem 3.1] it has been shown that the
non-averaged system (10.106) can be approximated by the averaged system
(10.107) up to an error O(εmin(3/2,2−ν) ) w.r.t. the Hul
θ -norm if L=O(1/εν )
378 10. Dynamics of pattern and the GL equation
cf. [Sch94a]. Moreover, we use the notations Zρ and Yρ for the respective
localized spaces.
Using |ρ b | ≤ b2 ρb and
ρb u 2 dx = − (ρb u ) u dx = − ρ b ( 12 u2 ) dx − ρb u u dx
1 2
= b2
2 ρb u dx − ρb uu dx ≤ 2 ρb u2 dx + ρb |uu | dx
we conclude
> ?
− ρb u(1 + ∂x ) u dx ≤
2 2
ρb (b4 /2 − 1)u2 + 2(1 + b2 )|uu | − u 2 dx,
R R
which yields the desired result after maximizing the integrand w.r.t. u .
b) Applying the estimates of part a) with b = 1 we find
d−1 2
ρ −u(1 + ∂x2 )2 u + du 2 dx ≤ ρ u + (4 + d)|uu | − u 2 dx.
2
Maximizing w.r.t. u gives the result with D = (2d − 2 + (4 + d)2 )/4.
With these estimates we can construct a global semiflow and an absorb-
ing ball for (10.109).
Theorem 10.8.2. The SH equation (10.109) defines a global semiflow u(t) =
St (u0 ) where for each t > 0 the nonlinear map St maps bounded sets in
Z = Hul 1 into bounded sets in H 2 . Moreover, for all α ∈ R there is a
ul
constant Δ1 (α) such that
lim sup u(t)H 1 ≤ Δ1 (α).
ul
t→∞
More precisely, there is a constant C independent of α ∈ [0, 1] such that
Δ1 (α) ≤ Cα1/4 .
Proof. The local existence and uniqueness of solutions and continuous de-
pendence on the initial condition in Hul 1 follow as in §10.2.2. We consider
≤ (4b2 + α)u2ρb − ρb u4 dx ≤ c1 − c2 u2ρb ,
R
d2
with c1 = 2 R ρb dx and c2 = d − 4b2 − α, and where b ≤ 1 was used and
d ≥ 0 is arbitrary. Assuming c2 > 0 and applying Gronwall’s Lemma to the
differential inequality yields
c1
u(t)2ρb ≤ e−2c2 t u(0)2ρb + (1 − e−2c2 t ).
c2
Letting t → ∞ and choosing the optimal d gives
lim sup u(t)ρb ≤ C α + b
2 2
ρb dx = Δ0 (α, b).
t→∞ R
With R ρb dx = C/b we obtain the desired estimate for α ∈ [0, 1] when
optimizing w.r.t. b = O(α1/2 ) ∈ (0, 1].
The estimate in Hul 1 is derived via the variation of constant formula
t
u(t) = e (t−τ )L
u(τ ) + e(t−r)L (αu(r) − u(r)3 )dr.
τ
Using the smoothing properties (10.111) and u2L∞ ≤ CuL2 uH 1 the
ul ul
nonlinear terms can be estimated as
erL u3 H 1 ≤erL L2 →H 1 u3 L2
ul ul ul ul
−1/4
≤Cr uL2 u2L∞ ≤ Cr−1/4 u2L2 uH 1 .
ul ul ul
For s ∈ {0, 1} we set es (t) = u(t)Huls and find the integral inequality
@ t
−1/4
(10.112) e1 (t) ≤C (t − τ ) e0 (τ ) + α (t − r)−1/4 e0 (r) dr
τ
t A
+ (t − r)−1/4 e0 (r)2 e1 (r) dr ,
τ
we have Iδ ≤ 2CE(1 + α)δ 3/4 . Inserting this result into (10.112) yields
@ A
e1 (τ + δ) ≤ C Eδ −1/4 + αEδ 3/4 + E 2 δ 5/8 Iδ .
This shows that u(t) is bounded in Hul1 and that the bound only depends on
the parameters and the L2ul bound E. For small E > 0 we have e1 ≤ CE.
Hence, solutions exist globally and for t → ∞ the bound E can be replaced
by Δ0 .
Theorem 10.8.3. The SH equation possesses an absorbing set, and a global
1 , H 1 )-attractor Aε in the sense of Definition 8.3.18.
(Hul ρ
is denoted with St and for (10.114) in Y = Hul with GT . To deal with the
ε 1
(10.118) Φδ (STε1 /δ2 (u0 )) − GT1 (A0 )Y ≤Cδ 1/4 .
Thus, we relax the regularity requirements for A0 , and we allow for an
initial error u0 − ψδ (A0 ). These two improvements allow to connect the
approximation result with the second result, which is the attractivity of the
GL set, i.e., of the set of all functions u ∈ Z having in leading order the
form of ψδ (A) for some appropriate A ∈ Y .
Theorem 10.8.5. (Attractivity) For each r0 > 0 there exists constants
C, T0 , R1 , δ0 > 0 such that for all 0 < ε ≤ δ ≤ δ0 we have
(10.119) distZ (STε0 /δ2 (BZ (δr0 )), ψδ (BY (R1 ))) ≤Cδ 5/4 ,
(10.120) distY (Φδ (STε0 /δ2 (BZ (δr0 ))), BY (R1 )) ≤Cδ 1/4 ,
where BZ (r) = {u ∈ Z : uZ ≤ r} and distZ (A, B) = supa∈A inf b∈B a −
bZ .
Both theorems are slight generalizations (due to the additional scaling
parameter δ ≥ ε) of [Sch94b, Lemmas 10-12] and have first been formulated
in [MS95, Theorem 4.3 and Theorem 4.2]. We review their proofs in §10.8.3.
As a first consequence of the two results and of the existence of the globally
attracting set for the GL equation, cf. Corollary 8.3.25, we have the global
existence and uniqueness of solutions in a neighborhood of the unstable
origin of the pattern forming system.
Theorem 10.8.6. (Global existence and uniqueness) There exist T0 ,
T1 , δ0 > 0 such that for all 0 ≤ ε ≤ δ ≤ δ0 and R0 sufficiently large we have
S(T0 +T1 )/δ2 (BZ (δR0 )) ⊂ BZ (δR0 ). Therefore, solutions St (u0 ) with initial
conditions u0 ∈ BZ (δR0 ) stay bounded and exist globally in time.
Remark 10.8.7. For the SH equation this is not a new result. However,
for instance for the 3D Couette-Taylor problem, where no weighted a priori
estimates are available, Theorem 10.8.6 gives a nontrivial global existence
and uniqueness result, cf. [Sch99b]. For pattern forming systems global
existence in the sense of Theorem 10.8.6 can be shown, whenever analogs to
Theorem 10.8.4, Theorem 10.8.5, and Corollary 8.3.25 can be established.
See Figure 10.14.
Moreover, it is possible to show that all solutions u(t) = Stε (u0 ) can be
shadowed by the lift of a pseudo-orbit in the GL equation. Here the notion
of pseudo-orbits is similar to but slightly different from the one in §2.4.3. A
(T1 , κ)-pseudo-orbit is pieced together from true orbits of time span T1 with
jumps of maximal size κ in between:
10.8. Attractors for pattern forming systems 385
Original system
GL equation
b2) T = 0 c2) T = T0
As already said, this will allow us to prove that the diameter of the
attractor of the SH equation in Hul1 is not only of size O(ε1/2 ), but of size
(a) initial condition and its Fourier modes (b) jumps in u and in A.
1
0.8
u0(x) |ût=0 |1/4
0.2 0.2
0.5 2A ( εx) 0.6
0 0.15
0.15
0.4 ||u−εψA||∞ ||A− φδ(u)||∞
0 0.1 0.1
0.2 0.05 0.05
−0.5
0
0 50 100 150 200 250 300 0 2 4 0 50 100 20 40 60 80
Again we work on x ∈ (0, l), l = 100π, X = εx ∈ (0, L), L = εl, and choose
the initial condition (10.123), but now we choose the smaller ε = 0.25.
Moreover, we reset A to Φε (u) if eu,∞ (t) > 2ε5/4 . The results are given in
Figure 10.16.
(a) initial condition and its Fourier modes (b) jumps in u and in A.
0.8 0.6
0.4 u (x)
0 |ût=0 |1/4
0.3
0.5
2A (ε x) 0.6
0.2 0
0.4
||u−εψA||∞
0.2 ||A−φ (u)||
0 0.4 0.3 δ ∞
0.2
−0.2 0.2 0.1
0.1
−0.4
0 100 200 300 0 2 4 50 100 150 200 50 100 150 200
0 50 100 150 0 2 4 0 20 40 60 0 20 40 60
t= 178 t= 178 t= 178 t= 178
u
ε ψ (A) 0.6
δ |û(k)| 1/4
0.5 1.5 1.4
|A−φ (u)|
0.4 δ
1 1.3 |A|
0
1.2 |Φ (u)|
δ 0.2
0.5
−0.5 1.1
0 50 100 150 0 2 4 0 20 40 60 0 20 40 60
t= 250 t= 250 t= 250 t= 250
1.52 0.05
|û(k)|1/4 |A| |A−φδ(u)|
0.5 1.5 |Φ (u)|
δ
u 1.5 0.04
0 ε ψδ(A) 1
1.48 0.03
0.5
−0.5
1.46
0.02
0 5 10 0 2 4 0 20 40 60 0 20 40 60
(t > 250, say), these O(ε ) terms can be seen more clearly in (b)
2
10.8.3. Ideas of the proofs. We want to prove that the set of initial
conditions for which the GL equation makes correct predictions is absorbing
in the sense of Theorem 10.8.5. In other words, given an arbitrary initial
conditions u0 ∈ BZ (δr0 ), we want to show that the associated solution
t → u(t) of the SH equation (10.109) near the Turing instability develops
in such a way that there exist a time
t and functions B = B(X) ∈ Hul 1 and
u(x,
t) = εB(εx) + c.c. + ε3 R(x).
Then we can use B as initial condition for the solution A ∈ C([0, T0 ], Hul
1 ) of
we find
S1/ε θs1 H 1 ≤ Cε−1 E1h u0 H 1 ≤ CR0 = O(1)
ul ul
and similarly
T1 /ε2
2 −τ )Λ
S1/ε θs2 H 1 ≤ S1/ε θE1h e(T1 /ε H 1 →H 1 dτ
ul ul ul
0
× sup ε−1 E1h (u3 (τ ))H 1
ul
t∈[0,T1 /ε2 ]
for ε → 0.
N
(10.127) u|t=O(1/ε2 ) = εβ(m) Am (εx)eimx + O(εN +1 )
m=0
and this turns out to be sufficient to prove (10.117). With similar estimates
we can prove
S1/ε θE1h sres (T0 /ε2 )H 1 ≤ C = O(1),
ul
which in the end implies (10.118).
2c3 α + ν 3/2
(10.128) A(T )L∞ ≤ (1 − e−νT ) + e−νT A0 L∞ .
ν 3c3
Moreover, for each τ ∈ [0, T ) we have
1 M (T −τ )
√
(10.129) A(T )H 1 ≤ C 1 + √ + Me T − τ A(τ )L∞ ,
ul T −τ
where M = α + 3c3 max{ α /c3 , A(τ )L∞ } and C is a universal constant
, c3 , and A.
independent of α
Proof. Let A(T, X) = r(T, X)eiφ(T,X) , which for r yields the equation
2
∂ T r = c2 ∂ X r − (∂X φ)2 r + α
r − c3 r 3 .
Using the maximum principle it is possible to compare r with the solution
a−c3 a3 . Hence, if A0 L∞ = a(0), then A(T )L∞ ≤ a(T ) for all
of ∂T a = α
392 10. Dynamics of pattern and the GL equation
since fδ (K) = 0 for |K| ≤ 1/(6δ). These estimates give the desired result.
Re λ
ε2
−kc kc k
∂t u = ε2 u − (1 + ∂x2 )2 u + εsu3 − u5
398 10. Dynamics of pattern and the GL equation
Exercises
10.1. Consider the Kuramoto-Shivashinsky-KdV equation
∂t u = −(1 + ∂x2 )2 u + ∂x u + ∂x3 u + ε2 u + ∂x (u2 ),
with x ∈ R, t ≥ 0, 0 < ε 1, and u(x, t) ∈ R. Make an ansatz
u(x, t) =εA1 (ε(x − ct), ε2 t)ei(x−ωt)
ε2
+ ε2 A2 (ε(x − ct), ε2 t)e2i(x−ωt) +
A0 (ε(x − ct), ε2 t) + c.c.
2
and derive equations for A0 , A1 , and A2 . Eliminate A0 and A2 to derive a GL
equation for A1 .
10.2. Compute the higher order GL approximations for the SH equations. How do
the higher order approximations for the abstract system in §10.3 look like?
10.3. Replace the assumption on page 334 on the critical curve of eigenvalues λ1 by
the assumptions that λ1 (kc , 0)=iω0 ∈iR, ∂k λ1 (kc , 0)= − icg , and ∂k2 Reλ1 (kc , 0)<0,
for a wave number k = kc > 0. Derive the GL equation for this situation.
10.4. Apply Theorem 10.4.3 to the KS equation by checking the assumptions (A0)-
θ
(A4) in Hul -spaces.
10.5. Prove that a2 and b2 in (10.106) can be eliminated from the terms of order
O(1) by a normal form transform, respectively averaging.
10.6. Let θ, θ0 ≥ 0 and let g(k) satisfy |g(k)| ≤ C|k − k0 |θ0 . Prove that
−1 (· − k0 ))L2 ≤ Cεθ0 −1/2 A
g(·)ε−1 A(ε L2 .
θ θ+θ 0
10.13. Compute for (10.95)-(10.97) the associated transformations for the pressure.
10.14. Use the symmetries (10.95)-(10.97) to show that in the associated amplitude
equation (10.106) in PR II only four and not eight different coefficients occur. In
order to do so make a GL ansatz in PR II and compute representations of the
transformations τa , S, and Rϕ on the level of the amplitudes.
Chapter 11
401
402 11. Wave packets and the NLS equation
uniform and not very accurate spatial discretization 1012 points, even if we
ignore the transverse directions and the temporal discretization. Hence, a
direct simulation of Maxwell’s equations which describe these electromag-
netic waves is very expensive, if not impossible. Therefore, before making
any numerical investigation, the system has to be analyzed and simpler,
numerically more suitable, models have to be derived.
It turns out that the multiple scaling character of the problem is not
only a curse, but also a blessing, since it allows to separate the dynamics
of the envelope from the dynamics of the carrier wave, such that by mul-
tiple scaling analysis the NLS equation can be derived for the description
of the slow modulations in time and space of the envelope of the spatially
and temporarily oscillating wave packet. Due to the immense reduction of
the dimension of the discretized problem by this procedure the NLS equa-
tion turned out to be a very successful model. Even though arguably its
most important application is in nonlinear optics, e.g., [Agr01], the NLS
equation has also been derived for water waves [Zak68, Osb10], for waves
in DNA [SH94b] and other discrete chains, for Bose-Einstein condensates
[Pel11], in plasma physics [Deb05, Chapter 10], and in many other fields
as a universal envelope or modulation equation, cf. also [Mil06, Chapter
10]. In this chapter we explain its justification by approximation theorems
for model problems. We explain its universal character, give an overview
about approximation results, and explain some applications.
11.1. Introduction
The Nonlinear Schrödinger (NLS) equation
(11.1) 2
∂T A = iν1 ∂X A + iν2 A |A|2 ,
with T ∈ R, X ∈ R, ν1 , ν2 ∈ R, and A(X, T ) ∈ C is a universal modulation
equation which can be derived via multiple scaling analysis in order to de-
scribe slow modulations in time and space of the envelope of a spatially and
temporarily oscillating wave packet. For instance, for the nonlinear wave
equation
(11.2) ∂t2 u = ∂x2 u − u − u3 , (x ∈ R, t ∈ R, u(x, t) ∈ R),
also called the cubic Klein-Gordon equation, the ansatz for the derivation
of the NLS equation is
(11.3) εψNLS = εA ε(x − cg t), ε2 t ei(k0 x+ω0 t) + c.c.,
where 0 < ε 1 is a small perturbation parameter, where cg is the group
velocity, and where the basic temporal and basic spatial wave number ω0
and k0 are related by the linear dispersion relation ω02 = k02 + 1. We obtain
11.1. Introduction 403
that the envelope A of the underlying carrier wave ei(k0 x+ω0 t) has to satisfy
in lowest order the NLS equation
(11.4) 2iω0 ∂T A = (1 − c2g )∂X
2
A − 3A |A|2 .
The dynamics of the NLS equation has been discussed in §8.1. The pulse
solutions found in §8.1.1 correspond to modulating pulse solutions in the
original system, cf. Figure 11.2.
cg
O(ε)
cp
O(ε−1 )
+ ε3 E3 (−A3 )
+ ε4 E(2cg ∂X ∂T A)
+ ε5 E(−∂T2 A) + c.c..
By choosing ω = ω0 and k = k0 to satisfy the linear dispersion relation
ω 2 = k 2 + 1,
by choosing cg to be the linear group velocity
d k0
cg = ω = ,
dk k=k0 ,ω=ω0 ω0
and by choosing A to satisfy the NLS equation
(11.6) 2iω0 ∂T A = (1 − c2g )∂X
2
A − 3A |A|2 ,
11.2. Justification in case of cubic nonlinearities 405
the first three lines in the residual cancel. However, we still have Res(εψNLS ) =
O(ε3 ).
Formal smallness of the residual. It turns out that by adding higher
order terms to the approximation εψNLS the residual can be made arbi-
trarily small, i.e., for arbitrary, but fixed n ∈ N with n ≥ 3 there exists an
approximation εψn with εψn −εψNLS = O(ε3 ) and Res(εψn ) = O(εn ). Since
εψn − εψNLS = O(ε3 ) the approximation εψn makes the same predictions as
εψNLS about the behavior of the solutions u of the original system. We will
show Res(εψn ) = O(εn ) for n = 4, 5. With these two examples the general
situation can be understood.
In order to obtain
(11.7) Res(εψ4 ) = O(ε4 )
we define
εψ4 = εψNLS + ε3 A3 ε(x − cg t), ε2 t E3 + c.c. .
We find
Res(εψ4 ) = ε3 E3 −A3 − (9ω02 − 9k02 − 1)A3 + O(ε4 ).
Due to the non-resonance 9ω02 − 9k02 − 1 = 9(k02 + 1) − 9k02 − 1 = 8 = 0 we
can choose A3 = −(9ω02 − 9k02 − 1)−1 A3 in order to achieve (11.7). In order
to achieve
(11.8) Res(εψ5 ) = O(ε5 )
we define
εψ5 = εψ4 + ε2 A12 (ε(x − cg t), ε2 t)E + ε4 A32 (ε(x − cg t), ε2 t)E3 + c.c.
where A12 and A32 are new functions to be chosen below. We find
Res(εψ5 ) =ε4 E −2iω0 ∂T A12 + (1 − c2g )∂X
2
A12
+ ε4 E −3A2 A12 − 6 |A|2 A12 − 2cg ∂X ∂T A
+ ε4 E3 ((9ω02 − 9k02 − 1)A32 − 3A2 A12 ) + O(ε5 ) + c.c..
By choosing A12 to satisfy the linearized NLS equation
(11.9) −2iω0 ∂T A12 + (1 − c2g )∂X
2
A12 − 3A2 A12 − 6 |A|2 A12 − 2cg ∂X ∂T A = 0
and A32 to satisfy
(9ω02 − 9k02 − 1)A32 − 3A2 A12 = 0
we achieve (11.8). In order to achieve Res(εψn ) = O(εn ) we choose
(m)
α
εψn = εα(m)+(j−1) Amj (X, T )Em
m=−N,...,N j=1
406 11. Wave packets and the NLS equation
where
s1 = 2ε3 E 3 A3 C 0 ≤ 2ε3 A3C 0 ,
b b
We can use the right-hand side of the NLS equation to estimate ∂T AC 1
b
and ∂T2 AC 0 . For instance we have
b
1
∂T AC 1 ≤ (1 − c2g )∂X
2
AC 1 + 3A3C 1 < ∞
b 2ω0 b b
and
sup εψN LS (t) − εψn (t)C 0 ≤ Cε3 .
b
t∈[0,T0 /ε2 ]
For (11.2) we have supt∈[0,T0 /ε2 ] εψN LS (t)−εψβ (t)C 0 ≤ Cε3 . If we look
b
in more detail at the approximation εψ5 we recognize that most regularity
is lost in (11.9) for A12 . Since ∂T2 A12 (·, T )C 0 has to be estimated, we need
b
A12 (·, T ) ∈ Cb4 , and so we need ∂X ∂T A(·, T ) ∈ Cb4 , respectively A(·, T ) ∈ Cb7 .
Since the structure of the approximation equations is the same for the next
orders we lose three derivatives in each step such that the estimates are
possible with θA = 3(n − 3) + 1. In fact θA can be chosen much smaller by
a number of simple tricks, cf. §11.5.1.
The equations for the error. Estimates for the residual, even in norms,
are only a necessary condition for showing that the NLS equation makes
correct predictions about the behavior of the original systems. By no means
they are sufficient. The errors can sum up in time and there are a number
of counter-examples, cf. [Sch95b, SSZ15], showing that formally derived
modulation equations make wrong predictions about the behavior of the
original system.
The error εβ R = u − εψ, the difference between the solution u and the
approximation εψ = εψn , with β and n suitably chosen, is estimated with
the help of Gronwall’s inequality. It satisfies
(11.11) ∂t2 R = ∂x2 R − R − 3ε2 ψ 2 R − 3εβ+1 ψR2 − ε2β R3 − ε−β Res(εψ).
Although there is local existence and uniqueness for (11.11) in Cbm -spaces
by the method of characteristics, here some crucial differences to Chapter
10 arise, since these spaces and this method are not suitable for obtaining
estimates on the long time scale O(1/ε2 ).
Estimates for the residual in Sobolev spaces. Sobolev spaces turn
out to be more suitable for (11.11). Hence, we assume that
A ∈ C([0, T0 ], H θA ) is a solution of the NLS equation with θA ≥ 0 suffi-
ciently large.
As a first step we have to re-estimate the residual in Sobolev spaces,
taking into account the scaling properties of the L2 -norm. As in §10.2.2 we
find
Res(εψNLS )H θ ≤C ε3 A(ε·)2C θ A(ε·)H θ
b
+ ε4 ∂X ∂T A(ε·)H θ + ε5 ∂T2 A(ε·)H θ
=O ε3 A(ε·)H θ+4 .
408 11. Wave packets and the NLS equation
However,
1/2 1/2
−1/2
A(ε·)L2 = |A(εx)| dx
2
=ε |A(X)| dX
2
= ε−1/2 AL2
R R
such that finally
Res(εψNLS (t))H θ = O ε5/2 AH θ+4 .
and
sup εψN LS (t) − εψn (t)H θ ≤ Cε3/2 .
t∈[0,T0 /ε2 ]
The equations for the error in Fourier space. For (11.2) simple
energy estimates are possible, cf. Exercise 11.1. However, in order to have a
method which also works for more general systems we use semigroup theory
and write the equations for the error as first order system in Fourier space.
We set β = n − 5/2, choose εψ = εψn and find
− 3ε2 ψ∗2 ∗ R
= −ω 2 R
∂t2 R − 3εβ+1 ψ ∗ R
∗3 + ε−β Res(εψ),
∗2 − ε2β R
√
where ω(k) = k 2 + 1. This is conveniently written as a first order system
1 = iω R
∂t R 2 ,
1 + ε2 f,
2 = iω R
∂t R
where
1
∗2 .
f = −3ψ ∗ R − 3εβ−1 ψ ∗ R ∗3 + ε−β−2 Res(εψ)
∗2 − ε2β−2 R
iω
This system is abbreviated in the following as
t) = Λ(k)R(k,
∂t R(k, t) + ε2 F(k, t),
with
0 iω(k) 0
Λ(k) = , F (k, t) = .
iω(k) 0 f (k, t)
11.2. Justification in case of cubic nonlinearities 409
and
1
· − k 1
·
1 = 2
ψ A
0
+ h.o.t. 1 ≤ C A
+ h.o.t.
Lθ
ε ε Lθ ε ε L1θ
≤ CAL1 + h.o.t. ≤ CAH 0 + h.o.t.
θ θ+1
Remark 11.2.5. Note that ψ 0 = O(ε−1/2 ) such that ψ has to be es-
Hθ
timated in the Lθ -norm, respectively ψ in the Cbθ -norm in order to get ε2
1
and as a consequence
sup u(t) − εψN LS (t)H θ < Cε3/2 .
t∈[0,T0 /ε2 ]
with f (x, t) = −u(x, t) − u(x, t)3 which is based on the solution formula for
the inhomogeneous wave equation. For t0 > 0 sufficiently small the right-
hand side F (u) is a contraction in the space C([−t0 , t0 ], H θ+1 ). Thus, there
exists a unique fixed point u∗ = F (u∗ ) which is a classical solution of (11.2)
if m ≥ 2.
11.3. The universality of the NLS equation 411
The solutions exist as long as the norm of the solutions stay bounded.
By using the error estimates as a priori estimates we can guarantee that
the solutions stay bounded for t ∈ [0, T0 /ε2 ] and so we can apply the local
existence and uniqueness again and again to guarantee the existence and
uniqueness of the solutions of the error equations which are obtained from
(11.2) by a smooth change of variables.
where
4(k) = 0 iω(k) (w)(k, 0
M , N t) = −1 ∗3 .
iω(k) 0 (k, t)
iω(k) u
This system is diagonalized forfixed wave number k. For (11.14) the asso-
1 1
ciated transformation S = √12 is independent of k and unitary,
1 −1
i.e., S−1 = S∗ . The transformed variable z = S∗ w
satisfies the diagonalized
system
(11.15) z + S∗ N
∂t z = Λ (S
z ),
with Λ(k) = diag(iω(k), −iω(k)). It turns out that the NLS equation can
be derived whenever the original system can be transformed in a system of
this form.
412 11. Wave packets and the NLS equation
with complex-valued kernels s2,n,n1 ,n2 (k, k−l, l), s3,n,n1 ,n2 ,n3 (k, k−l, l−m, m),
etc. We have for instance
s2,n,n1 ,n2 (k, k−l, l) =
fn∗ (k), e−ikx N2 [fn1 (k−l)ei(k−l)x , fn2 (l)eilx ]L2
where fn∗ (k) is the associated adjoint eigenfunction w.r.t. the scalar product
·, ·L2 .
Derivation of the NLS equation for (11.15). Taking the Fourier
transform of the ansatz in physical space leads to the ansatz
−1 k − k0 2
z(k, t) = εε A1 , ε t eiω(k0 )t eicg (k−k0 )te1
ε
(11.16)
−1 k + k0 2
+εε A−1 , ε t e−iω(k0 )t eicg (k+k0 )te2
ε
for (11.15), where
1 0
e1 = and e2 = .
0 1
11.3. The universality of the NLS equation 413
The notation εε−1 refers to the amplitude scaling ε and the wave number
scaling ε−1 , but in the following we shorten this to εε−1 = 1. Since the
Fourier modes of the wave packet are concentrated in an O(ε) neighborhood
of the basic wave numbers ±k0 the evolution of the wave packet will be
strongly determined by the curves ±ω at ±k0 . At eiω(k0 )t eicg (k−k0 )te1 we
find
1 + iεcg K A
iω(k0 )A 1 + ε2 ∂T A
1
1 + iε∂k ω(k0 )K A
=iω(k0 )A 1 + i ε2 ∂ 2 ω(k0 )K 2 A
1
k
2
3i 1 ∗ A
−1 + O(ε3 ),
+ ε2 A1 ∗ A
4ω(k0 )
where k = k0 + εK, A 1 = A1 (K, T ), and where we used
k − l − k0 iω(k0 )t icg (k−l−k0 )t l − k0
A ,T e e A , T eiω(k0 )t eicg (l−k0 )t dl
ε ε
R
k − 2k 0
=ε A − m, T A(m, T )e2iω(k0 )t eicg (k−2k0 )t dm .
ε
R
At ε0 and ε1 we obtain the linear dispersion relation and the linear group
velocity. At ε2 we obtain a NLS equation. Undoing the transformation
w = Sz and u = w1 gives a multiple of the original approximation
(k, t) = √
u
1
A1 k − k0 , ε2 t eiω(k0 )t eicg (k−k0 )t
2 ε
k + k0 2 −iω(k0 )t icg (k+k0 )t
+A−1 ,ε t e e .
ε
for |n| ≥ 2 and complex valued functions A n,j , where An,−j = An,j in
physical space. With k − k0 = εK we find at ε2 for the modes concentrated
at k0 that
1,1 (K, T )
∂T A
=i∂k2 ω1 (k0 )K 2 A1,1 (K, T )/2
∞
+2 s2,1,1,n (k0 , k0 , 0) A n,0 (κ, T ) dκ
1,1 (K − κ, T )A
n∈Z −∞
∞
+2 s2,1,1,n (k0 , −k0 , 2k0 ) A n,2 (κ, T ) dκ
1,−1 (K − κ, T )A
n∈Z −∞
∞ ∞
+ 3s3,1,1,1,1 (k0 , k0 , k0 , −k0 ) 1,1 (K − κ1 , T )
A
−∞ −∞
1,1 (κ1 − κ2 , T )A
×A 1,−1 (κ2 , T ) dκ2 dκ1 ,
11.3. The universality of the NLS equation 415
such that
∂t v = Av − K (u)NQ (u) = −iΔu − ui∇u · ∇u
In order to obtain an evolutionary problem for v we have to invert the
transformation v = u − K(u). For small v this is possible by applying the
implicit function theorem.
We apply this idea to the equation for the error and eliminate the O(ε)-
term 2εψ ∗ R
1 with a normal form transformation. In order to do so we first
diagonalize (11.25) by introducing
R
ψ 1 1 1
R=S −1 1
, Ψ=S −1
, where S = √ .
R2 1 2 1 −1
iω(k) ∂t ψ(k))
We find
(11.27) ∂t R = ΛR + 2εB(Ψ, R) + εβ B(R, R) + ε−β RES(εΨ),
with Λ a symmetric linear map and B(·, ·) a bilinear map, which are given
in Fourier space by
iω 0 −1 0
Λ = , RES(εΨ) = S ,
0 −iω Res(εψ)
U, U U
) = 1 S −1 B(S , SU
), B(
U
, V ) = 0
B( iω 1 ∗ V1 ,
U
= (U
where U 2 ). Then we make a near identity change of variables
1 , U
(11.28) w = R + εQ(Ψ, R)
with Q an autonomous bilinear map. This gives
∂t w =∂t R + εQ(∂t Ψ, R) + εQ(Ψ, ∂t R)
=ΛR + 2εB(Ψ, R) + εβ B(R, R) + ε−β RES(εΨ) + εQ(∂t Ψ, R)
+ εQ(Ψ, ΛR + 2εB(Ψ, R) + εβ B(R, R) + ε−β RES(εΨ))
=Λw − εΛQ(Ψ, R) + 2εB(Ψ, R) + εβ B(R, R) + ε−β RES(εΨ)
+ εQ(∂t Ψ, R) + εQ(Ψ, ΛR + 2εB(Ψ, R) + εβ B(R, R) + ε−β RES(εΨ)),
and so
∂t w = Λw + ε (−ΛQ(Ψ, R) + Q(∂t Ψ, R)
(11.29)
+Q(Ψ, ΛR) + 2B(Ψ, R)) + O(ε2 ).
In order to eliminate the dangerous term 2εB(Ψ, w) we have to find a bilinear
Q such that
(11.30) −ΛQ(Ψ, R) + Q(∂t Ψ, R) + Q(Ψ, Λw) + 2B(Ψ, R) = 0.
11.4. Quadratic nonlinearities 419
In this form this equation is hard to analyze. For its simplification we first
use that
√ εA1 ε(x − cg t), ε2 t ei(k0 x+ω0 t) + O(ε2 )
ε∂t Ψ = 2∂t
εA−1 ε(x − cg t), ε2 t e−i(k0 x+ω0 t) + O(ε2 )
√ iω0 εA1 ε(x − cg t), ε2 t ei(k0 x+ω0 t) + ε2 G1
= 2
−iω0 εA−1 ε(x − cg t), ε2 t e−i(k0 x+ω0 t) + ε2 G−1
1 = O(1). Hence, (11.30) is given in lowest order by
with G L θ
with bjmn = bjmn (k, k − l, l) some smooth kernel. Thus, we make the same
ansatz for the j-th component of Q, namely
(Q(ψ, R))j = qmn
j
(k, k − l, l)ψm (k − l)R
n (l) dl,
m,n=1,2 R
j j
with qmn = qmn (k, k − l, l) some kernel which we have to compute. Inserting
these representations in (11.31) yields the relations
(11.32) i(ωj (k) − ω1 (k0 ) − ωn (l))j
q1n (k, k − l, l) = 2bj1n (k, k − l, l),
and
(11.33) i(ωj (k) − ω2 (−k0 ) − ωn (l))j
q2n (k, k − l, l) = 2bj2n (k, k − l, l),
with ω1,2 (k) = ±ω(k).
Example 11.4.3. For system (11.26), where the approximation is replaced
by a general function and where only one curve of eigenvalues is involved,
the associated relation would be
q(k, k − l, l) = b(k, k − l, l),
i(ω(k) − ω(k − l) − ω(l))
with ω(k) = −k 2 and b(k, k − l, l) = i(k − l)l. We find
ω(k) − ω(k − l) − ω(l) = −k 2 + (k − l)2 + l2 = −2(k − l)l
such that q(k, k − l, l) = − 12 .
and
(11.35) inf inf |(ωj (k) − ω2 (−k0 ) − ωn (l))| ≥ C > 0
j,n∈{1,2} k,l∈R,|k−l|≤δ
for this δ > 0 fixed. The validity of the non-resonance conditions can be
checked graphically by looking for intersections of the curves k → ±ω(k)
and k → ω(k0 ) ± ω(k − k0 ), see Figure 11.4. Since the asymptotes k → ±k
and k → ω(k0 ) ± (k − k0 ) to these curves are separated, for no value of k0 a
quadratic resonance occurs.
Since
sup sup |bjmn (k, k − l, l)| ≤ C < ∞,
j,m,n∈{1,2} k,l∈R,|k−l|≤δ
(11.34) immediately implies
sup sup |j
qmn (k, k − l, l)| ≤ C < ∞.
j,m,n∈{1,2} k,l∈R,|k−l|≤δ
As a consequence we obtain
1 R
(Q(ψ, R))H θ ≤ CψC θ RH θ ≤ Cψ 2.
b
L L θ θ
Thus, the transformation (11.28) can be inverted with the help of Neumann’s
series for ε > 0 sufficiently small. We denote the inverse by R = Rε (w) =
O(1). Therefore, (11.29) transforms into
(11.36) ∂t w = Λw + ε2 F
11.5. Extension of the theory 421
with
ε2 F =εβ B(Rε (w), Rε (w)) + ε−β RES(εΨ) + ε2 Q(G, Rε (w))
+ εQ(Ψ, 2εB(Ψ, Rε (w)) + εβ B(Rε (w), Rε (w)) + ε−β RES(εΨ)).
Since ε2 F obeys the same estimates as the one in Lemma 11.2.4 the rest of
the proof of the approximation property from §11.2 applies line for line to
(11.36).
Proof. This follows immediately from the fact that the left-hand side of this
inequality can be estimated by
ε
≤ sup g(k)(1+ ) − max(θ+θ∞ ,θ0 ) θ −1 ·) 2
(1+|k−k0 |) ε−1 A(ε
k∈R |k−k0 | Lmax(θ+θ ,θ )
∞ 0
|εK| θ 0 (1+|k0 +εK|) θ+θ ∞
≤ sup C min( ,
) ε−1/2 A(·)
θ θ+θ ∞ L2max(θ+θ ,θ )
K∈R (1+|K|) 0 (1+|K|) ∞ 0
where the loss of ε−1/2 is due to the scaling properties of the L2 -norm.
422 11. Wave packets and the NLS equation
for g(k) = 1 − χ(k). Hence, the modified approximation and the original
approximation are O(εθ−1/2 ) close to each other in H θ if the ansatz functions
A are in H θ . Moreover, we have
Lemma 11.5.3. Let θ, θ0 ≥ 0, θ∞ ∈ R, and let g(k) satisfy
|g1 (k)| ≤ C min(|k − k0 |θ0 , (1 + |k|)θ∞ ).
Then
−1 (· − k0 )) 2 ≤ Cεθ0 −1/2 A
g1 (·)χ(·)ε−1 A(ε 2 .
L θ L θ0
Before we explain this in more detail we derive the NLS equation for the
Boussinesq model (11.40). As before we make the ansatz
εψNLSbouss =εA1 ε(x − cg t), ε2 t ei(k0 x+ω0 t) + c.c.
(11.42) + ε2 A2 ε(x − cg t), ε2 t e2i(k0 x+ω0 t) + c.c.
+ ε2 A0 ε(x − cg t), ε2 t .
can be solved w.r.t. A0 since (1 − c2g )2 = 0 and w.r.t. A2 since −4ω02 + 4k02 +
16ω02 k02 = 0. Inserting the solutions for A0 and A2 into the equation for A1
finally yields the NLS equation
(11.43) 2iω0 (1 + k02 )∂T A1 = (1 − c2g − ω02 )∂X
2
A1 + γA1 |A1 |2 ,
with
4k02 k04
γ= + .
1 − c2g −ω02 + k02 + 4ω02 k02
The error εβ R = u − εΨ satisfies
∂t2 R = ∂x2 R + ∂x2 ∂t2 R + 2ε∂x2 (ΨR) + O(ε2 ).
Writing this as a first order system gives two equations of the form
− m, t)R
∂t Rj (k, t) = iωj (k)Rj (k, t) + ερj (k) ψ(k j (m, t) dm + O(ε2 ),
R
with ωj (0) = ρj (0) = 0, but non-vanishing ωj (0) and ρ j (0). The approxi-
mation εψ is concentrated at the wave numbers ±k0 . Hence, a subsystem
is given by
j (k0 , t) =iωj (k0 )R
∂t R 0 , t)R
j (k0 , t) + ερj (k0 ) ψ(k j (0, t) + h.o.t.,
∂t R
j (0, t) + ερj (0)ψ(−k
j (0, t) =iωj (0)R
0 , t)Rj (k0 , t) + h.o.t..
∂t R
j (0, t) + ρj (0)ψ(−k
j (0, t) =iωj (0)R
0 , t)Rj (k0 , t) + h.o.t..
426 11. Wave packets and the NLS equation
Hence, in the first equation the nonlinear terms do not make problems any-
more. However, in the second equation we have now terms of order O(1)
which can be eliminated in the full system with the argument from above.
The terms of order O(ε) resulting from this transformation in the second
equation can either be eliminated by another transformation or are of long
wave form, i.e., of a similar form as ε∂x (B(εx)R(x)) in physical space. Such
terms can be estimated by energy estimates to have an O(ε2 ) influence on
the dynamics, cf. Chapter 12.
This can be made rigorous for the full system by making the ansatz
= u − εΨ
εβ ϑ(k)R
for the error function R, where ϑ(k) = min(ε+|k|/δ, 1) with δ > 0 sufficiently
small, but independent of 0 < ε 1. This has been carried out in [Sch98a]
with a correction explained in [DS06].
The above idea has been transferred to the water wave problem without
surface tension in case of finite depth in [DSW16]. However, the water
wave problem is a quasilinear problem. Below we will explain the additional
difficulties occurring for such systems. But also the so called FPU system
falls into this class, cf. Exercise 11.4. In [Sch10] it has been explained that
the proofs of the approximation theorems given for the PDE systems can be
transferred almost line for line to the FPU system by looking at the Fourier
transformed FPU system.
with
−1 |k−mk |
(11.44) 1/w(k) = sup |e−αε 0
|.
m∈Z
Hence, the Fourier modes associated to the resonant wave numbers are ex-
ponentially small initially, i.e., of order O(exp(−rε−1 )) for an r > 0, in-
dependent of 0 < ε 1. The quadratic resonances will lead to growth
rates O(exp(εt)) for solutions of order O(ε). Hence, it takes a time of order
O(1/ε2 ) to have O(exp(−rε−1 ))O(exp(εt)) = O(1), i.e., it takes the time
scale of the NLS equation for the error to grow to the order of the NLS ap-
proximation. This idea can be used to prove error estimates on an O(1/ε2 )
time scale for the validity of the NLS approximation also in case of additional
non-trivial resonances, if the solutions of the NLS equation are analytic in
a strip in the complex plane, and if the set of wave numbers resonant to k0
is separated from the set of integer multiples of the basic wave number k0 .
This idea can be made rigorous by making the coefficient α time-dependent,
i.e., by choosing α(t) = α0 − β̌ε2 t in (11.44). This idea has been explained
in [Sch98c] and carried out in [DHSZ16].
The second approach is based on a more detailed analysis of the reso-
nances. Consider a basic wave number k1 = k0 , resonant to wave numbers
k2 and k3 . The ansatz
u(x, t) = εA1 (εt)ei(k1 x+ω1 t) ϕ1 + εA2 (εt)ei(k2 x+ω2 t) ϕ2 + εA3 (εt)ei(k3 x+ω3 t) ϕ3 ,
with vectors ϕj , then yields a three wave interaction (TWI) system
∂T A1 = iγ1 A2 A3 , ∂T A2 = iγ2 A1 A3 , ∂T A3 = iγ3 A1 A2 ,
with coefficients γj ∈ R, associated to the resonances, cf. [Sch05, §3.3]. In
[Sch05, Theorem 3.8] a NLS approximation theorem has been shown in case
that the subspace {A2 = A3 = 0} associated to the wave number k1 = k0 is
stable in the TWI system, cf. Figure 11.6. The proof is based on a mixture
of normal form transforms for the non-resonant wave numbers and energy
estimates for the resonant wave numbers. In [DS06] the ideas of [Sch98a]
and [Sch05] are brought together to handle Boussinesq equations which
model the water wave problem in case of small positive surface tension.
There is also a counter-example [Sch05, §4.1] showing that the NLS
equation fails to approximate solutions in the original system in case of an
unstable k0 -subspace in the associated TWI system and periodic boundary
conditions in the original system. This idea has been carried out in [SSZ15]
for the water wave problem with surface tension and periodic boundary
conditions showing that there exists a continuum of wave numbers and val-
ues of surface tension where the NLS approximation does not make correct
predictions.
428 11. Wave packets and the NLS equation
ImA2
ImA1
ImA3
Figure 11.6. The phase portrait of the TWI system in the invariant
subspace Re A1 = Re A2 = Re A3 = 0. The energy surface is an ellipsoid
since due to conservation of energy not all j have the same sign. The
axes are invariant subspaces associated to the wave numbers kj . There
are one unstable and two stable subspaces.
The situation on the whole real line for an unstable resonance is still
open. In this case the different group velocities ω (kj ) at the resonant
wave numbers kj no longer can be neglected. For a thorough discussion
see [Sch05, §4.2]. A recent attempt to understand this situation can be
found in [MN13].
that the original quasilinear system can no longer be handled after applying
the transformation (see §11.4). This can be seen as follows. Writing the er-
ror equations as first order system gives nonlinear terms with growth rates
proportional to |k| for |k| → ∞. This results in the normal form transform
in a nominator proportional to |k| and in a denominator proportional to
ω(k) − ω(k0 ) − ω(k − k0 ) = O(1)
for |k| → ∞. Thus, the normal form transform is of the form identity plus
a term which is small but loses one derivative. Hence, the normal form
transform can no longer be inverted with Neumann’s series.
So far there are only a few quasilinear systems where approximation
results for the NLS approximation could have been established. One example
is where the right-hand side of the quasilinear
√ dispersive wave system only
loses half a derivative, i.e., a factor of k in Fourier space, as a result of the
normal form transformation. In this case the elimination of the quadratic
terms is still possible and the transformed system can be handled with the
Cauchy-Kowalevskaya theorem [SW11]. The Lagrangian formulation of
the water wave problem in case of finite depth and zero surface tension
falls into this class, cf. [DSW16]. In case of zero surface tension and
infinite depth such a result has been established in [TW12, Tot15] by
finding a special transformation which allows to eliminate all quadratic terms
for this particular system without loss of regularity. Another example is
in the context of the KdV equation where the result can be obtained by
applying the Miura transformation [Sch11], cf. Exercise 11.5. In [CS13]
numerical evidence is given that the NLS approximation is valid for (11.45).
Very recently it turned out that the solutions of the transformed quasilinear
system can be estimated by more clever energy estimates [Dül16, CW16,
DH16].
Although relevant systems can be handled with the last approach a va-
lidity theory for general dispersive wave systems with quasilinear quadratic
terms is still an open problem.
b) follows from
2π
χ2 2
(L()w, w)L2 (χ−1 dx) = |(∂x + i)w|2 + |w| dx > 0, (w = 0).
1
0 χ1
2 (R/2πZ) →
c) follows from the fact that for fixed the operator L() : Hper
L2per (R/2πZ) is elliptic, i.e., L() has a compact resolvent and so L() has
discrete spectrum with ∞ the only accumulation point. Due to the self-
adjointness all eigenvalues are real and semi-simple. Due to the positive
definiteness all eigenvalues are non-negative.
Lemma 11.6.2. Except of intersection points the curves of eigenvalues →
ωn () of L() are smooth.
Figure 11.7. The left panel shows the multivalued inverse of the map
λ → D(λ) and the right panel the associated curves of eigenvalues k →
ω 2 (k) for Example 11.6.3.
11.6. Pulse dynamics in photonic crystals 433
For frequencies ω with ω 2 = λ which fall into a spectral gap the incoming
wave is damped in the photonic crystal with some exponential rate w.r.t.
the depth of penetration x. Hence, photonic crystals can be used as filters.
This is one of the reasons why the wings of a butterfly show their colorful
appearance.
Remark 11.6.4. There is a relation between the regularity of the coeffi-
cients χ1 and χ2 on one side and the size of the spectral gaps in
{ωn (l) : l ∈ (−1/2, 1/2], n ∈ Z \ {0}} ⊂ R
on the other side. For continuous or even smoother χ1 and χ2 the spectral
gaps become smaller as n increases. According to [Eas73, Nti76], the size
of the gaps which are found at ω ∼ n decays at least with 1/nθ+1 for n → ∞
if χ1 ∈ Cbθ and χ2 ∈ Cbθ−2 , i.e., the more regular χ1 and χ2 are, the faster
the gaps close.
Remark 11.6.5. Spectral gaps can be obtained from a spatially homo-
geneous situation by adding small spatially periodic perturbations to the
coefficients. As an example consider the eigenvalue problem
(∂x + i)2 w(x) + ω 2 (1 + 2ε cos(2x))w(x) = 0,
with w(x) = w(x + 2π), ∈ [− 12 , 12 ), and 0 ≤ ε 1. For ε = 0 the problem
is given by
(∂x + i)2 w + ω 2 w = 0,
and can be solved by w(x) = einx with n ∈ Z and associated eigenvalues
ωn2 () = (n + )2 . Hence, at (, ω) = (0, 1) there is a crossing of the curves
of eigenvalues. All single eigenvalues ω vary smoothly w.r.t. small ε since
2ε cos(2x)· is a small perturbation of the operator (∂x + i)2 . Hence, the
smooth curves → ωn () will only vary slightly w.r.t. ε. However, at the
crossing points the curves can split. As an example we consider the point
(, ω) = (0, 1). We use 2 cos 2x = e2ix + e−2ix , make the ansatz w(x) =
n∈2Z+1 cn e 2 . We obtain
inx , and set ω 2 = 1 + ω
which are linear w.r.t. c1 and c−1 . Inserting this into (11.52)-(11.53) gives
2 )c±1 + ε(1 + ω
0 =(∓2 − 2 + ω 2 )c∓1 + εO(|| + |
ω |2 + |ε|)O(|c1 | + |c−1 |),
or equivalently,
−2 − 2 + ω 2 + h.o.t. (1 + ω 2 )ε + h.o.t. c1 0
= .
(1 + ω 2 )ε + h.o.t. 2 − 2 + ω 2 + h.o.t. c−1 0
In order to have non-trivial solutions we need a vanishing determinant. We
find
ω 2 )2 − ε2 + h.o.t. = 0,
(
i.e., (ω 2 )1/2 = ±ε+h.o.t., i.e., there is a splitting of the eigenvalues. In higher
space dimensions spectral gaps cannot be obtained by a small perturbation
of the spatially homogeneous situation. For examples of spectral gaps in 2D
and 3D see [Kuc93, BFL+ 07].
11.6.3. Bloch transform. To derive and justify the NLS equation for
(11.48) we follow [BSTU06] and adapt the Fourier space approach for the
constant coefficient problem (11.2) from §11.2 to Bloch space. For Schwartz
functions u ∈ S, the Bloch transform is defined by
(11.55) (, x) = (T u)(, x) =
u ( + j),
eijx u
j∈Z
By construction we have
(11.57) (, x) = u
u (, x + 2π) and u
(, x) = u
( + 1, x)eix .
The Bloch transform turns out to be an isomorphism between H θ (R, C) and
L2 ((−1/2, 1/2], H θ ([0, 2π), C)), cf. [RS75b, Sca99], where
1/2
1/2
uL2 ((−1/2,1/2],H θ ([0,2π),C)) =
u(, ·)2H θ [0,2π] d .
−1/2
(11.60)
j1 ( − 1 , t)
×u uj2 (1 − 2 , t)
uj3 (2 , t) d2 d1 ,
where
bjj1 j2 j3 (, − 1 , 1 − 2 , 2 )
=
fj (, ·), χ3 (·)fj1 ( − j1 , ·)fj2 (j1 − j2 , ·)fj3 (j2 −, ·)L2 (χ−1 dx) .
1
Since the nonlinear terms have some convolution structure we have a system
as in §11.3. Thus, we can proceed exactly as in §11.3 in order to derive a
NLS equation. However, due to the special structure of (11.60) we keep the
second order system and proceed as in §11.2 and make the ansatz
−1 · − 0 2 −1 · + 0 2
n0 (, t) = εε A1
u 1
, ε t E + εε A−1 , ε t E−1
ε ε
for a n0 ∈ N where Ej = ejiωn0 (0 )t eiωn0 (0 )(−j0 )t . In physical space the
ansatz corresponds to
(11.61) u(x, t) = εψA (x, t) = εA(ε(x+cg t), ε2 t)fn0 (0 , x)ei0 x eiωn0 (0 )t +c.c.,
436 11. Wave packets and the NLS equation
×A 1 (κ1 − κ2 )A
1 (κ − κ1 )A −1 (κ2 ) dκ2 dκ1
∞ ∞
→ bnn00 n0 n0 (0 , 0, 20 , −0 ) A 1 (κ1 − κ2 )A
1 (κ − κ1 )A −1 (κ2 ) dκ2 dκ1
−∞ −∞
and the symmetry of the kernel. The derivation from (11.62) is consistent
with the derivation from the associated first order system since for instance
−(λ n0 (0 ) − 2(ωn 0 (0 ))2 )/(4iωn0 ) = − ((ωn0 (0 )2 ) − 2(ωn 0 (0 ))2 )/(4iωn0 )
than 1 to a function on the complete real axis with period 1. Then we make
the modified ansatz
1 · , ε2 t E1 )() + εε−1 P(χ(·)A
1 (, t) = εε−1 P(χ(·)A
u −1 · , ε2 t E−1 )()
ε ε
and find the NLS equation
(11.64) 2iω1 (0)∂T A 1 = −(λ 1 (0) − 2(ω1 (0))2 )κ2 A 1 ∗ A
1 /2 + γ A 1 ∗ A
−1 ,
∂t2 u ∂x )
⊥ (, x, t) = − L(, u⊥ (, x, t) + χ3 (x)
u3 (, x)
− Ec ()
f1 (, ·), χ3 (·)
u3 (, ·, t)L2 (χ−1 dx) f1 (, x),
1
where Ec ∈ C0∞ , with Ec () ∈ [0, 1], with Ec () = 1 for ∈ [−1/5, 1/5], and
χ() = 0 for || ≥ 2/5. Note that χ is defined on the Fourier wave numbers,
where Ec is defined on the Bloch wave numbers.
We add higher order terms to the ansatz to make the residual smaller,
i.e., we consider
1 (, t) =εε−1 P(χ(·)A
u 1 · , ε2 t E1 )() + εε−1 P(χ(·)A −1 · , ε2 t E−1 )()
ε
ε
3 −1 · 2 3
+ ε ε P(χ(·)A3 , ε t E )()
ε ·
+ ε ε P(χ(·)A−3 , ε3 t E−3 )(),
3 −1
ε
⊥ 3 −1 ⊥
(, x) =ε ε u
u 1 ( , x, ε t)E + ε3 ε−1 u
2 1
⊥ 2
−1 ( , x, ε t)E
−1
ε ε
3 −1 ⊥
+ ε3 ε−1 u⊥ 2 3
3 ( , x, ε t)E + ε ε u −3 ( , x, ε2 t)E−3 .
ε ε
As before we find A 1 as a solution of the NLS equation (11.64) and A −1 as
a solution of the complex conjugate equation. Moreover we choose
−9ω 2 (0)A 3 (κ, T )
1
3 (κ, T ) +
f1 (0, ·), χ3 (·)(f1 (0, ·))3 2 −1 A
= − λ1 (0)A ∗3
L (χ dx) 1 (κ, T ),
1
u⊥
−ω12 (0)1 (, x, t)
438 11. Wave packets and the NLS equation
= − L(εκ, u⊥ ∗2
∂x )1 (κ, x, T ) + 3χ3 (x)(A1 f1 ) ∗ A−1 f−1 )(κ, x)
1 f1 )∗2 ∗ A
−Ec ()
f1 (εκ, ·), 3χ3 (·)(A −1 f−1 )(κ, ·, T ) 2 −1 f1 (εκ, x),
L (χ dx)
1
u⊥
−9ω12 (0)3 (, x, t)
= − L(εκ, u⊥
∂x )3 (κ, x, T ) + χ3 (x)(A1 f1 ∗ A1 f1 ∗ A1 f1 )(κ, x)
1 f1 )∗3 (κ, ·, T ) 2 −1 f1 (εκ, x),
− Ec ()
f1 (εκ, ·), χ3 (·)(A L (χ dx)
1
and A−3 , u
⊥ ⊥
−1 , and u−3 as the solutions of the associated complex conjugate
equations. By this choice we eliminate all nonlinear terms w.r.t. A 1 and
A−1 in the part of the residual belonging to the u equation. This has the
⊥
advantage that we do not need more information about the spectrum and
the associated eigenfunctions, especially we avoid to estimate derivatives of
the eigenfunctions w.r.t. .
In order to solve the equations for A 3 , u
⊥ ⊥
1 , and u3 we need the non-
resonance conditions
−9ω12 (0) = − λ1 (0),
−ω12 (0) ∈spec(−L(εκ, ∂x ))|{f1 (εκ,·)}⊥ ,
−9ω12 (0) ∈spec(−L(εκ, ∂x ))|{f1 (εκ,·)}⊥ .
If these conditions are not satisfied for = εκ for which Ec () = 1 we can
make the support of Ec smaller to satisfy the non-resonance conditions. This
is possible if
∂x ))|{f (0,·)}⊥ ,
−ω12 (0) ∈spec(−L(0, 1
∂x ))|
−9ω12 (0) ∈spec(−L(0, {f1 (0,·)}⊥ .
where
Res(u) = −∂t2 u + χ1 ∂x2 u − χ2 u + χ3 u3 .
The error ε3/2 R = u − εψ satisfies
∂t2 R = χ1 ∂x2 R − χ2 R + 3ε2 χ3 ψ 2 R + 3ε5/2 χ3 ψR2 + ε3 χ3 R3 + ε−3/2 Res(εΨ).
Multiplying
∞ this equation with χ−1
1 ∂t R and performing the integration
−∞ . . . dx yields
∂t E ≤ C1 ε2 E + C2 ε5/2 E 3/2 + C3 ε3 E 2 + C4 ε2 ,
11.6. Pulse dynamics in photonic crystals 439
11.6.5. Gap solitons. Another popular class of models for light propa-
gation in photonic crystals starts directly with a Nonlinear Schrödinger (or
Gross–Pitaevsky) equation in the form
(11.65) iEt = −ΔE + V (x)E + σ|E|2 E, V : R2 → R, x ∈ R2 , t ∈ R,
with E = E(x, t) ∈ C, σ = ±1, and where w.l.o.g. the potential V is 2π–
periodic in each coordinate. Here E is the “out–of–plane” electric field, i.e.,
perpendicular to the 2D periodicity structure, and the amplitude of V is
called the contrast of the material. An interesting problem then is to search
for so called gap solitons E(x, t) = φ(x)e−iωt , where φ ∈ C is a localized
solution of the stationary problem
(11.66) (−Δ + V (x) − ω)φ + σ|φ|2 φ = 0.
Localized means |φ(x)| → 0 exponentially as |x| → 0, and this implies
that ω has to lie in a gap of the essential spectrum of the operator L :=
−Δ + V (x), hence the name “gap soliton”. These have been discussed the
physics literature since the early 1990ties, see, e.g., [Ace00] for a review.
In 1D, where gaps open for small contrasts as discussed above, they are
typically described by so called coupled mode equations, cf. Exercise 11.9. In
440 11. Wave packets and the NLS equation
two and more space dimensions gaps are more difficult to open [Kuc93], and
in particular one needs a finite contrast. Rigorous proofs and asymptotics for
2D gap solitons, based on the reduction of (11.65) to systems of homogeneous
NLS equations, here also called coupled mode equations, can be found for
instance in [DPS09] (for the case of a separable potential) and in [DU09]
(for general V ).
For linearly polarized light, e.g. E = u(x3 , t)ex1 and P = p(x3 , t)ex1 we find
∇ · P = 0, ∇ · E = 0, and so
(11.76) ∂x23 u = ∂t2 u + ∂t2 p,
where u(x3 , t) ∈ R and p(x3 , t) ∈ R.
Modeling the polarization. In order to have an evolutionary problem we
have to close (11.75) or (11.76) by a constitutive law P = P (E) or p = p(u),
respectively. There exist various models. Basically the law of motion of a
particle with coordinates x = x(t) of mass m and charge q is given by
d2 dx
m 2
x = q(E + × B).
dt dt
In the simplest model the polarization is modeled as an oscillator and the
influence of the magnetic field B is neglected. Thus we suppose that for an
atom placed in an electric field the center x = x(t) of the electron density
obeys the equation 2
d 2
m x + ω0 x = qe E,
dt2
with qe the elementary charge, ω0 some normalized temporal wave number,
and m the mass of the electron. On the continuum level in some mean-field
limit for the electron position we find for the polarization
(11.77) ∂t2 P + ω02 P = dE,
with a constant d ∈ R. In case of damping by thermalization we have
2
d d 2
m x + γ x + ω0 x = qe E
dt2 dt
for a γ > 0. On the continuum level we find
(11.78) ∂t2 P + γ∂t P + ω02 P = dE.
n
(11.81) P = Pj ,
j=0
where the dj > 0 are constants taking into account different masses, different
numbers of atoms, etc., of the various kinds of atoms. In order to solve
this system uniquely we need initial conditions for E|t=0 , ∂t E|t=0 , Pj |t=0 and
∂t Pj |t=0 for j = 0, . . . , n. We refer to [SU03b] for a mathematical analysis of
light pulse propagation in glass fibers with this modeling in case of damping,
i.e., γj > 0.
Remark 11.7.2. a) (11.80)-(11.82) models isotropic media. Anisotropic
media can be modeled for instance by choosing γj , ωj2 , dj , and rj as tensors.
b) In the linear case rj = 0 in (11.82), the constitutive laws can be solved
explicitly. For E = E0 eiωt we find P = P0 eiωt + P(t), with P(t) → 0 with
some exponential rate for t → ∞ and
P0 = α(ω)E0
where
n
dj
α(ω) = .
j=0
−ω 2 + iγj ω + ωj2
The numbers γj , dj , and ωj can be used to fit the constitutive law to ex-
perimental data. There is a remarkable good agreement with experimental
observations.
c) More generally, in the linear case there always is a Green’s function
χ1 such that P can be expressed in terms of E, i.e.,
t
(11.83) P (t) = χ1 (t − τ )E(τ ) dτ.
−∞
In isotropic materials χ1 is scalar. In glass fibers the above constants and
also χ1 only depend on the transverse variables, in photonic crystals there
is a periodic dependence on the spatial variables. Although there is no
Green’s function for the nonlinear system, similar to the linear situation the
constitutive law for the polarization is very often modeled by
t t
(11.84) P (t) = χ1 (t − τ )E(τ ) dτ + χ3 (t − τ )|E(τ )|2 E(τ ) dτ,
−∞ −∞
i.e., again (11.81)-(11.82) is replaced by (11.84).
d) In nonlinear optics very often time and space are interchanged. Due
to the finite size of the fibers and due to the experimental data which can be
measured initial conditions are posed at one end of the fiber, namely at x =
0, and one is interested in the solution at the end of the fiber, namely at x =
444 11. Wave packets and the NLS equation
possesses solutions
which yields
−k 2 + ω 2 ω2 uk 0
= .
−d ω02 − ω 2 pk 0
We have non-trivial solutions if the determinant vanishes, i.e., if
(−k 2 + ω 2 )(ω02 − ω 2 ) + dω 2 = 0.
We find four curves of solutions ω = ω1,2,3,4 (k) which are sketched in Figure
11.8. The Fourier transformed system can be written as first order system
and then diagonalized, leading to
∂t
c1 (k, t) =iω1 (k)
c1 (k, t) + nonlinear terms ,
..
.
∂t
c4 (k, t) =iω4 (k)
c4 (k, t) + nonlinear terms ,
such that the Maxwell-Lorentz system falls in the abstract class of systems
considered in §11.3 for which an NLS equation can be derived.
Remark 11.7.3. Other models for the polarization are in use. The choice
∂t2 p = −u − u3 leads to the Klein-Gordon model (11.2) for which we justified
the NLS approximation in §11.2. Hence, the pulse dynamics present in
the NLS equation is present in the models used for the description of the
propagation of light pulses in glass fibers, too. Experimental observations
confirm this approximation and modeling of reality.
11.7. Nonlinear optics 445
where we assumed for simplicity that there are no resonant wave numbers
kj1 , . . . , kj4 with kj1 +. . .+kj4 = 0 and ωj1 +. . .+ωj4 = 0. If this assumption
is not satisfied there will be additional coupling terms Aj2 Aj3 Aj4 in the
equation for A−j1 which however can be handled as explained below, too.
At a first view there seems to be a full coupling between all equations.
However, looking more closely at the coupling terms shows that they have
446 11. Wave packets and the NLS equation
where the Aj satisfy decoupled equations and there are explicit formulae
[CBSU07] for the pulse shifts εψj and the phase shifts εΩj . The internal
dynamics of the wave packets (described via the Aj ) and the interaction
dynamics of the wave packets (described via ψj and Ωj ) can be separated
up to very high order. An almost complete description of the interaction of
general localized NLS described wave packets can be found for the nonlinear
wave equation in [CBCSU08] and in [CBS12] for general dispersive wave
systems. As a consequence the shift of the underlying carrier wave and
the shift of the envelope both can be shown to be of order O(ε) instead of
O(1) w.r.t. the original x-variable. See also [SUW11] for an application to
oscillator chains.
Further Reading. The description of waves in nonlinear optics by mod-
ulation equations is an active field of research. Three main current topics
are:
• ultra-short pulses [SW04, CdR15, PS13, New16], see also Ex-
ercise 11.10;
11.7. Nonlinear optics 447
b) Plug in the ansatz εψv (x, t) = εA(ε(x − ct), ε2 t)ei(kx+ωt) + c.c. into the mKdV
equation and derive a NLS equation for A.
c) Justify this approximation by energy estimates and conclude the validity of an
approximation theorem for the approximation of the KdV equation by the NLS
equation [Sch11].
11.6. The dispersion relation of the water wave problem with surface tension is
given by
ω 2 = (k + σk3 ) tanh(k),
with surface tension parameter σ ≥ 0. Show that for σ ∈ (0, 1/3) beside k = 0 and
k = k0 there are two additional resonances k1 and k2 with k0 + k1 + k2 = 0 and
ω0 + ω1 + ω2 = 0.
11.7. Consider the system for the resonant three wave interaction
Ȧ1 = iγ1 A2 A3 , Ȧ2 = iγ2 A1 A3 , Ȧ3 = iγ3 A1 A2 ,
for Aj (t) ∈ C and coefficients γj ∈ R.
a) Prove that this is a Hamiltonian system with
An An
qn = sgn(γn ) , pn = , H = i(A1 A2 A3 + A1 A2 A3 ).
|γn | |γn |
b) Prove that the quantities
A1 A1 A2 A2 A1 A1 A3 A3 A1 A2 A3 A3
J1 = − , J2 = − , J3 = −
γ1 γ2 γ1 γ3 γ2 γ3
are conserved under the flow of the system.
c) Prove for the Poisson brackets that {H, J1 } = {H, J2 } = {J1 , J2 } = 0. Use this
to show the complete integrability of the Hamiltonian system.
11.8. Solve the eigenvalue problem ∂x u(x) = v, ∂x v(x) = −s(x)λu(x), with the 1-
periodic function s(x) = χ[0,6/13] + 16χ(6/13,7/13) + χ[7/13,1] (x mod 1). Compute the
monodromy matrix Cλ and the discriminant D(λ) = trace Cλ . Hint. On intervals
where s is constant, the problem can be solved explicitly. We look for u ∈ Cb1 .
11.9. Consider the nonlinear wave equation ∂t2 u = ∂x2 u + u + 2ε2 cos(2x)u − u3 ,
with spatially periodic perturbed coefficients, i.e., 0 ≤ ε 1. For ε > 0 a spectral
gap of order O(ε2 ) occurs, which is too small to derive an NLS equation. In this
case with the ansatz
u(x, t) = εa(ε2 x, ε2 t)ei(x−ω0 t) + εb(ε2 x, ε2 t)e−i(x+ω0 t) + c.c.
derive the coupled mode system, cf. [SU01]
−2iω0 ∂T a =2i∂x a + ib − 3a|a|2 − 6a|b|2 ,
−2iω0 ∂T b = − 2i∂x b + ia − 3b|b|2 − 6b|a|2 .
11.10. If the pulses become very narrow, then the so called short pulse equation
∂ξ ∂τ A = A + ∂ξ2 (A3 ),
11.7. Nonlinear optics 449
can be derived. Consider the quasilinear wave equation ∂t2 u = ∂x2 u + u + ∂x2 (u3 ),
and make the ansatz
x−t
u(t, x) = 2εA(τ, ξ), τ = εt, ξ =
2ε
to derive the short pulse equation, cf. [PS13].
11.11. Show that the inhomogeneous Maxwell equations
∂t E = ∇ × B − J, ∇ · E = ρ, ∂t B = −∇ × E, ∇ · B = 0,
can be transformed to the inhomogeneous wave equations
∂t2 u − Δu = ρ, ∂t2 A − ΔA = J,
where A is a vector potential of B, i.e., ∇ × A = B, and u is a scalar potential of
E, i.e., E + ∂t A = −∇u, see Remark 11.7.1.
Hint. Consider a suitable gauge transform, i.e., adding ∇λ to A and subtracting
−∂t λ from u.
Chapter 12
It is the purpose of this chapter to explain the role of the KdV equation as
a long wave modulation equation for various dispersive wave systems. Such
approximations have a long history in science, and they not only play a
role in the description of the water wave problem but also for other nonlin-
ear dispersive systems such as the equations of plasma physics or the FPU
model.
One of the simplest systems where the KdV equation can be derived is
the so called Boussinesq equation
(12.1) ∂t2 u = ∂x2 u − ∂x4 u + ∂x2 (u2 ),
with x ∈ R, t ∈ R, and u(x, t) ∈ R. The long wave ansatz
u(x, t) = ε2 v(ξ, τ ) = ε2 v(εx, εt),
where 0<ε1 is a perturbation parameter, yields ∂τ2 v=∂ξ2 v+O(ε2 ), hence
(12.2) ∂τ2 v = ∂ξ2 v
in lowest order. The linear wave equation (12.2) is solved by
v(ξ, τ ) = v+ (ξ − τ ) + v− (ξ + τ ),
with functions v+ and v− satisfying
v(ξ, 0) = v+ (ξ) + v− (ξ) and ∂τ v(ξ, 0) = −v+ (ξ) + v− (ξ).
Hence, the formal long wave approximation predicts a wave moving to the
left and a wave moving to the right. However, this approximation is formally
only valid on a time scale of order O(1/ε). Nevertheless the right and
451
452 12. Long waves and their modulation equations
left moving wave will be separated for spatially localized initial conditions.
Afterwards the wave packets will evolve individually. For each of the wave
packets a KdV equation can be derived, and these are formally valid on the
much longer time scale of order O(1/ε3 ). In detail, inserting the ansatz
(12.3) u(x, t) = ε2 A+ (ε(x − t), ε3 t) + ε2 A− (ε(x + t), ε3 t).
into the Boussinesq equation and equating the coefficient at ε6 to zero yields
the system of two decoupled KdV equations
(12.4) 2∂T A+ − ∂X
3
A+ + ∂X (A2+ ) = 0, −2∂T A− − ∂X
3
A− + ∂X (A2− ) = 0,
one for each of the wave packets. There are some coupling terms which
however turn out to be of higher order for spatially localized initial condi-
tions. Estimates that this formal KdV approximation and true solutions of
the Boussinesq model stay close together over the natural KdV time scale
are a non-trivial task since solutions of order O(ε2 ) have to be estimated on
an O(1/ε3 ) time scale. This question will be discussed in §12.1.
It turns out that the set of KdV equations can be derived whenever
the dispersion relation for harmonic waves ei(kx+ω(k)t) gives two curves of
eigenvalues k → ω± (k) with ω± (0) = 0, and the nonlinear terms also vanish
at the spatial wave number k = 0. This is the reason why the KdV equation
occurs as a universal modulation equation, as discussed in §12.2. There, we
introduce the water wave problem, the equations of plasma physics and the
FPU system and show that they belong to the class of systems for which the
KdV equation can be derived. There are many other modulation equations
which can be derived in the long wave limit. Some of them are independent of
the small perturbation parameter ε such as the inviscid Burgers equation or
the Whitham system and some depend on the small perturbation parameter
ε such as the regularized long wave equation or the Benjamin-Bona-Mahony
system. We will comment on the use of these equations and on their validity
in §12.3. In §12.4 we explain that in the so called long wave limit, where
a big number of these modulation equations have been derived, the above
splitting in two decoupled KdV equations is sufficient to describe the original
system.
the derivative in front of the KdV nonlinearity. This approach has been
used in [KN86], but as pointed out in [CSS92] the error estimates for the
KdV approximation there have only been shown on an O(1/ε) time scale
instead of the O(1/ε3 ) KdV time scale. However, as pointed out in [Sch96b]
the approach of [KN86] can be extended to the correct O(1/ε3 ) KdV time
scale. This approach turns out to be very robust, but does not give optimal
results, cf. [CDS15] for another example.
In this book we will present the second approach which in general gives
better estimates and is conceptionally easier, but less robust. For the Boussi-
nesq equation (12.1) the KdV approximation can be justified with a simple
energy estimate. An energy is constructed in such a way that the influence
of the terms of order O(ε2 ) on the growth of the solutions can be estimated
to be of order O(ε3 ). The papers [Cra85, SW00b, SW02, Dül12] where
the KdV approximation has been justified for the water wave problem work
this way. The major difficulty in transferring this proof from the Boussinesq
equation to the water wave problem or other applications is to bring together
the energy estimates with the local existence and uniqueness theory for the
various systems. For the Boussinesq equation our approximation result is
as follows.
Theorem 12.1.1. Fix θ > 1/2 and let A+ = A ∈ C([0, T0 ], H θ+4 ) and
A− = 0 be solutions of the system of KdV equations (12.4). Then there
exist ε0 > 0 and C > 0 such that for all ε ∈ (0, ε0 ] we have solutions u of
(12.1) with
sup u(·, t) − ε2 A(ε(· − t), ε3 t)H θ ≤ Cε7/2 .
t∈[0,T0 /ε3 ]
Lemma 12.1.2. Fix θ ≥ 0 and let A+ = A ∈ C([0, T0 ], H θ+6 ) and A− = 0
be solutions of the system of KdV equations (12.4). Then there exist ε0 > 0,
Cres such that for all ε ∈ (0, ε0 ) we have
sup Res(ε2 Ψ(·, t, ε))H θ ≤ Cres ε15/2
t∈[0,T0 /ε3 ]
and
sup ∂x−1 Res(ε2 Ψ(·, t, ε))H θ ≤ Cres ε13/2 .
t∈[0,T0 /ε3 ]
All terms which can be written as a time derivative are included in E. The
others are estimated, namely
(∂t R) Res(ε2 Ψ) dx ≤∂t RL2 Res(ε2 Ψ)L2 ,
R
(∂τ Ψ)(∂x R)2 dx ≤∂τ ΨL∞ ∂x R2 2 ,
L
R
(∂t R)(∂X Ψ)∂x R dx ≤∂X ΨL∞ ∂t RL2 ∂x RL2 ,
R
(∂t R)(∂ 2 Ψ)R dx ≤∂ 2 ΨL∞ ∂t RL2 RL2 ,
X X
R
(∂x R)2 (∂t R) dx ≤∂x RL∞ ∂t RL2 ∂x RL2
R
≤C∂x RH 1 ∂t RL2 ∂x RL2 ,
We can estimate
(∂t ∂ −1 R)∂ −1 Res(ε2 Ψ) dx ≤∂t ∂ −1 RL2 ∂ −1 Res(ε2 Ψ)L2 ,
x x x x
R
(∂τ Ψ)R2 dx ≤∂τ ΨL∞ R2 2 .
L
R
456 12. Long waves and their modulation equations
As in §10.3 and §11.3 it turns out that Fourier transform is the key for
the understanding of the universality. Hence, we consider the Boussinesq
equation (12.1) in Fourier space. The Fourier transform u satisfies
(12.8) (k, t) = −ω 2 (k)
∂t2 u u∗2 (k, t),
u(k, t) − ρ(k)
√
where ω(k) = k 2 + k 4 sign(k) and ρ(k) = k 2 . By introducing w(k) =
1
(k)) we rewrite (12.8) into the first order system
u(k), ω(k) ∂t u
(
This system is diagonalized for fixed wave number k. For (12.9) the associ-
ated transformation
=√ 1 1 1
U
2 1 −1
−1 = U
is independent of k and unitary, i.e., U ∗ , cf. §11.1 where the same
transformation has been used a number of times. The transformed variable
∗w
z = U satisfies the diagonalized system
(12.10) z + U
∂t z = Λ ∗N
(U
z),
with Λ(k) = diag(iω(k), −iω(k)). In accordance with (12.3), for (12.10) we
make the ansatz
2 −1 k 3
z(k, t) = ε ε A+ , ε t eickte1
ε
(12.11)
2 −1 k 3
+ε ε A− , ε t e−ickte2 ,
ε
where
1 0
e1 = and e2 = .
0 1
Since the Fourier modes of the wave packets are concentrated in an O(ε)
neighborhood of the wave number k = 0, the evolution of the wave packets
will strongly be determined by the curves ±ω at k = 0. At eickte1 we find
1 + ε4 ∂T A
iε2 cK A 1
1 + i ε4 ∂ 3 ω(0)K 3 A
= iε2 ∂k ω(0)K A 1 − ε4 ∂k ρ(0) iK A
1 ∗ A
1 + O(ε6 ),
k
6 ω(0)
where k = εK. At ε2 we obtain the linear group velocity. At ε4 we obtain
a KdV equation.
458 12. Long waves and their modulation equations
ω ω
k→
k3 /6
k→ ω(k)
k k
12.2.1. The water wave problem. The most famous system where the
KdV equation can be derived is the so called water wave problem. Following
§12.1, there are no conceptional difficulties to justify this approximation,
but the realization of this approach for the water wave problem is rather
lengthy and involves a big number of estimates. Therefore, it is the only
goal of this section to explain that the water wave problem falls into the
class of systems for which the KdV equation can be derived. References to
the KdV approximation for the full water wave problem can be found at the
end of the chapter.
The water wave problem consists in finding the irrotational flow of an
inviscid incompressible fluid in an infinitely long canal of fixed finite depth
with an unknown free top surface subject to gravitational forces. The bot-
tom is impermeable and for expository reasons surface tension is neglected
in this subsection. The coordinates are denoted in horizontal direction by
x1 ∈ R and in the vertical bounded direction by x2 . The fluid fills the
domain Ω(t) in between the bottom {(x1 , −h) : x1 ∈ R} and the unknown
free top surface Γ(t). In the Eulerian formulation the free surface Γ(t) is
12.2. The universality of the KdV equation 459
x2
Γ (t)
0
x1
Ω (t) u2
u1
−1
x2 ) = ck cosh(k(1 + x2 )),
φ(k,
x =0 (k) = ck cosh(k) and ∂x
and so Φ| 2 2 φx2 =0 (k) = kck sinh(k). Therefore,
we have
∂x
2 φx2 =0 (k) = k(tanh k)Φ(k).
1.5
1
0.5
ω(k)
0
−ω(k)
−0.5
−1
−1.5
−3 −2 −1 0 1 2 3
k = 0. Hence the water wave problem falls into the class of systems described
in §12.2.
The KdV approximation. In case of positive surface tension (12.17)
changes into
1 ∂x1 η
(12.20) ∂t φ = ((∂x1 φ)2 + (∂x2 φ)2 ) − gη + σ∂x1 ,
2 1 + (∂x1 η)2
where σ is the surface tension parameter. The linear dispersion relation
then modifies into
(12.21) ω 2 = (k + σk 3 ) tanh(k).
There are essentially three situations, namely σ ∈ [0, 1/3), σ = 1/3, and
σ > 1/3. Expanding the curves of eigenvalues k → ω1,2 (k) at the wave
number k = 0 gives
1 1
ω1 (k) = k + (σ − )k 3 + O(k 5 ).
2 3
Hence, the sign in front of the cubic terms changes at σ = 1/3. At the same
value the inflection points of the curves k → ω1,2 (k) disappear. See Figure
12.4.
ω ω
k k
Figure 12.4. The curves of eigenvalues k → ω1,2 (k) of the water wave
problem with surface tension with σ ∈ (0, 1/3) in the left panel and
σ > 1/3 in the right panel.
462 12. Long waves and their modulation equations
12.2.2. The FPU system. The most famous conservative lattice differ-
ential equation where the KdV and NLS (see Exercise 11.4) equation can
be derived is the so called Fermi-Pasta-Ulam (FPU) system
(12.23) ∂t2 qj = W (qj+1 (t) − qj (t)) − W (qj (t) − qj−1 (t)) , j ∈ Z .
It was first studied numerically by Fermi, Pasta, and Ulam [FPU55] for
a finite set of oscillators in order to see how energy was spread through
the various modes of the system by the nonlinear coupling via the inter-
particle forces which are described by the potential function W : R → R.
They found that at low energy most trajectories did not “thermalize” as
expected, but rather exhibited a regular motion. This observation has been
explained in [ZK65], where Kruskal and Zabusky derived the KdV equation
as a formal approximation to the FPU system and in studying the KdV
equation numerically they found soliton dynamics. A rigorous proof that
long waves in the FPU system can be approximated via the KdV equation
has been given in [SW00a].
12.2. The universality of the KdV equation 463
where ω 2 has been defined in (12.28) and where in the convolution integrals
has to be used. Hence, for the derivation of the
the 2π-periodicity of u
NLS and KdV equations and for the justification of the associated approx-
imations we can proceed as in §11.1 and as above. A proof that the NLS
approximation makes correct predictions for the FPU model can be found
464 12. Long waves and their modulation equations
in [Sch10]. The KdV and NLS approximation has been justified for the
poly-atomic FPU model in [CCPS12].
We explain in the next subsection that at this point the analysis goes beyond
the formal calculations and makes a decision which of the equations should
be taken to approximate the long wave limit. The answer of [SW00b,
SW02] is that two ε-independent decoupled KdV equations are sufficient
to describe this limit. On the O(1/ε3 ) time scale for small ε > 0 no other
small amplitude dynamics can be found in the original system via the other
ε-dependent modulation equations.
Proof. For notational simplicity we only consider the case θ = 2. For the
approximation
ε2 Ψ(x, t, ε) = ε2 A+ (ε(x − ct), ε3 t) + ε2 A− (ε(x + ct), ε3 t)
we find using the previous calculations that
Res(ε2 Ψ) = −ε8 ∂T2 A+ − ε8 ∂T2 A− + 2ε6 ∂X
2
(A+ A− ).
As before we have
4∂T2 A+ =∂X
3 3
(∂X A+ − ∂X (A2+ )) − 2∂X (A+ (∂X
3
A+ − ∂X (A2+ ))).
12.4. The long wave limit 469
C
≤ A+ (·, T )H22 A− (·, T )H22 .
1 + (ε−2 T )2
Therefore, we have with the previous arguments
Lemma 12.4.2. Fix θ ≥ 0, and let A+ ∈ C([0, T0 ], H2θ+6 ) and A− ∈
C([0, T0 ], H2θ+6 ) be solutions of the system of KdV equations (12.4). Then
there exist ε0 > 0 and Cres > 0 such that for all ε ∈ (0, ε0 ) we have
1
Res(ε2 Ψ(·, t, ε))H θ ≤ Cres (ε15/2 + ε11/2 )
1 + (εt)2
and
1
∂x−1 Res(ε2 Ψ(·, t, ε))H θ ≤ Cres (ε13/2 + ε9/2 )
1 + (εt)2
for all t ∈ [0, T0 /ε3 ].
Remark 12.4.4. The analogous result for the water wave problem gives an
almost complete description of its long-time behavior in the long wave limit,
cf. [SW00b, SW02]. On a time scale O(1/ε) the solutions split up into
two wave packets, one moving to the right and one to the left. These wave
packets evolve independently as solutions of the system of KdV equations.
Their long-time behavior can be computed explicitly with the help of the
inverse scattering transform, cf. Remark 8.2.10. Some solitons which are
ordered w.r.t. their height evolve out of a dispersive remainder, see Figure
12.5, cf. [EvH81]. Approximation results such as Theorem 12.1.1 imply
that the same behavior will be observed in (12.1) in the long wave limit.
The approximation results from [SW00b, SW02] show the same behavior
for the water wave problem.
<−−− −−−>
Figure 12.5. The long time behavior for the water wave problem in the
long wave limit for t = T0 /ε3 with T0 large. The solutions of the water
wave problem in the long wave limit splits up into two wave packets,
one moving to the right and one to the left, where each of these wave
packets evolves independently as a solution of a KdV equation. For large
T0 solitons evolve out of a dispersive remainder.
Further Reading. There are various formulations of the water wave prob-
lem which differ in the chosen parametrization of the top surface. For the
different formulations there are a number of local existence and unique-
ness theorems. For the Eulerian formulation in §12.2.1, local existence and
uniqueness results have been shown for instance in [Shi76, KN79] with
the Cauchy-Kowalevslaya approach, and in [Lan05] with a Nash-Moser ap-
proach. For all other formulations the proofs are adaptions of the local ex-
istence and uniqueness theory for quasilinear hyperbolic equations [Kat75].
For the Lagrangian formulation such results have been shown for instance in
[Nal74, Yos82, Yos83, Cra85, Wu97, Wu99, SW00b, Igu01, SW02].
For the arc-length description of Γ(t) a local existence and uniqueness the-
orem has been shown in [AM05]. The theorems can be distinguished w.r.t.
2D or 3D, finite or infinite depth, with or without surface tension, and reg-
ularity of the initial conditions. See also [ABZ14]. Recently, a number of
almost global and global existence and uniqueness results have been estab-
lished [AL08, Wu09, Wu11, GMS12, AD15]. We refer to [Lan13] as
recent textbook.
12.4. The long wave limit 471
Exercises
12.1. Prove that solutions u(x, t) = v(x + t, ε2 t) of the regularized long wave
equation
∂t u = ∂x u + ε2 ∂x3 u + ε2 ∂x (u2 )
can be approximated via the solutions v of the associated KdV equation on an
O(1/ε2 ) time scale.
12.2. Use simultanuosly
ε2 ∂ξ3 Y = ε2 ∂ξ2 ∂τ Y + O(ε4 ) and ε2 ∂ξ3 Z = ε2 ∂ξ2 ∂τ Z + O(ε4 )
in (12.41)-(12.42) with different prefactors αj , and bring these terms to the right-
hand side. Then use
∂τ Y − αε2 ∂ξ2 ∂τ Y = (1 − αε2 ∂ξ2 )∂τ Y
472 12. Long waves and their modulation equations
together with
−1
1 − α1 ε2 ∂ξ2 −α2 ε2 ∂ξ2 1 0 α1 ∂ξ2 α2 ∂ξ2
= + ε2 + O(ε4 ).
−α3 ε2 ∂ξ2 1 − α4 ε2 ∂ξ2 0 1 α3 ∂ξ2 α4 ∂ξ2
to manipulate the coefficients on the right-hand side further and derive an approx-
imation system which is no longer symmetric w.r.t. to the interchange of Y and
Z.
12.3. Discuss the form of the smooth curves ω1,2 = ω1,2 (k) defined by
ω 2 = (k2 + μk4 )/(1 + k2 )
for |k| → ∞ and |k| → 0. Compute ∂kj ω1 |k=0 for j = 0, . . . , 4. Is there a change of
sign?
12.4. Consider the two-dimensional Boussinesq model
∂t2 u = Δu + ∂t2 Δu + Δ(u2 )
with u = u(x, y, t) ∈ R, x, y, t ∈ R, and Δ = ∂x2 + ∂y2 for modeling three-
dimensional water waves. By making the ansatz u = ε2 A(ε(x − t), ε2 y, ε3 t) derive
the Kadomtsev-Petviashvili (KP) equation
3
∂X ∂T A = ∂X (ν1 ∂X A + ν2 A∂X A) + ν3 ∂Y2 A,
with X = ε(x − t), Y = ε2 y. Compute the values of the coefficients ν1 , ν2 , and
ν3 ∈ R. Remark. The solutions of the KP equation describe unidirectional waves
slowly modulated in the direction normal to the direction of propagation. For
this model problem it has been pointed out in [GS01b] that the approximation
property for the KP equation depends strongly on the chosen initial condition of
the KP equation.
12.5. Derive the Bernoulli equation (12.17) from the Euler equations (12.13).
12.6. Derive a KdV equation for ∂t2 u = ∂x2 u + ∂x2 ∂t2 u − μ(∂x4 u + ∂x4 ∂t2 )u + ∂x2 (u2 )
by making the ansatz u(x, t) = ε2 A(ε(x − c0 t), ε3 t).
3
12.7. Derive a Kawahara equation ∂T A = c1 ∂X 5
A + c 2 ∂X A + c3 ∂X (A2 ) with T ∈ R,
X ∈ R, coefficients cj ∈ R, and amplitude A(X, T ) ∈ R for
∂t2 u = ∂x2 u + ∂x2 ∂t2 u − μ(∂x4 u + ∂x4 ∂t2 u) + ∂x2 (u2 ),
in case 1 + μ = ε2 ν, by making the ansatz u(x, t) = ε4 A(ε(x − c0 t), ε5 t).
12.8. Find solitary wave solutions of the Boussinesq equation
∂t2 u − ∂x2 u + 3∂x2 (u2 ) − ∂x4 u = 0
in the form u(x, t) = a sech2 (b(x−ct)). Discuss the well-posedness of this Boussinesq
equation.
Chapter 13
Center manifold
reduction and spatial
dynamics
In Chapters 10-12, our focus has been on the approximation of the dynam-
ics of complicated PDEs on the real line or cylindrical domains by simple
modulation equations. Often, by this approximation we described special
solutions such as bifurcating spatially periodic pattern, solitary waves, mod-
ulating front solutions, or modulating pulse solutions. It is not clear that
the corresponding exact solutions really exist in the original system, too.
The last decades saw big efforts and successes in the construction of these
special solutions using the Lyapunov-Schmidt reduction and the center man-
ifold reduction. There are a number of overview articles [Van89, VI92] and
textbooks [Car81, HI11] where especially center manifold theory and its
applications are well explained. We concentrate on the aspects of the theory
and its applications which have to do with modulation equations.
473
474 13. Center manifold reduction and spatial dynamics
For u ∈ 2,θ sufficiently small the right-hand side of (13.3) turns out to
be a contraction in a space of slowly exponentially growing functions. In the
finite-dimensional case etBc , etBs , etBu are well-defined and continuous from
Rd to Rd . In the infinite-dimensional case this has to be assumed. For our
purposes it is sufficient to assume the existence of β+ > 0 and β− < 0 such
13.1. The center manifold theorem 475
that (etBs )t≥0 and (etBu )t≤0 define C0 -semigroups in some space 2,θ which
satisfy
i) etBu 2,θ →2,θ ≤M eβ+ t , ∀ t ≤ 0,
ii) etBs 2,θ →2,θ ≤M eβ− t , ∀ t ≥ 0.
with a constant M . Moreover, the nonlinear terms gc , gs , and gu should be
C r+1 as functions from 2,θ to 2,θ . According to §5.1.2 in case of diagonal
matrices Bu and Bs the assumptions on the semigroups follow from the
position of the spectral values, for instance by assuming that
β+ = inf{Re λ : λ ∈ σ(Bu )} > 0, and β− = sup{Re λ : λ ∈ σ(Bs )} < 0.
Since Bc is finite-dimensional we have for all ε > 0 the existence of a M > 0
such that
(13.4) etBc 2,θ →2,θ ≤ M eε|t| , ∀ t ∈ R.
By assuming the existence of etBc , with this last estimate we can construct
a center manifold also in case that Bc is infinite-dimensional, cf. [GS01b,
GS05]. Similar to §5.1.2 the assumptions on the semigroup and nonlinearity
can be weakened.
The cut-off. Center manifolds in general only exist in a small neighbor-
hood of the fixed point, here u = 0. In order to construct them, we modify
the original system outside this neighborhood without changing the dynam-
ics close to the fixed point. To do so, we modify the nonlinearity g outside
a neighborhood of u = 0 by some cut-off function χ ∈ C ∞ ([0, ∞), R) with
the properties (i) 0 ≤ χ(r) ≤ 1 for all r ∈ [0, ∞), (ii) χ(r) = 1, if r ∈ [0, 1]
and (iii) χ(r) = 0, if r ≥ 2. We define gρ (u) = g(u)χ(ρ−1 u2,θ ). Since
gρ (u) = g(u) for u ∈ Bρ = {u ∈ 2,θ : u2,θ ≤ ρ} the vector fields of the
modified system and of the original system (13.1) coincide for all u ∈ Bρ .
In particular we have
Lemma 13.1.2. Let g ∈ C k (2,θ , 2,θ ) for a k ≥ 1 and gρ defined as above.
Then gρ ∈ C k (2,θ , 2,θ ) and
lim sup D
gρ (u)2,θ →2,θ = 0.
ρ→0 u∈2,θ
ii) For the existence proof we use the variation of constant formula (13.3)
which we abbreviate as
(13.5) u = Suc (0) + KG(u)
where
(Suc (0))(t) = etBc uc (0), G(u)(t) = g(u(t)),
and
t
(Ky)(t) = e(t−τ )Bc yc (τ )) dτ
0
t ∞
+ e(t−τ )Bs ys (τ )) dτ − e(t−τ )Bu yu (τ )) dτ.
−∞ t
We prove the existence of a fixed point of (13.5) by showing that the right-
hand side is a contraction in the metric space
Yη = {u ∈ C 0 (R, 2,θ ) : yη = sup e−η|t| y(t)2,θ < ∞}.
t∈R
for y1 , y2 ∈ Yη .
Proof. The first statement is obvious, since g is bounded. From the mean
value theorem it follows that
G(y1 ) − G(y2 )Yη ≤ sup e−η|t| g(y1 (t)) − g(y2 (t))2,θ
t∈R
−η|t|
≤ sup e DgC 0 y1 (t) − y2 (t)2,θ ≤ DgC 0 y1 − y2 Yη
b b
t∈R
for y1 , y2 ∈ Yη .
Lemma 13.1.6. For each η ∈ (ε, β) the map K is a bounded linear operator
from Yη into Yη , i.e., there is a function γ : (ε, β) → R such that
KYη →Yη ≤ γ(η).
∞
Proof. We write the last two terms of K as ∞ κ(t−τ )y(τ ) dτ . For η ∈ (ε, β)
and y ∈ Yη it follows
e−η|t| (Ky)(t)2,θ
6 t ∞ 7
−η|t| (t−τ )Bc η|τ | η|τ |
≤yη sup e e e dτ + κ(t − τ )e dτ
t∈R 0 −∞
6 t ∞ 7
(t−τ )Bc −η|t−τ | η|t−τ |
≤yη sup e e dτ + κ(t − τ )e dτ
t∈R 0 −∞
6 ∞ 0
τ Bc −ητ
≤yη max e e dτ, e e dτ
τ Bc ητ
0 −∞
∞ 7
η|τ |
+ κ(τ )e dτ
−∞
0 1
≤yη M (ε) (η − ε)−1 + 2(β − η)−1
which implies the assertion.
The last three estimates show that, for δ0 > 0 sufficiently small, the map
F (u) := Suc (0) + KG(u)
is a contraction in Yη for all η ∈ (ε, β) and g ∈ Cb1 (Rd ) with contraction
constant Kη DgC 0 < 1. Hence, F possesses a unique fixed point which
b
is a solution of (13.5). We write this solution of (13.5) as u = Ψ(Suc (0)).
iii) We define
ψ(uc (0)) := Ψ(Suc (0))(0) and set h = (0, ψs , ψu ).
478 13. Center manifold reduction and spatial dynamics
The continuity of Ψ implies the continuity of ψ and h. Since for g ∈ Cb1 the
map Ψ is Lipschitz-continuous the same is true for ψ and h. Thus, we have
proved the existence of a Lipschitz-continuous center manifold.
Remark 13.1.7. a) To complete the proof of Theorem 13.1.3 it remains to
prove that h is in C r−1 . This turns out to be rather complicated, because G
is in general not differentiable from Yη into Yη . However, it turns out that
G is k-times continuously differentiable from Yη1 into Yη2 , if η1 > kη2 . For
the heuristics of this consider g(x) = xk and x(t) = eη|t| .
b) Stable and unstable manifolds can be constructed similarly,
cf. [Van89]. For instance, in case of a stable manifold a fixed point is
constructed for the map
F (u) := Sus (0) + Ks G(u)
in the space
Zη+ = {u ∈ C 0 (R+ , 2,θ ) : uη = sup eηt u(t)2,θ < ∞},
t≥0
are smooth maps from 2,θ to 2,θ if θ ≥ 1. Hence, all assumptions for the
application of the center manifold theorem as stated in Theorem 13.1.3 are
satisfied. As in the finite-dimensional case, we add the equation ε̇ = 0 in
order to apply the center manifold theorem not only for ε = 0, but also for
small ε = 0. As a consequence there exist smooth functions hj such that for
the solutions on the center manifold
(13.8) uj = hj (u1 , ε2 ) = O(u31 ).
with π
2 3
γ= sin4 (x) dx = .
π 0 4
Hence, a supercritical pitchfork bifurcation occurs.
480 13. Center manifold reduction and spatial dynamics
for k ∈ (2Z + 1) \ {−1, 1}, it is obvious that the coefficients in the center
manifold reduction can be obtained by the coefficients from the GL approx-
imation.
13.2. Local bifurcation theory on bounded domains 481
Obviously the same construction is possible for all spatial periods close
to 2π.
482 13. Center manifold reduction and spatial dynamics
where
ν0 − ν2 q 2 i(qX+φ)
A0 [q, φ](X) = − e
ν3
is an equilibrium of the associated GL equation (10.103). The critical wave
number kc of the most unstable pattern U 1 (z)eikc x and the bifurcation
kc ,0
parameter 0 ≤ ε2 1 have been introduced in §10.7. The function h from
the center manifold reduction satisfies h[q, ε](εA0 , εĀ0 , ε) = O(ε2 ). Since
our center manifold theorem, Theorem 13.1.3, has not been formulated to
apply in this situation, more abstract versions like [HI11, Theorem 2.9] have
to be used for the reduction.
eigenvalues form two real sequences going to ±∞. See Figure 13.1.
Im
Re
Hence, the initial value problem for (13.12) is ill-posed in every 2,θ -
space. Nevertheless, the center manifold theorem, Theorem 13.1.3, can be
applied. The equations for n ≥ 2 can be written as
u n = λn vn , vn = λn un + λ−1
n gn (u0 , u1 , . . .).
This system can be diagonalized by introducing cn = un + vn and c−n =
un − vn which satisfy
c n = λn cn + gn , c −n = −λn c−n − gn .
Since functions satisfying Dirichlet boundary conditions on (0, π) can be
extended to odd 2π-periodic functions w.r.t. y, and since odd functions to
the power three are odd again, we can consider (13.11) alternatively in the
invariant subspace of odd 2π-periodic functions w.r.t. y. As a consequence
the nonlinearity is a smooth map from 2,θ to 2,θ for every θ > 1/2. Hence,
all assumptions of Theorem 13.1.3 are satisfied and so there exist smooth
functions uj = hj (u1 , α) for j ≥ 2, such that on the center manifold the
reduced system is given by
3
u 1 = (α + 1)u1 + g1 (u1 , h2 (u1 , α), . . .) = αu1 − u31 + O(u51 ).
4
Ignoring the higher order terms yields
3
(13.14) u 1 = v1 , v1 = (α + 1)u1 − u31 .
4
For α + 1 > 0 small, this system possesses two homoclinic orbits at the
origin which correspond to spatially localized solutions of the original elliptic
problem (13.11).
484 13. Center manifold reduction and spatial dynamics
13.4. Applications
The center manifold theorem in combination with spatial dynamics is nowa-
days a well established tool which has successfully been used in many appli-
cations. Examples are the construction of solitary surface water waves, of
standing light pulses in photonic crystals, or of breather solutions in lattice
differential equations. In this section we concentrate on the aspects which
have to do with modulation equations when the center manifold theorem
is applied, and explain how the spectrum of the linearized spatial dynam-
ics formulation is related to the spectrum of the linearized time-dependent
problem and how the reduced systems on the center manifold are related to
the associated modulation equations. We will concentrate on the construc-
tion of traveling wave solutions, modulating front solutions, and breather
solutions. We refrain from rewriting the literature and therefore skip all
functional analytic difficulties which have to be overcome in applying the
center manifold theorem and the discussion of the reduced system.
13.4.1. Solitary waves for the water problem. The center manifold
theorem in combination with spatial dynamics has especially been used in
the construction of solitary waves for the water wave problem as introduced
in §12.2.1, cf. [IK90, Ioo95]. The water wave problem can be written as
evolutionary system
i) The reduction in case σ = 0. For |c| < 1 there are two intersections of
k → ω± (k) and k → −ck at wave numbers k with |k| > 0, which lead to two
central eigenvalues. Since the spatial dynamics formulation (13.17) of the
water wave problem can be written as an infinite-dimensional Hamiltonian
system it can be expected that an infinite-dimensional version of Lyapunov’s
subcenter theorem can be applied. Therefore, the existence of spatially
periodic traveling waves can be established, cf. [Str26]. If |c| approaches
1 these two central eigenvalues collide in zero and leave the imaginary axis
along the real axis. For |c| > 1 there exists a homoclinic orbit to the origin.
This scenario is analogous to the following example.
vanish identically, these two central eigenvalues play no role and can be
eliminated. In fact, (13.21) can be integrated twice, i.e.,
(13.22) c2 v = v + c2 ∂ξ2 v + v 2 .
The constants of integration vanish since we are interested in solitary waves.
We write (13.22) as the first order system
(13.23) ∂ξ v0 = v1 , ∂ξ v1 = c−2 ((c2 − 1)v0 − v02 ).
We have two fixed points (v0 , v1 ) = (0, 0) and (v0 , v1 ) = (c−2 (c2 − 1), 0). For
c2 < 1 the origin is a center surrounded by a family of periodic solutions.
For c2 > 1 the origin is a saddle, and by looking at the phase portrait we
find a homoclinic solution at the origin which corresponds to the solitary
wave we are interested in and which exists for velocities c2 > 1.
From Remark 13.4.2 and the fact that the associated KdV equation pos-
sesses solitary waves Asol and the reversibility of the water wave problem it
follows that the reduced system on the center manifold possesses a homo-
clinic orbit for every c, with |c| > 1 not to big. By making the previous
ideas rigorous the following result [Ioo98] can be shown.
Theorem 13.4.3. There exists a δ0 > 0 such that for δ ∈ (0, δ0 ) the wa-
ter wave problem without surface tension possesses a solitary wave solution
U (x, y, t) = V (x − ct, y) with lim|ξ|→∞ V (ξ, y) = 0 satisfying
ii) The reduction in case σ > 1/3. For |c| > 1 there are two intersection
points of k → ω± (k) and k → −ck at wave numbers k with |k| > 0 which
lead to two central eigenvalues. The existence of spatially periodic traveling
waves can be established [Zei71] by applying an infinite-dimensional version
of Lyapunov’s subcenter theorem. If |c| approaches 1 these two central
eigenvalues collide in zero and leave the imaginary axis along the real axis.
For |c| < 1 there exists a homoclinic orbit to the origin. This scenario is
exactly as in the case σ = 0. However, as we already know from the KdV
approximation the solitary wave for σ > 1/3 is a wave of depression and not
a wave of elevation.
iii) The reduction in case 0 < σ < 1/3. For small values of |c| except
of the trivial intersection at k = 0 there are no further intersection points.
For a value |c| = cmin < 1 the curve k → −ck is tangent to k → ω(k) for a
k ≥ 0 which implies that two pairs of eigenvalues collide on the imaginary
axis. At this point the phase and group velocity of the associated NLS
approximation coincide. Thus, the modulating pulse solutions described
by the NLS approximation become traveling waves and can rigorously be
established with the present center manifold approach. For cmin < |c| < 1
there are four intersection points at wave numbers k with |k| > 0 which
lead to four central eigenvalues. The existence of spatially periodic traveling
waves can be established also in this case [Zei71]. As |c| approaches 1, two of
the four central eigenvalues collide in zero and leave the imaginary axis along
13.4. Applications 489
the real axis. The solitary waves in this case can be obtained via the KdV
approximation. However, the solitary waves do not decay to zero for |ξ| →
∞, but have some small oscillatory tails. The reason for this fact is that it
is very unlikely that a one-dimensional unstable manifold intersects with a
one-dimensional stable manifold in a four-dimensional space. However, the
three-dimensional center-unstable manifold and three-dimensional center-
stable manifold intersect due to the reversibility of the water wave problem
and so for |ξ| → ∞ the solutions in the intersection converge towards the
center manifold which is filled with periodic solutions.
15
ω(k)
ck
1
c2k
10
c3k
0
0 5 10 15 20
1
Im
c Re
σ
1/3
In order to use the center manifold theorem we have to compute the central
eigenvalues μ = ik ∈ iR which are solutions of
2
−cik = − 1 + (ik + im)2 + α = λ(k + m, α).
For c = 0 and α = 0 there are infinitely many intersection points. Since the
eigenvalue curve λ possesses maxima in k+m = ±1, the solution k = ±1−m
is double for each of these ms, which leads to a Jordan block of size two in
the spatial dynamics formulation. Hence, for c = 0 and α = 0 there are
infinitely many eigenvalues on the imaginary axis. Thus, at a first view a
reduction to a finite-dimensional center manifold does not seem possible.
However, at a second view an interesting phenomenon occurs. In order to
see the phenomenon take one of the front solutions
A(X, T ) = Af (X −
cT )
of the associated GL equation with
√
lim Af (ξ) = 0 and lim Af (ξ) = 1/ 3
ξ→∞ ξ→−∞
equation we also discuss the existence of breathers for nonlinear wave equa-
tions with spatially periodic coefficients.
The non-persistence of breathers for perturbations of the sine-
Gordon equation. The sine-Gordon equation
(13.25) ∂t2 u = ∂x2 u − sin(u),
with t ∈ R, x ∈ R, and u(x, t) ∈ R, originally came up in differential geome-
try in the form ∂ξ ∂η u = sin(u), describing surfaces with a constant negative
curvature. It was found to govern the propagation of a dislocation in a
crystal whose periodicity is represented by sin u, it was posed as a tentative
model of an elementary particle, and it was shown to be an equivalent form
of the so called Thirring model, cf. [DJ89]. It turns out to be a completely
integrable Hamiltonian system.
Remark 13.4.5. The name sine-Gordon equation is a pun referring to the
Klein-Gordon equation
(13.26) ∂t2 u = ∂x2 u − m2 u,
which is just the linear wave equation with the additional term −m2 u, where
m is the rest mass of the particle. The Klein-Gordon equation is the lin-
earized version of the sine-Gordon equation and of the nonlinear Klein-
Gordon equation (11.2), which in physics is often called φ4 -model, and has
been derived in 1928 as a relativistic version of the Schrödinger equation
describing free particles, i.e., it is invariant under Lorentz transformations.
Hence it attempts to unite quantum mechanics and special relativity, but it
has some serious flaws; see [Law90] for a very basic introduction.
for arbitrary x0 , t0 , and λ, with 0 < |λ| < 1. Due to the dynamical behavior
of these solutions they are called breather solutions. Surprisingly, it turns
13.4. Applications 493
out that the sine-Gordon equation is the only nonlinear wave equation
(13.28) ∂t2 u = ∂x2 u − u + g(u),
with g : R → R a smooth, odd function which satisfies g(u) = O(u3 ) and
g (0) > 0 for which such breather solutions exist [Den93, BMW94]. In
the following we explain why this ’non-persistence of breathers’ result holds.
Moreover, we explain a number of positive results about the existence of
generalized breather solutions.
Remark 13.4.6. With the ansatz
u(x, t) = εA(εx, ε2 t)eit + c.c.
the NLS equation
2
2i∂T A = ∂X A − A|A|2
can be derived which possesses localized time-periodic solutions. From the
approximation results in §11.1 it is known that these solutions approximately
exist in the nonlinear wave equations, too, on a time interval of length
O(1/ε2 ). However, the sine-Gordon equation up to rescaling is the only
of the nonlinear wave equations for which these solutions exist as localized
time-periodic solution for all times. In fact, taking the limit λ = 1 − ε2 → 1
in (13.27) shows that sine-Gordon breathers can be approximated by the
localized time-periodic solutions of the associated NLS equation.
(x) satisfy
The Fourier coefficients u
k (x) = ∂x2 u
−ω 2 k 2 u k (x) − u
k (x) + gk (u)(x),
where 2π/ω
ω
gk (u)(x) = e−iωkt g(u(x, t)) dt.
2π 0
The linearization of this spatial dynamics formulation possesses solutions of
k (x) = eλx with
the form u
λ2 = −ω 2 k 2 + 1.
From Remark 13.4.6 it follows that for small amplitude solutions we have
approximately ω = 1. Since we have an odd nonlinearity we only have to
consider k ∈ 2Z + 1. Hence, at the bifurcation point all eigenvalues are
on the imaginary axis. For ω a little bit smaller than 1 there is one stable
and one unstable eigenvalue and still infinitely many central eigenvalues.
Breather solutions lie in the intersection of the one-dimensional stable and
494 13. Center manifold reduction and spatial dynamics
Diffusive stability
d
(14.1) u = −u + up ,
dt
(14.2) ∂t u = ∂x2 u + up .
faster than the linear ones, and so it can be expected that the nonlinear
problem behaves asymptotically as predicted by the linear one.
497
498 14. Diffusive stability
asymptotic decay
cf. Exercise 14.1. Hence only for p > 3 the nonlinear terms vanish faster than
the linear ones, and so only for p > 3 it can be expected that the nonlinear
problem behaves asymptotically as predicted by the linear one. Nonlineari-
ties which can be controlled with the help of the linearized problem in this
way are called irrelevant w.r.t. diffusion or simply irrelevant. Nonlinearities
exactly at the boundary, p = 3, are called critical. Surprisingly, it turns out
that for many interesting and rather complicated problems from physics the
nonlinear terms are irrelevant. It is has been observed at the beginning of
the 1990s that the concept of diffusive behavior and irrelevant nonlinearities
plays an important role in stability questions of pattern forming systems.
The plan of this chapter is as follows. In §14.1 we recall from §7.1 the
most important properties of the linear diffusion equation from a slightly dif-
ferent point of view, and introduce the concept of irrelevant nonlinearities
and a number of different methods to establish the irrelevance of nonlin-
earities. Then in §14.2.1 we show the occurrence of diffusive behavior and
irrelevant nonlinearities in the real GL equation. In §14.2.2 we explain how
this approach can be transferred to pattern forming systems in order to show
the diffusive stability of spatially periodic equilibria in the SH equation, the
Couette-Taylor problem or Bénard’s problem. The rest of this section is
devoted to critical nonlinearities. In §14.3.1 we discuss exponentially long
transient behavior in unstable Poiseuille flow, and in §14.3.2 we consider the
Burgers equation as limit equation. Examples are the inclined film problem
or self-similar mixing of phases in pattern forming systems where the group
velocity depends on the wave number. In §14.4 we introduce phase diffusion
equations, which are modulation equations occurring in diffusive stability
theory. Finally, in §14.5 we revisit similarities and differences between dif-
fusive and dispersive dynamics.
(14.3) ∂t u = ∂x2 u,
14.1. Linear and nonlinear diffusive behavior 499
for all θ ≥ 2. Using the fact that Fourier transform is an isomorphism from
Hmθ to H m , cf. Lemma 7.3.31, we obtain equivalently
θ
√ √ √
0 (0)e−x /4 H 2 ≤ Ct−1/2 .
2
(14.6) t u(x t, t) − π u
θ
Most of the above ideas hold if we have a linear evolution operator eλ(k)t with
eigenvalues λ(k) ∼ −k 2 for k → 0. This is the reason why diffusive behavior
can be observed in a big variety of problems, as we will see. Textbooks about
self-similar solutions in various parabolic problems are [Bar96, SR10].
500 14. Diffusive stability
(14.9) − (k/2)∂k w
−k 2 w = f,
− λw
with f ∈ H2θ . The eigenfunctions ψs (k) = k s e−k to the real eigenvalues
2
14.1.3. L1 -L∞ estimates. This method relies on the Lq -Lp estimate (14.5),
the variation of constants formula, and suitable estimates for the nonlinear-
ity.
Lemma 14.1.2. Let p > 3. For all C > 0 there exists a δ > 0 such that
solutions u of ( 14.10) with u0 L1 + u0 L∞ ≤ δ satisfy
for all t ≥ 0.
Proof. We introduce
a(t) = sup u(s)L1 , b(t) = sup (1+s)1/2 u(s)L∞ .
0≤s≤t 0≤s≤t
such that
2
(1+t)1/2 et∂x u0 L∞ ≤ u0 L1 + u0 L∞ .
For the nonlinear terms we use up L∞ ≤ upL∞ and up L1 ≤ uL p−1
∞ uL1 .
We obtain
t (t−s)∂ 2 p
(1+t)
1/2
0e
x u (s) ds
∞ ≤ (1+t)
1/2 t e(t−s)∂x2
0 L1 →L∞ u L1 ds
p
t/2 L
≤ ((1+t)1/2 0 (t−s)−1/2 (1+s)−(p−1)/2 ds
t
+(1+t)1/2 t/2 (t−s)−1/2 (1+s)−(p−1)/2 ds) · b(t)p−1 a(t)
t/2
≤ ((1+t)1/2 0 (t−t/2)−1/2 (1+s)−(p−1)/2 ds
t
+(1+t)1/2 t/2 (t−s)−1/2 (1+t/2)−(p−1)/2 ds) · b(t)p−1 a(t)
≤ C1 b(t)p−1 a(t),
with a constant C1 independent of t for p > 3, and
t (t−s)∂ 2 p t (t−s)∂ 2
e x u (s) ds ≤ 0 e
x 1
L →L1 u L1 ds
p
0 1
t L
≤ 0 (1+s)−(p−1)/2 ds · b(t)p−1 a(t) ≤ C2 b(t)p−1 a(t),
with a constant C2 independent of t for p > 3. Thus,
a(t) ≤ a(0)+b(0)+|c|C2 b(t)p−1 a(t), b(t) ≤ a(0)+b(0)+|c|C1 b(t)p−1 a(t).
502 14. Diffusive stability
Introducing y(t) = a(t) + b(t) and adding the two inequalities yields
(14.13) y(t) ≤ 2y(0) + (C1 + C2 )y(t)p .
For y(0) < δ, with δ > 0 sufficiently small, the two curves y → y and
y → y(0) + (C1 + C2 )y p possess two intersection points y1 , y2 , with y1 < y2 .
Inequality (14.13) is valid for y ≤ y1 and y ≥ y2 . Since y(0) ≤ y1 and since
we have continuity of t → y(t) we must have y(t) ≤ y1 for all t ≥ 0.
According to Remark 14.1.1 we cannot expect faster decay rates even
if the initial conditions vanish much faster than a general L1 -function for
|x| → ∞, and for L∞ -initial conditions there is no decay. For functions
with an intermediate spatial decay, e.g., L2 initial conditions, we obtain the
following. On a linear level we have via Lemma 7.3.19 that
u(t)L∞ = G(t) ∗ u0 L∞ ≤ G(t)L2 u0 L2 ,
x2
1
with G(x, t) = √4πt e− 4t . We find
1/2
1 − x2 2 1 − 2x2 √ 1/2
G(t)L2 = √ e 4t dx = e 4 dy t ∼ t−1/4 ,
R 4πt R 4πt
the following result easily can be proved almost line for line as above.
Corollary 14.1.3. Let p > 5. For all C > 0 there exists a δ > 0 such that
solutions u of (14.10) with u0 L2 + u0 L∞ ≤ δ satisfy
u(t)L2 ≤ C and u(t)L∞ ≤ C(1+t)− 4
1
2
in
case Ω = R. We introduce
the functionals I(u) = R u dx, J(u) =
2 2 2 2
R (∂x u) dx and K(u) = R (∂x u) dx. For initial conditions with an L -
∞
decay for |x| → ∞, due to the L -L estimate (14.5) we can only expect to
2
Note that the method also allows to handle the boundary case p = 5. In
higher space dimensions the critical exponent obtained with this approach
can be reduced to p = 1 + 4/d, cf. Exercise 14.4.
where we choose w|τ =0 ∈ H22 . Moreover, w.l.o.g. for our purposes we set
c = 1. In H22 we have a zero eigenvalue, and the rest of the spectrum
is left of the line {λ ∈ C : Reλ = −1/2}, cf. Remark 14.1.1. Since the
solutions of the linearized problem are uniformly bounded and since the
nonlinear terms vanish with an exponential rate there exist δ, C > 0 such
that supτ ≥0 w(τ )H22 < C for the solutions w of (14.15) if w|τ =0 H22 < δ
and p > 3. If we denote by w0 the part of w belonging to the eigenvalue 0
and with w1 the rest of w, then by integration of the variation of constant
formula w.r.t. time we can conclude that
3−p 1 3−p
w0 (·, τ ) = wlim e−(·) }τ
2 /4
+ O(e 2
τ
) and w1 (τ ) = O(emax{− 2 , 2 )
for τ → ∞, where wlim ∈ R is a constant only depending on the initial
conditions. This leads to the following convergence result, cf. [Way97].
Theorem 14.1.5. Let p > 3. There exist δ, C > 0 such that for all solutions
u of (14.10) with u|t=0 H22 ≤ δ, there exists a wlim ∈ R such that
√ √
t u(· t, t) − wlim e−(·) /4 H22 ≤ C(1+t)− max{1/2,(3−p)/2} for all t ≥ 0.
2
approach to give a
Proof of Theorem 14.1.5. In Fourier space (14.10) is given by
(14.19) = −k 2 u
∂t u ∗p ,
+u
where u∗p denotes the p-times convolution of a function u ∗p = u
, i.e., u ∗. . .∗
. Corresponding to (14.16) we rescale u
u to u −n
(L k, L2n τ ) for a
n (k, τ ) = u
fixed L > 1. Note that due to the scaling properties of Fourier transform
. We obtain
there is no Ln in front of u
(14.20) n = −k 2 u
∂τ u ∗p
n + Ln(3−p) un ,
14.1. Linear and nonlinear diffusive behavior 505
where a factor Ln(1−p) in front of the convolution terms comes from the
substitution in the integrals. In order to simplify notation we denote all
constants with the same symbol C if they can be chosen independently of
L. We solve (14.19) via (14.20) by the following renormalization procedure.
(RG)n→n+1 : Equation (14.20) is solved on the time interval [1/L2 , 1].
n+1 (k, 1/L2 ) = u
Then u n (k/L, 1) is taken as initial condition for n + 1.
For solving (14.20) we use the variation of constant formula
n (k, τ ) = e−k
2 (τ −1/L2 )
u n−1 (k/L, 1)
u
(14.21) τ
e−k ∗p
2 (τ −s)
+Ln(3−p) 1/L2 un (k, s) ds.
e−(·)
2 (τ −1/L2 )
≤ sup C 2
un−1 (·/L, 1)H22 ≤ CL5/2
un−1 (·, 1)H22 .
b
τ ∈[1/L2 ,1]
This estimate for the semigroup and Sobolev’s embedding theorem show for
the nonlinear terms
τ
n(3−p) −k2 (τ −s) ∗p
sup L e n (·, s) ds ≤ CLn(3−p) Rnp .
u
τ ∈[1/L2 ,1] 1/L 2 2
H2
with An = Πun |τ =1 towards zero and the existence of the limit limn→∞ An .
By definition we have ρn |k=0 = 0. The variables An and ρn satisfy
1
∗p
e−k (1−s) u
n(3−p) 2
An+1 =An + Π(L n+1 (k, s) ds),
1/L2
1
−k2 (1−1/L2 )
e−k ∗p
2 (1−s)
ρn+1 (k) =e ρn (k/L) + L n(3−p)
un+1 (k, s) ds
1/L2
+ e−k
2 (1−1/L2 )
An ψ(k/L) − An+1 ψ(k).
For the terms on the right-hand side we get
1
∗p
e−(·) (1−s) u
2
L n(3−p)
n+1 (·, s) dsH22 ≤ CRn L
p n(3−p)
,
1/L2
|Πu| ≤ CuH22 ,
e−(·)
2 (1−1/L2 )
ρn (·/L)H22 ≤ (C/L)
ρn H22 ,
e−(·)
2 (1−1/L2 )
An ψ(·/L) − An+1 ψ(·)H22 ≤ C|An+1 − An |.
Using these estimates we obtain
|An+1 − An | ≤CLn(3−p) Rnp ,
ρn+1 H22 ≤(C/L)
ρn H22 + CLn(3−p) Rnp .
Proof. Using the notation from the proof of Theorem 14.1.5 the statement
follows from
√
u(·, t)L∞ ≤Ct−1/2 e−(·) /(4t) L∞ + Ct−1 ρ(·/ t, t)L∞
2
After showing the linear diffusive behavior we have to establish the irrel-
evance of the nonlinearity w.r.t. to this diffusive behavior. We stay a little
bit more abstract than necessary at this point and expand the nonlinear
terms into
(
pc N vc , vc ) + B2,2 (
v ) =B2,1 ( vc , vs ) + B3 (
vc , vc , vc ) + gc (
vc , vs ),
(
Ps N vc , vc ) + gs (
v ) =B2,3 ( vc , vs ),
where the B2,j are symmetric bilinear maps, where B3 is a symmetric trilin-
ear map, and where gc and gs stand for the remaining terms. The splitting
is motivated as follows. If vc decays like t−1/2 , then vs , which is slaved to
vc , decays at least like t−1 . Then gc = O(| vc |4 + |
vc |2 |
vs | + |
vs |2 ) decays like
−2
t and gs = O(| vc | + |
3 vc ||
vs | + |
vs | ) decays like t
2 −3/2 . They are therefore
both irrelevant w.r.t. diffusive behavior.
14.2. Diffusive stability of spatially periodic equilibria 511
(14.29) c ,
vc = w −1 Es B2,3 (
vs = −L vc , vc ) + w
s
c ) =B2,1 (w
N1 ( w c , w
c ) + B2,2 (w −1 Es B2,3 (w
c , −L c , w
c ))
c , w
+ B3 ( w c , w
c ),
c , w
N2 ( w s ) =gc (w −1 Es B2,3 (w
c , −L c , w
c ) + w
s ) + B2,2 (w
c , w
s ),
c , w
N3 ( w s ) =gs (w −1 Es B2,3 (
c , −L vc , vc ) + w −1 Es B2,3 (w
s ) + ∂t (L c , w
c )).
By this transformation N2 decays like t−2 and N3 decays like t−3/2 such
that they are both irrelevant w.r.t. diffusive behavior. In order to prove the
irrelevance of N1 (wc ) we write it as
c ) = K2 (k, k − l, l)w
N1 ( w c (k − l)w
c (l) dl
R
+ K3 (k, k − l, l − m, m)wc (k − l)w
c (l − m)w
c (m) dm dl,
R R
with ν < 2 fixed arbitrarily close to 2, and denote many different con-
stants with the same symbol C if they can be chosen independently of
ac (T ), . . . , bs (T ), and T .
From the previous representations and Lemma 7.3.19 we find
c )L1 ≤C(|k|w
N1 (w c 2L1 + w
c L1 |k|2 w
c L1 + w
c 2L1 |k|w
c L1 ),
c , w
N2 (w s )L1 ≤C(w
c 4L1 + w
c L1 w
s L1 + w
s 2L1 ),
c , w
N3 (w s )L1 ≤C(w
c 3L1 + w
c L1 w
s L1 + w
s 2L1 + w
c L1 ∂T w
c L1 ),
and
c )L∞ ≤C(|k|w
N1 (w c L1 |k|w
c L∞ + w
c L1 |k|2 w
c L∞
+ w
c 2L1 |k|w
c L∞ ),
c , w
N2 (w s )L∞ ≤C(w
c 3L1 w
c L∞ + w
c L∞ w
s L1 + w
s L1 w
s L∞ ),
c , w
N3 (w s )L∞ ≤C(w
c 2L1 w
c L∞ + w
c L∞ w
s L1 + w
s L1 w
s L∞
+ w
c L∞ ∂T w
c L1 ),
c L1 on the right-hand side will be replaced via
where ∂T w
c L1 ≤ C(|k|2 w
∂T w c L1 + N1 (w
c )L1 + N2 (w
c , w
s )L1 ).
c has compact support in Fourier space, |k|2 w
Since w c can be estimated in
terms of |k| w
ν c for every ν ∈ [0, 2).
14.2. Diffusive stability of spatially periodic equilibria 513
Similarly we find
T
e μ c (T −τ )
N2 (w s )(τ ) dτ
c , w ≤ C(b3c (T )ac (T )+ac (T )bs (T )+as (T )bs (T )).
0 L∞
Similarly we find
T
(1+T )1/2 e μc (T −τ )
s )(τ ) dτ
c , w
N2 (w
0 L1
≤ C(b3c (T )ac (T ) + ac (T )bs (T ) + as (T )bs (T )).
514 14. Diffusive stability
A3) We estimate
T
ν/2
(1+T ) e μc (T −τ ) ν
c )(τ ) dτ
|k| N1 (w ∞
0 L
T
≤(1+T )ν/2 eμc (T −τ ) |k|ν L∞ →L∞ N1 (w c )(τ )L∞ dτ
0
T
≤(1+T ) ν/2
C (T − τ )−ν/2 (1+τ )−3/2 dτ · (cc (T )dc (T ) + b2c (T )cc (T ))
0
≤C(cc (T )dc (T ) + b2c (T )cc (T )).
T T /2 T
again by separating 0 = 0 + T /2 . Similarly we find
T
ν/2
(1+T ) e μc (T −τ )
|k| N2 (w
ν
s )(τ ) dτ
c , w
0 L∞
≤ C(b3c (T )ac (T ) + ac (T )bs (T ) + as (T )bs (T )).
C) We set
r(T ) = ac (T ) + bc (T ) + cc (T ) + dc (T ) + cc,ν (T ) + dc,ν (T ) + as (T ) + bs (T ).
Summing up all estimates yields an inequality
r(T ) ≤ r(0) + f (r(T ))
where f is at least quadratic in its argument. Comparing the curves r → r
and r → δ + r2 it is easy to see that r cannot go beyond 2δ if δ > 0 is
sufficiently small. Hence, if r(0) < δ, with δ > 0 sufficiently small, we have
the existence of a C > 0 such that r(T ) ≤ C for all T ≥ 0.
We will apply the scheme of the proof a second time, namely for showing
the diffusive stability of the time-periodic solution A(X, T ) = e−iβt of the
complex GL equation (14.26). In order to do so we introduce B(X, T ) =
A(X, T )eiβt which satisfies
2
∂T B = (1 + iα)∂X B + (1 + iβ)B − (1 + iβ)B|B|2 .
This equation possesses the equilibrium B = 1 for which we would like to
prove diffusive stability. The deviation v = B − 1 satisfies
(14.34) ∂T v = Lv + N (v)
where
2
Lv =(1 + iα)∂X v + (1 + iβ)v − (1 + iβ)(2v + v),
N (v) = − (1 + iβ)(v 2 + 2vv + v 2 v).
The eigenvalue problem Lv = μv is solved by Fourier modes v(x, t) =
v1,2 (k)eikx with eigenvalues μ1,2 (k). In the Eckhaus stable domain we have
constants σc > 0 and σs > 0 such that
Reμ1 (k) ≤ −σc k 2 and Reμ2 (k) ≤ −σs
for all k ∈ R. System (14.34) has exactly the same properties as System
(14.27). Therefore, we can follow the proof of Theorem 14.2.1 starting at
(14.28) line for line. The formal calculations for the irrelevance of the nonlin-
ear terms can be found in the subsequent §14.4.2. With the same arguments
14.2. Diffusive stability of spatially periodic equilibria 517
Remark 14.2.6. Obviously, with the presented method the diffusive stabil-
ity of all Eckhaus-stable equilibria of (14.22) and of all Eckhaus-stable time-
periodic solutions of (14.26) can be established. We already explained that
whenever the L1 -L∞ approach works the renormalization approach works,
too. See [BK92] for a detailed proof for the real GL equation. The renor-
malization approach gives more informations than the L1 -L∞ approach, but
needs more localization of the initial conditions.
Theorem 14.2.7. There exists a δ > 0 such that for all p ∈ (0, 1/2) there
exists a C1 > 0 such that for all (φ0 , s0 ) ∈ H22 × H22 with φH22 + sH22 ≤ δ
there exists a unique global solution (φ, s) ∈ C([1, ∞), H22 × H22 ) of (14.24)
to initial conditions (φ0 , s0 ), and a φlim ∈ R such that
(14.35)
T 1/2 φ(T 1/2 ·, T )−φlim e−· H22 + T 1/2 s(T 1/2 ·, T )H22 ≤ C1 T −1/2+p ,
2 /4
compact resolvent (L − μI)−1 for some μ ∈ C. From this, for fixed , it
follows the existence of a discrete set of eigenvalues
{μj () ∈ C : j ∈ N, μj ≥ μj+1 → −∞ for j → ∞}
and a corresponding set of eigenfunctions {fj () : j ∈ N}. We normalize fj
such that fj (, ·)L2 (T) = 1.
For ε = 0 we have uper,ε = 0 and so the eigenvalue problem can be solved
explicitly. The well known curve λ(k) = −(1−k 2 )2 is cut into pieces of length
one. These pieces are plotted via k = m+, with m ∈ Z and ∈ [−1/2, 1/2),
as functions w.r.t. , i.e., we consider the spatial homogenous situation
artificially as spatially periodic. Spectral perturbation theory for fixed , cf.
[Kat95] or Lemma 11.6.2, yields the smoothness of the curves of eigenvalues
with the intersection points as exception. Hence, it is easy to see that for
fixed 1 = O(1) > 0 for ε → 0 there exist ε0 > 0 and σ0 (1 ) = O(1) > 0
for ε → 0, such that for all ε ∈ (0, ε0 ) and all || > 1 all eigenvalues satisfy
μj () < −σ0 . Except of the two curves μ1 , μ2 touching zero this is also true
for all other eigenvalues, i.e., supj≥3 sup∈[−1/2,1/2) μj () < −σ0 .
In order to understand what happens with the two curves μ1 , μ2 for
ε > 0 we proceed as follows. The eigenvalue problem
v = L v − μ
G(, ε, μ) v=0
always has the trivial solution v = 0 for all , μ, and ε. Solutions can bifur-
cate from this trivial branch if (∂vG(, ε, 0))−1 does not exist. We know al-
ready that ∂w G(0, 0, 0) = L0 |ε=0 is not invertible and has a two-dimensional
kernel spanned by U1 = sin x and U2 = cos x. Thus, we apply the Lyapunov-
Schmidt reduction method to compute the bifurcating solutions. Let P be
the orthogonal projection on this kernel and let v = aU1 + bU2 + V with
(1−P )V = V. The solution of the hyperbolic part (1−P )G(, ε, μ, a, b, V) = 0
is denoted by V = V(, ε, μ, a, b). Inserting this into P G(, ε, μ, a, b, V) = 0
gives the bifurcation equation
a
G1 = 0,
b
where
ρ−μ −iδ O(2 + |μ|) O(|| + |μ|)
G1 = + O(ε ) 4
,
iδ ρ + c(ε) − μ O(|| + |μ|) O(2 + |μ|)
with
ρ = −42 − 4 , δ = −43 , c(ε) = −2ε2 + O(ε4 ).
Computing μ1,2 (, ε) in such a way that the determinant of G1 vanishes,
gives
520 14. Diffusive stability
where χ() = 1 for || ≤ 1 /2 and χ() = 0 for || > 1 /2, allows to separate
the diffusive modes from the exponentially damped modes. Moreover, define
v = v − (
Ps () v)fc ().
pc ()
We write v = vc fc + vs and use the projections to separate (14.39) into
two parts, namely
(
∂t vc = μc vc + pc N v ), s vs + Ps N
∂t vs = LP (
v ).
s we have the estimates
For the semigroups generated by μc and LP
eμc t vc L1 ≤eμc t L∞
vc L1 ≤
vc L1 ,
eμc t vc L∞ ≤eμc t L∞
vc L∞ ≤
vc L∞ ,
vc L∞ ≤ Ct−1/2
eμc t vc L1 ≤eμc t L1 vc L∞ ,
eLPs t vs L1 ≤Ce−σs t/2
vs L1 ,
eLPs t vs L∞ ≤Ce−σs t/2
vs L∞ ,
with a constant σs > 0, where we used the abbreviation
L1 = L1 ([−1/2, 1/2), H 1 (T)) and L∞ = L∞ ([−1/2, 1/2), H 1 (T)).
By the choice H 1 (T) the nonlinear terms are well-defined subsequently. Af-
ter showing the linear diffusive behavior we have to establish the irrelevance
of the nonlinearity w.r.t. to this diffusive behavior. With the same motiva-
tion as before we expand the nonlinear terms into
(
pc N vc , vc ) + B2,2 (
v ) =B2,1 ( vc , vs ) + B3 (
vc , vc , vc ) + gc (
vc , vs ),
(
Ps N vc , vc ) + gs (
v ) =B2,3 ( vc , vs )
14.2. Diffusive stability of spatially periodic equilibria 521
where the B2,j are symmetric bilinear maps, B3 symmetric trilinear maps,
and where gc and gs stand for the remaining terms. In order to prove the
irrelevance all these terms we make a change of coordinates. We set
(14.40) c ,
vc = w −1 Es B2,3 (
vs = −L vc , vc ) + w
s
and obtain
c = μc w
∂t w c + N1 (w
c ) + N2 (w c , w
s ),
(14.41)
s = Ls Ps w
∂t w s + N3 (wc , w
s ),
where
c ) =B2,1 (w
N1 ( w c , w
c ) + B2,2 (w −1 Es B2,3 (w
c , −L c , w
c )) + B3 (w
c , w
c , w
c ),
c , w
N2 ( w s ) =gc (w −1 Es B2,3 (w
c , −L c , w
c ) + w
s ) + B2,2 (w
c , w
s ),
c , w
N3 ( w s ) =gs (w −1 Es B2,3 (
c , −L vc , vc ) + w −1 Es B2,3 (w
s ) + ∂t (L c , w
c )).
By this transformation N2 decays like t−2 and N3 decays like t−3/2 such that
they are both irrelevant w.r.t. diffusive behavior.
In order to prove the irrelevance of N1 (w c ) we write it as
c ) = K2 (, − 1 , 1 )
N1 ( w vc ( − 1 )
vc (1 ) d1
R
+ K3 (, − 1 , 1 − 2 , 2 )
vc ( − 1 )
vc (1 − 2 )
vc (2 ) d2 d1 ,
R R
with smooth kernels K2 (, − 1 , 1 ) and K3 (, − 1 , 1 − 2 , 2 ). As be-
fore, the faster the kernels vanish near (k1 , . . . , kj ) = (0, . . . , 0), the more
irrelevant are the associated terms. We have
(14.42) K2 (0, 0, 0) = 0, K3 (0, 0, 0, 0) = 0
by considering the SH equation in the space of 2π-spatially periodic func-
tions. There exists a one-dimensional center manifold associated to family
of equilibria {uper,ε (· + x0 ) : x0 ∈ R}, cf. §13.2.2. The flow on the center
manifold is given by
(14.43) c (0) = μ1 (0)w
∂t w c (0) + β2 w
c (0)2 + β3 w
c (0)3 + . . . ,
with coefficients βj ∈ R. Since the flow on the center manifold is trivial,
we must have ∂t wc (0, t) = 0 and so β2 = β3 = 0. Since after the transform
(14.40) the right-hand side of (14.41) in the space of 2π-spatially periodic
functions for ws = 0 coincides with the right-hand side of (14.43) up to
terms of order O(|wc |4 ) we must have
β2 = K2 (0, 0, 0), β3 = K3 (0, 0, 0, 0)
and so (14.42). So it remains to prove that the first derivatives of K2 van-
ish at (0, 0, 0), too. This property can be computed explicitly for the SH
522 14. Diffusive stability
equation using that the curve of eigenfunctions f1 () has the following useful
property.
Lemma 14.2.9. The eigenfunction f1 can be expanded as f1 () = g0 +ig1 +
O(2 ), where g0 , g1 ∈ H 1 (T) are real-valued functions.
Proof. We have g0 = ∂x uper,ε /∂x uper,ε (x)L2 (0,π) . The operator L is of the
form L = A+iB +O(2 ), where A and B are real -independent operators.
The eigenvalues μ1 satisfy μ1 () = O(2 ). Let h1 ∈ L2 (T) also be real-valued.
Inserting the ansatz f1 () = g0 + ig1 + h1 + O(2 ) into L f1 () = μ1 ()f1 ()
gives for the terms of O() the condition iAg1 + Ah1 + iBg0 = 0. Thus,
h1 ∈ kerA, i.e., h1 g0 = β∂x uper,ε for a β ∈ R. Since we have normalized
f1 we have to choose h1 = 0.
Using the above representation of the projection P1 () shows
K2 (, − m, m) = 3 f1 (, x)f1 ( − m, x)f1 (m, x)
u0 (x) dx.
R
Using Lemma 14.2.9 shows that
6
K2 (, −m, m) = 3 g0 (x)g0 (x)g0 (x)
u0 (x)
R
−ilg1 (x)g0 (x)g0 (x)
u0 (x)+g0 (x)i(−m)g1 (x)g0 (x)
u0 (x)
7
+g0 (x)g0 (x)img1 (x)u0 (x) + O(l2 +(−m)2 +m2 ) dx.
Note that u0 (x) is an even function, and so g0 (x) = ∂x uper /∂x uper L2 (0,2π)
is an odd function. Thus, the integral over the zero order terms goes over an
odd function and vanishes. Since additionally the first order terms cancel,
we have shown
|K2 (, −m, m)| ≤C|2 + (−m)2 + m2 |,
|K3 (, − m, m − k, k)| ≤C(|| + | − m| + |m − k| + |k|).
Therefore, we have a system with exactly the same properties as for the GL
equation, and so we get
Theorem 14.2.10. The stationary solution u = uper,ε of (10.4) is diffu-
sively stable in the following sense. For all C > 0 there exists a δ > 0 such
that for solutions u = uper,ε + v of (14.22) with v |t=0 L1 +
v |t=0 L∞ ≤ δ
we have
v (t)L∞ ≤ C,
v (t)L1 ≤ C(1+t)−1/2
for all t ≥ 0.
Corollary 14.2.11. For all C > 0 there exists a δ > 0 such that solutions
u = uper,ε + v of (14.22) with
v |t=0 L1 + v|t=0 L1 ≤ δ satisfy
v(t)L∞ (R,R) ≤ C(1+t)−1/2
for all t ≥ 0.
nonlinearity show the same decay rates as diffusion but the limiting profile
is no longer a Gaussian.
The analogue to Theorem 14.3.1 and Proposition 14.3.2 has been proved
in [SU03a] for Poiseuille flow. This makes the measurement of the critical
Reynolds number Rc a delicate experiment. Initial perturbations may be
amplified on a short transient time scale due to the non-normality of the
linearization around the laminar flow, see, e.g., [SH94a]. After the short
time transient growth the solutions seem to decay for a very long time, i.e.,
at criticality for an exponentially long time. The better the experiment is
performed the more and more this observation plays a role and the longer
it takes to observe the final growth of localized perturbations. For more
details see Figure 14.1.
stable
Re λ
unstable
kc k
Rc R
rest of spectrum
can be found. The second term on the right-hand side of (14.46) is estimated
as follows. We have
t
1/2
(1+t) e (t−s)∂x2
∂x (u (s)) ds
2
∞
0 L
t/2
2
≤ (1+t)1/2 e(t−s)∂x ∂x L1 →L∞ u2 L1 ds
0
t
2
+ (1+t)1/2 e(t−s)∂x ∂x L∞ →L∞ u2 L∞ ds
t/2
t/2
≤ ((1+t)1/2 (t−s)−1 (1+s)−1/2 ds · b(t)a(t)
0
t
+(1+t)1/2 (t−s)−1/2 (1+s)−(p−1)/2 ds) · b(t)2
t/2
t/2
≤ ((1+t)1/2 (t−t/2)−1 (1+s)−1/2 ds · b(t)a(t)
0
t
+(1+t) 1/2
(t−s)−1/2 (1+t/2)−1 ds) · b(t)2
t/2
≤ C1 (b(t)a(t) + b(t)2 ),
with a constant C1 independent of t, and
t (t−s)∂ 2 t (t−s)∂ 2
e x ∂ (u2 (s)) ds
1 ≤ 0 e
x∂ 1
x L →L1 u L1 ds
2
0 x
t L
≤ 0 (t − s)−1/2 (1+s)−1/2 ds · b(t)a(t) ≤ C2 b(t)a(t),
with a constant C2 independent of t such that finally
a(t) ≤a(0) + b(0) + C2 (b(t)a(t) + b(t)p−1 a(t))
b(t) ≤a(0) + b(0) + C1 (b(t)a(t) + b(t)2 + b(t)p−1 a(t)).
If a(0) + b(0) < δ, with δ > 0 sufficiently small, we have the existence of a
C > 0 such that a(t), b(t) ≤ C for all t ≥ 0.
Also here, whenever the L1 -L∞ approach works, the renormalization
approach works too. In [BKL94] the following diffusive stability result has
been shown with the discrete renormalization approach from §14.1.6.
Theorem 14.3.4. Let b ∈ (0, 1/2). Then there exist C, δ > 0 such that for
all initial conditions u0 of (14.45) with u0 H22 ≤ δ there is a z > −1 such
that the solution u of (14.45) with u|t=1 = u0 satisfies
√ √
tu(x t, t) − fz∗ (x)H22 ≤ Ct−1/2+b ,
hence in particular
u(x, t) − t−1/2 fz∗ (t−1/2 x)L∞ ≤ Ct−1+b .
14.4. Phase diffusion equations 529
y, v g
h(t, x)
x, u
with a laminar flow and a flat surface. For sufficiently small Reynolds num-
ber this flow turns out to be stable with a self-similar decay as in Theorem
14.3.4. Using the discrete renormalization approach from §14.1.6, this has
been proved for the so-called integrated boundary layer (IBL) approximation
of the inclined film problem in [Uec03], for the free surface Navier-Stokes
problem in [Uec07], and for the IBL over wavy bottoms in [HSU12], which
adapts the Bloch wave analysis from §14.2.2 to quasilinear systems.
systems these are phase diffusion equations for the description of wave num-
ber modulations of a family of spatially periodic equilibria and depending
on the scaling the Burgers equation or a conservation law for the description
of wave number modulations of a family of traveling waves where the group
velocity depends in a non-trivial way on the wave number. For conservative
systems Whitham’s equations for the description of wave number and am-
plitude modulations of a family of traveling waves appear, cf. [Whi99]. We
explain the derivation of these equations for the real and complex GL equa-
tion. Moreover, we explain that due to the reconstruction of the solutions
from the wave number, approximation results can only be expected locally,
but not globally, in space. These equations give a feeling about the diffu-
sive dynamics which can be expected in more complicated pattern forming
systems.
14.4.1. The phase diffusion equation for the real GL equation. The
real GL equation
(14.47) 2
∂T A = ∂X A + A − A|A|2 ,
with X ∈ R, T ≥ 0, and A(X, T ) ∈ C possesses the two-dimensional family
of stationary solutions
(14.48) A = Aper [q, φ0 ](X) = 1 − q 2 eiqX+iφ0 ,
which is parameterized by φ0 , q ∈ R. In §8.3.6 we discussed the linearization
around these equilibria and found diffusive modes and exponentially damped
modes. In order to derive a nonlinear effective equation for the diffusive
modes alone, namely the so called phase diffusion equation, we introduce
polar coordinates
A(X, T ) = r(X, T )eiφ(X,T )
in the real GL equation and obtain
∂T r = ∂X2 r + r − (∂ φ)2 r − r 3 ,
X
(14.49) 2 2(∂X r)(∂X φ)
∂T φ = ∂X φ + r .
We introduce the deviation (s, φ) from the equilibrium (r, φ) = (1, 0) by
setting r = 1 + s and find
∂T s = ∂X 2 s − 2s − (∂ φ)2 − (∂ φ)2 s − 3s2 − s3 ,
X X
(14.50) 2 2(∂X s)(∂X φ)
∂T φ = ∂X φ + 1+s .
We replace the equation for the phase φ by an equation for the local wave
number ψ = ∂X φ and obtain
2 s − 2s − ψ 2 − ψ 2 s − 3s2 − s3 ,
∂T s = ∂X
(14.51) 2 (∂X s)ψ
∂T ψ = ∂X ψ + 2∂X 1+s .
14.4. Phase diffusion equations 531
with τ = δ 2 T and ξ = δX. After neglecting the terms of order O(δ 2 ) for
small ψ̌, the implicit function theorem yields that the first equation
(14.55) ∂τ ψ̌ = ∂ξ2
h(ψ̌, q)
where h(ψ̌, q) = h(ψ̌ + q). For each q, let k → μ1,2 (k, q) denote the smooth
curves of eigenvalues corresponding to the Fourier wave numbers k for the
linearization of (14.49) around (r, φ) = ( 1 − q 2 , qX). In particular, let
μ1 (k, q) = −k 2 − (1 − q 2 ) + (1 − q 2 )2 + 4q 2 k 2
denote the critical curve for which μ1 (0, q) = 0. Then the phase-diffusion
equation (14.54) must give at lowest order a linear diffusion equation with
diffusion coefficient
Theorem 14.4.1. Let 2 ≤ m ≤ n − 2. Then there exists a Cψ∗ > 0 such that
the following is true. Let ψ ∗ be a solution of the phase diffusion equation
(14.54) with
sup ψ ∗ (τ )Hul ∗
n ≤ C ,
ψ
τ ∈[0,τ0 ]
Then we obtain
|A(X, T ) − Aapp (X, T )|
X
≤(1 + š(δX, δ T )) exp i
2
ψ̌(δX , δ 2 T ) dX
0
X
∗
− (1 + s (δX, δ T )) exp i
2
ψ ∗ (δX , δ 2 T ) dX
0
X
≤(1 + š(δX, δ T )) exp i
2
ψ̌(δX , δ 2 T ) dX
0
X
− (1 + š(δX, δ T )) exp i
2
ψ ∗ (δX , δ 2 T ) dX
0
X
+ (1 + š(δX, δ 2 T )) exp i ψ ∗ (δX , δ 2 T ) dX
0
X
− (1 + s∗ (δX, δ 2 T )) exp i ψ ∗ (δX , δ 2 T ) dX
0
14.4. Phase diffusion equations 533
X
≤|1 + š(δX, δ T ))| exp i
2
ψ̌(δX , δ 2 T ) dX
0
X
− exp i ψ ∗ (δX , δ 2 T ) dX
0
∗
+ |s (δX, δ T )) − š(δX, δ 2 T ))|
2
X
≤C |ψ̌(δX , δ 2 T ) − ψ ∗ (δX , δ 2 T )| dX + Cδ 2
0
X
≤ Cδ 2 dX + Cδ 2 ≤ Cδ 2 (1 + |X|)
0
using the approximation result of Theorem 14.4.1.
By adding higher order terms in (s∗ , ψ ∗ ), the approximation can be
improved to hold uniformly on space intervals with length of order O(δ −p )
with p ≥ 0 arbitrary.
š = s∗ (ψ̌). Inserting this into the second equation gives the conservation
law
(14.60) ∂τ ψ̌ = ∂ξ (−2βs∗ (ψ̌) − αψ̌ 2 − βs∗ (ψ̌)2 ) = ∂ξ h(ψ̌)
where h : R → R is smooth. Estimates that (14.60) makes correct pre-
dictions about the dynamics of the complex GL equation can be found in
[MS04a].
p−1
up H s ≤CuH [s/2]+1 uH ,
s
p−2
up W s,1 ≤CuW [s/2]+1,∞ uH [s/2]+1 uH .
s
t/2
≤(1 + t) d/2
C(t − τ )−(d/2) u(τ )W
p−2
s∞ ,∞ u(τ )H s2 dτ
2
0
t
p−1
s∞ ,∞ u(τ )H s dτ
d/2
+ (1 + τ ) Cu(τ )W
t/2
− d
t/2
t 2
(1 + τ )− 2 (p−2) dτ b(t)p−2 c(t)2
d
≤(1 + t) d/2
C
0 2
t d
d/2 t − 2 (p−1)
+ (1 + t) C 1+ dτ b(t)p−1 c(t)
t/2 2
≤C2 b(t)p−2 c(t)2 + b(t)p−1 c(t) ,
where we used Sobolev’s embedding theorem for s2 > s∞ + d2 and for the
existence of the integrals we assume d2 (p − 2) > 1 and d2 (p − 1) ≥ d2 + 1. For
t ∈ (0, 1] we have
t 1
(1+t)d/2 e i(t−τ )Λ
N (u(τ )) dτ
s ,∞ ≤ 2 d/2
C ei(t−τ )Λ N (u(τ ))H s2 dτ
0 W ∞ 0
1
≤2d/2 C N (u(τ ))H s2 dτ ≤ C3 b(t)p−1 c(t)
0
Therefore we find
c(t) ≤c(0) + C1 b(t)p−1 c(t),
b(t) ≤C(c(0) + b(0)) + (C2 + C3 )(b(t)p−1 c(t) + b(t)p−2 c(t)).
Hence y(t) = b(t) + c(t) satisfies y(t) ≤ Cy(0) + Cy(t)p . It follows as above
that supt≥0 y(t) ≤ 2Cy(0).
Theorem 14.5.1. Consider (14.61) with p > 2 + d2 , s2 > s∞ + d2 , s1 >
s∞ + d + 2, and s∞ ≥ [ s22 ] + 1. Then there exists δ > 0 and C > 0 such that
the following holds. If
U0 H s2 + U0 W s∞ ,∞ + U0 W s1 ,1 ≤ δ,
then the associated solution U with U |t=0 = U0 exists globally and satisfies
u(t)W s∞ ,∞ ≤ C(1 + t)− 2 .
d
u(t)H s2 ≤ C and
Remark 14.5.2. Although (14.61) can be solved forward and backward in
time, Theorem 14.5.1 has a direction in time. The reason is the growth
of the W s1 ,1 -norm as time evolves. Since the W s1 ,1 -norm appears in the
assumptions of Theorem 14.5.1 the application of the Theorem cannot be
iterated.
Remark 14.5.3. The condition p > 1+2/d was used in the estimates for the
H s2 -norm. The stronger condition p > 2+2/d was only used in the estimates
for the W s∞ ,∞ -norm. By a normal form transform [Sha85] the nonlinearity
538 14. Diffusive stability
Exercises
14.1. For u(x, t) = ct−1/2 e−x
2
/(4t)
and t → ∞, prove the estimates
−3/2
sup |∂t u| ∼ t , sup |∂x2 u| ∼ t−3/2 , and sup |up | ∼ t−p/2 .
x∈R x∈R x∈R
14.2. Compute the first four Hermite polynomials via the inverse Fourier transform
j = (ik)j e−k for j = 0, 1, 2, 3, cf. Exercise 7.19.
2
of u
14.3. Prove the estimates (14.11) and (14.12).
14.4. Transfer the Lyapunov approach from §14.1.4 to 2D and 3D spatial domains.
14.5. Consider ∂t u = −∂x4 u, with x ∈ R, t ≥ 0, u(x, t) ≥ 0. Make an ansatz
1 x
u(x, t) = t1/4 v( t1/4 ) for finding self-similar solutions. Show that v satifies
1 1
−v (4) + ξv + v = 0.
4 4
Show that v(k) = e−k satisfies the Fourier transformed equation. Discuss the
4
−k4 ikξ
smoothness
and decay of the function G(ξ) = c R e e dk where c ∈ R is chosen
such that G(ξ)dξ = 1.
Bibliography
541
542 Bibliography
[BN85] N. Bekki and K. Nozaki, Formations of spatial patterns and holes in the
generalized Ginzburg-Landau equation, Physics Letters A 110 (1985), no. 3,
133–135.
[BNSZ14] M. Beck, T. T. Nguyen, B. Sandstede, and K. Zumbrun, Nonlinear stability
of source defects in the complex Ginzburg-Landau equation, Nonlinearity 27
(2014), no. 4, 739–786.
[Bou77] M. B. Boussinesq, Essai sur la théorie des eaux courantes, Inst. France (séries
2) 23 (1877), 1–680.
[Bou99] J. Bourgain, Global solutions of nonlinear Schrödinger equations, American
Mathematical Society, Providence, RI, 1999.
[BR13] D. Blömker and M. Romito, Local existence and uniqueness for a two-
dimensional surface growth equation with space-time white noise, Stochastic
Anal. Appl. 31 (2013), no. 6, 1049–1076.
[Bra83] M. Bramson, Convergence of solutions of the Kolmogorov equation to travel-
ling waves, Mem. Amer. Math. Soc. 44 (1983), no. 285.
[Bri13] Th. J. Bridges, A universal form for the emergence of the Korteweg-de Vries
equation, Proc. R. Soc. Lond. Ser. A Math. Phys. Eng. Sci. 469 (2013),
no. 2153, 20120707, 15.
[BS07] J. Bitzer and G. Schneider, Approximation and attractivity properties of the
degenerated Ginzburg-Landau equation, J. Math. Anal. Appl. 331 (2007),
no. 2, 743–778.
[BSTU06] K. Busch, G. Schneider, L. Tkeshelashvili, and H. Uecker, Justification of the
Nonlinear Schrödinger equation in spatially periodic media, ZAMP 57 (2006),
1–35.
[BV90] A. V. Babin and M. I. Vishik, Attractors of partial differential equations in
an unbounded domain, Proceedings of the Royal Society of Edinburgh 116A
(1990), 221–243.
[BV92] , Attractors of evolution equations, North Holland, 1992.
[BX08] W. Brock and A. Xepapadeas, Diffusion-induced instability and pattern for-
mation in infinite horizon recursive optimal control, Journal of Economic
Dynamics and Control 32 (2008), no. 9, 2745–2787.
[BX10] , Pattern formation, spatial externalities and regulation in coupled
economic–ecological systems, Journal of Environmental Economics and Man-
agement 59 (2010), no. 2, 149–164.
[Car66] L. Carleson, On convergence and growth of partial sumas of Fourier series,
Acta Math. 116 (1966), 135–157.
[Car81] J. Carr, Applications of centre manifold theory, Applied Mathematical Sci-
ences, Vol. 35. New York Heidelberg Berlin: Springer Verlag. XII. 142 p. DM
32.00; $ 14.60 (1981)., 1981.
[CBCSU08] M. Chirilus-Bruckner, C. Chong, G. Schneider, and H. Uecker, Separation
of internal and interaction dynamics for NLS-described wave packets with
different carrier waves, J. Math. Anal. Appl. 347 (2008), no. 1, 304–314.
[CBS12] M. Chirilus-Bruckner and G. Schneider, Detection of standing pulses in peri-
odic media by pulse interaction, J. Differential Equations 253 (2012), no. 7,
2161–2190.
[CBSU07] M. Chirilus-Bruckner, G. Schneider, and H. Uecker, On the interaction of
NLS-described modulating pulses with different carrier waves, Math. Methods
Appl. Sci. 30 (2007), no. 15, 1965–1978.
Bibliography 545
(2000), no. 11-12, 745–753, Special issue on the occasion of the 125th anniver-
sary of the birth of Ludwig Prandtl.
[ES02] , Non-linear stability of modulated fronts for the Swift-Hohenberg equa-
tion, Comm. Math. Phys. 225 (2002), no. 2, 361–397.
[Eva98] L. C. Evans, Partial differential equations, Graduate Studies in Mathematics,
vol. 19, American Mathematical Society, Providence, RI, 1998.
[EvH81] W. Eckhaus and A. van Harten, The inverse scattering transformation and
the theory of solitons. An introduction, North-Holland Mathematics Studies,
vol. 50, North-Holland Publishing Co., Amsterdam, 1981.
[EW91] J.-P. Eckmann and C. E. Wayne, Propagating fronts and the center manifold
theorem, Comm. Math. Phys. 136 (1991), no. 2, 285–307.
[EW94] , The nonlinear stability of front solutions for parabolic partial differ-
ential equations, Comm. Math. Phys. 136 (1994), 285–307.
[EWW97] J.-P. Eckmann, C. E. Wayne, and P. Wittwer, Geometric stability analysis of
periodic solutions of the Swift-Hohenberg equation, Comm. Math. Phys. 190
(1997), 173–211.
[Fef06] C. L. Fefferman, Existence and smoothness of the Navier-Stokes equation,
The millennium prize problems, Clay Math. Inst., Cambridge, MA, 2006,
pp. 57–67.
[Fen79] N. Fenichel, Geometric singular perturbation theory for ordinary differential
equations, J. Differential Equations 31 (1979), no. 1, 53–98.
[FH54] K. O. Friedrichs and D. H. Hyers, The existence of solitary waves, Commun.
Pure Appl. Math. 7 (1954), 517–550.
[Fib15] G. Fibich, The nonlinear Schrödinger equation. singular solutions and optical
collapse, Applied Mathematical Sciences, vol. 192, Springer, Cham, 2015.
[Fis37] R. Fisher, The advance of advantageous genes, Ann. Eugenics 7 (1937), 355–
369.
[Fit69] R. FitzHugh, Mathematical models of excitation and propagation in nerve,
H.P. Schwan, ed., Biological Engineering, 1969, pp. 1–85.
[FL96] A. S. Fokas and Q. M. Liu, Asymptotic integrability of water waves, Phys.
Rev. Lett. 77 (1996), no. 12, 2347–2351.
[FLS64] R. P. Feynman, R. B. Leighton, and M. Sands, The Feynman lectures on
physics. Vol. 2: Mainly electromagnetism and matter, Addison-Wesley Pub-
lishing Co., Inc., Reading, Mass.-London, 1964.
[FMW92] W. E. Fitzgibbon, J. J. Morgan, and S. J. Waggoner, Weakly coupled semi-
linear parabolic evolution systems, Ann. Mat. Pura Appl. (4) 161 (1992),
213–229.
[Fow97] A. C. Fowler, Mathematical models in the applied sciences, Cambridge Uni-
versity Press, 1997.
[FPU55] E. Fermi, J. Pasta, and S. M. Ulam, Studies in nonlinear problems, Technical
Report Los Alamos Sci. Lab.., 1955.
[FRMT01] C. Foias, R. Rosa, O. Manley, and R. Temam, Navier-Stokes equations and
turbulence, Cambridge: Cambridge University Press, 2001.
[Gal94] Th. Gallay, Local stability of critical fronts in nonlinear parabolic partial dif-
ferential equations, Nonlinearity 7 (1994), no. 3, 741–764.
[Gal11] G. P. Galdi, An introduction to the mathematical theory of the Navier-Stokes
equations. Steady-state problems, second ed., Springer Monographs in Math-
ematics, Springer, New York, 2011.
550 Bibliography
[HI11] M. Haragus and G. Iooss, Local bifurcations, center manifolds, and nor-
mal forms in infinite-dimensional dynamical systems, Universitext. London:
Springer; Les Ulis: EDP Sciences. xi, 329 p. , 2011.
[HK91] J. K. Hale and H. Kocak, Dynamics and bifurcations, Springer-Verlag, 1991.
[HN01] F. Hamel and N. Nadirashvili, Travelling fronts and entire solutions of the
Fisher-KPP equation in RN , Arch. Ration. Mech. Anal. 157 (2001), no. 2,
91–163.
[HNM14] S. Hata, H. Nakao, and A. S. Mikhailov, Sufficient conditions for wave insta-
bility in three-component reaction-diffusion systems, Prog. Theor. Exp. Phys.
(2014), 013A01.
[HO98] Y. Hayase and T. Ohta, Sierpinski gasket in a reaction–diffusion system,
Phys. Rev. Let. 81 (1998), 1726.
[HO00] , Self-replicating pulses and Sierpinski gaskets in excitable media,
Phys. Rev. E 62 (2000), 5998–6003.
[Hör83] L. Hörmander, The analysis of linear partial differential operators. I. Distri-
bution theory and Fourier analysis, Grundlehren der Mathematischen Wis-
senschaften, vol. 256, Springer-Verlag, Berlin, 1983.
[Hor03] D. Horstmann, From 1970 until present: the Keller-Segel model in chemotaxis
and its consequences. I, Jahresber. Deutsch. Math.-Verein. 105 (2003), no. 3,
103–165.
[Hor04] , From 1970 until present: the Keller-Segel model in chemotaxis and
its consequences. II, Jahresber. Deutsch. Math.-Verein. 106 (2004), no. 2,
51–69.
[Hoy06] R. B. Hoyle, Pattern formation. An introduction to methods, Cambridge Uni-
versity Press, Cambridge, 2006.
[HPT00] J. K. Hale, L. A. Peletier, and W. C. Troy, Exact homoclinic and heteroclinic
solutions of the Gray-Scott model for autocatalysis, SIAM J. Appl. Math. 61
(2000), no. 1, 102–130.
[HR90] J. Hale and G. Raugel, Lower semicontinuity of the attractor for a singularly
perturbed hyperbolic equation, J. Dyn. Differ. Equations 2 (1990), no. 1, 19–67.
[HS12] M. Haragus and A. Scheel, Dislocations in an anisotropic Swift-Hohenberg
equation, Comm. Math. Phys. 315 (2012), no. 2, 311–335.
[HSD04] M. W. Hirsch, S. Smale, and R. L. Devaney, Differential equations, dynamical
systems, and an introduction to chaos, second ed., vol. 60, Elsevier/Academic
Press, Amsterdam, 2004.
[HSU12] T. Häcker, G. Schneider, and H. Uecker, Self-similar decay to the marginally
stable ground state in a model for film flow over inclined wavy bottoms, Electr.
J. Diff. Eq. 61 (2012), 1–51.
[HSZ11] T. Häcker, G. Schneider, and D. Zimmermann, Justification of the Ginzburg-
Landau approximation in case of marginally stable long waves, J. Nonlinear
Sci. 21 (2011), no. 1, 93–113.
[HT00] K.-H. Hoffmann and Qi Tang, Ginzburg-Landau Phase Transition The-
ory and Superconductivity, International Series of Numerical Mathematics,
Birkhäuser, 2000.
[HZ11] H. Hofer and E. Zehnder, Symplectic invariants and Hamiltonian dynamics,
Modern Birkhäuser Classics, Birkhäuser Verlag, Basel, 2011, Reprint of the
1994 edition.
Bibliography 553
[IA98] G. Iooss and M. Adelmeyer, Topics in bifurcation theory and applications. 2nd
ed., Advanced Series in Nonlinear Dynamics. 3. Singapore: World Scientific.
viii, 186 p., 1998.
[Igu01] T. Iguchi, Well-posedness of the initial value problem for capillary-gravity
waves, Funkcial. Ekvac. 44 (2001), no. 2, 219–241.
[IK90] G. Iooss and K. Kirchgaessner, Bifurcation d’ondes solitaires en présence
d’une faible tension superficielle. (Bifurcation of solitary waves, for a low
surface tension), C. R. Acad. Sci., Paris, Ser. I 311 (1990), 265–268 (French).
[IM91] G. Iooss and A. Mielke, Bifurcating time-periodic solutions of Navier-Stokes
equations in infinite cylinders, J. Nonlinear Sci. 1 (1991), no. 1, 107–146.
[Ioo95] G. Iooss, Capillary-gravity water-waves problem as a dynamical system, A.
Mielke (ed.) et al., Structure and dynamics of nonlinear waves in fluids. Pro-
ceedings of the IUTAM/ ISIMM symposium, Hannover, Germany, August
17-20, 1994. Singapore: World Scientific. Adv. Ser. Nonlinear Dyn. 7, 42-57
(1995), 1995.
[Ioo98] , Travelling water-waves, as a paradigm for bifurcations in reversible
infinite dimensional “dynamical” systems, Doc. Math., J. DMV (1998), 611–
622.
[IS16] S. Iyer and B. Sandstede, Mixing in Reaction-Diffusion Systems: Large Phase
Offsets, arXiv:1610.06527, 2016.
[Jam03] G. James, Centre manifold reduction for quasilinear discrete systems, J. Non-
linear Sci. 13 (2003), no. 1, 27–63.
[JGK93] C. K. R. T. Jones, R. Gardner, and T. Kapitula, Stability of travelling waves
for non-convex scalar viscous conservation laws, Commun. Pure Appl. Math.
46 (1993), no. 4, 505–526.
[JK84] F. John and S. Klainerman., Almost global existence to nonlinear wave equa-
tions in three space dimensions, Commun. Pure Appl. Math. 37 (1984), 443–
455.
[JNRZ14] M. A. Johnson, P. Noble, L. M. Rodrigues, and K. Zumbrun, Behavior of
periodic solutions of viscous conservation laws under localized and nonlocalized
perturbations, Invent. Math. 197 (2014), no. 1, 115–213.
[Joh91] F. John, Partial differential equations, Springer, 1991.
[Joh02] R. S. Johnson, Camassa-Holm, Korteweg-de Vries and related models for
water waves, J. Fluid Mech. 455 (2002), 63–82.
[Jon84] C. K. R. T. Jones, Stability of the travelling wave solution of the FitzHugh-
Nagumo system, Trans. Amer. Math. Soc. 286 (1984), no. 2, 431–469.
[JR94] B. Jakobsen and F. Rosendahl, The Sleipner Platform Accident, Structural
Engineering International 4 (1994), no. 3, 190–193.
[JZ10] M. A. Johnson and K. Zumbrun, Nonlinear stability of periodic traveling wave
solutions of systems of viscous conservation laws in the generic case, J. Differ.
Equations 249 (2010), no. 5, 1213–1240.
[JZ11] , Nonlinear stability of periodic traveling-wave solutions of viscous con-
servation laws in dimensions one and two, SIAM J. Appl. Dyn. Syst. 10
(2011), no. 1, 189–211.
[Kal88] L. A. Kalyakin, Asymptotic decay of a one-dimensional wavepacket in a non-
linear dispersive medium, Math. USSR Sbornik 60 (1988), no. 2, 457–483.
[Kal89] , Long wave asymptotics. Integrable equations as asymptotic limits of
non- linear systems, Russ. Math. Surv. 44 (1989), no. 1, 3–42.
554 Bibliography
[Kap94a] T. Kapitula, On the nonlinear stability of plane waves for the Ginzburg-
Landau equation, Commun. Pure Appl. Math. 47 (1994), no. 6, 831–841.
[Kap94b] , On the stability of travelling waves in weighted L∞ spaces, J. Differ.
Equations 112 (1994), no. 1, 179–215.
[Kat75] T. Kato, Quasi-linear equations of evolution, with applications to partial dif-
ferential equations, Spectral theory and differential equations (Proc. Sympos.,
Dundee, 1974; dedicated to Konrad Jörgens), Springer, Berlin, 1975, pp. 25–
70. Lecture Notes in Math., Vol. 448.
[Kat81] , The Cauchy problem for the Korteweg-de Vries equation, Nonlinear
partial differential equations and their applications, Res. Notes in Math.,
vol. 53, Pitman, 1981, pp. 293–307.
[Kat95] , Perturbation theory for linear operators. Reprint of the corr. print.
of the 2nd ed. 1980, reprint of the corr. print. of the 2nd ed. 1980 ed., Berlin:
Springer-Verlag, 1995.
[Kaw72] T. Kawahara, Oscillatory solitary waves in dispersive media, J. Phys. Soc.
Japan (1972), 260–264.
[KC13] G. Kozyreff and S. J. Chapman, Analytical results for Front Pinning between
an Hexagonal Pattern and a Uniform State in Pattern-Formation Systems,
Phys. Rev. Letters 111(5) (2013), 054501.
[KDL90] E. Knobloch and J. De Luca, Amplitude equations for travelling wave convec-
tion, Nonlinearity 3 (1990), no. 4, 975–980.
[KdV95] D. J. Korteweg and G. de Vries, On the change of form of long waves ad-
vancing in the rectangular canal and a new type of long stationary waves,
Phil.Magaz. 39 (1895), 422–443.
[Kee88] J. P. Keener, Principles of applied mathematics, Addison-Wesley, 1988.
[Kel77] H. B. Keller, Numerical solution of bifurcation and nonlinear eigenvalue prob-
lems, Application of bifurcation theory, Proc. adv. Semin., Madison/Wis.
1976, 359-384, 1977.
[Kel93] S. H. Kellert, In the wake of chaos. Unpredictable order in dynamical sys-
tems, Science and its Conceptual Foundations, University of Chicago Press,
Chicago, IL, 1993.
[KH97] A. Katok and B. Hasselblatt, Introduction to the modern theory of dynamical
systems, Cambridge University Press, 1997.
[Kir82] K. Kirchgaessner, Wave-solutions of reversible systems and applications, J.
Differ. Equations 45 (1982), 113–127.
[KN79] T. Kano and T. Nishida, Sur les ondes de surface de l’eau avec une justifi-
cation mathématique des équations des ondes en eau peu profonde, J. Math.
Kyoto Univ. 19 (1979), no. 2, 335–370.
[KN86] , A mathematical justification for Korteweg-de Vries equation and
Boussinesq equation of water surface waves, Osaka J. Math. 23 (1986), no. 2,
389–413.
[Kno92] E. Knobloch, Nonlocal amplitude equations, Pattern formation in complex
dissipative systems (Kitakyushu, 1991), World Sci. Publ., River Edge, NJ,
1992, pp. 263–274.
[Koc15] H. Koch, Nonlinear dispersive equations, H. Koch, D. Tartaru, M. Visan, eds,
Dispersive Equations and Nonlinear Waves, Birkhäuser, 2015.
[Kol27] A. Kolmogoroff, Sur la convergence des séries de fonctions orthogonales,
Math. Z. 26 (1927), no. 1, 432–441.
Bibliography 555
[Ler34] J. Leray, Sur le mouvement d’un liquide visqueux emplissant l’espace, Acta
Math. 63 (1934), no. 1, 193–248.
[LeV92] R. J. LeVeque, Numerical methods for conservation laws, second ed.,
Birkhäuser Verlag, Basel, 1992.
[LL91] L. D. Landau and E. M. Lifschitz, Lehrbuch der theoretischen Physik
(“Landau-Lifschitz”). Band VI, fifth ed., Akademie-Verlag, Berlin, 1991, Hy-
drodynamik. [Hydrodynamics], Translated from the Russian by Wolfgang
Weller and Adolf Kühnel, Translation edited by Weller and with a foreword
by Weller and P. Ziesche.
[LLS13] D. Lannes, F. Linares, and J.-C. Saut, The Cauchy problem for the Euler-
Poisson system and derivation of the Zakharov-Kuznetsov equation, Studies
in phase space analysis with applications to PDEs. In part selected papers
based on the presentations at a meeting, Bertinoro, Italy, September 2011,
New York, NY: Birkhäuser/Springer, 2013, pp. 181–213.
[Log15a] J. D. Logan, Applied partial differential equations, Undergraduate Texts in
Mathematics, Springer, Cham, 2015, Third edition.
[Log15b] , A first course in differential equations, Undergraduate Texts in Math-
ematics, Springer, Cham, 2015, Third edition.
[Lor63] E. N. Lorenz, Deterministic non-periodic flow, J. Atmos. Sci. (1963).
[LP09] F. Linares and G. Ponce, Introduction to nonlinear dispersive equations, New
York, NY: Springer, 2009.
[LSAC08] D. J. B. Lloyd, B. Sandstede, D. Avitabile, and A. R. Champneys, Localized
hexagon patterns of the planar Swift-Hohenberg equation, SIAM J. Appl. Dyn.
Syst. 7 (2008), no. 3, 1049–1100.
[Lun95] A. Lunardi, Analytic semigroups and optimal regularity in parabolic prob-
lems, Progress in Nonlinear Differential Equations and their Applications,
16, Birkhäuser Verlag, Basel, 1995.
[Lyn04] S. Lynch, Dynamical systems with applications using MATLAB, Birkhäuser,
2004.
[Man91] B. B. Mandelbrot, Die fraktale Geometrie der Natur, Birkhaeuser, 1991.
[Man92] P. Manneville, Dissipative structures and weak turbulence, Academic Press,
1992.
[Mar55] V. A. Marchenko, On the reconstruction of the potential energy from phases of
the scattered waves, Dokl. Akad. Nauk SSSR 104 (1955), 695–698 (Russian).
[Mar07] P. A. Markowich, Applied partial differential equations: a visual approach,
Springer, Berlin, 2007.
[Maz11] Vl. Maz’ya, Sobolev spaces with applications to elliptic partial differen-
tial equations, Grundlehren der Mathematischen Wissenschaften, vol. 342,
Springer, Heidelberg, 2011.
[MB02] A. J. Majda and A. L. Bertozzi, Vorticity and incompressible flow, Cam-
bridge Texts in Applied Mathematics, vol. 27, Cambridge University Press,
Cambridge, 2002.
[MBK14] W. W. Mohammed, D. Blömker, and K. Klepel, Multi-scale analysis of SPDEs
with degenerate additive noise, J. Evol. Equ. 14 (2014), no. 1, 273–298.
[Mel98] I. Melbourne, Derivation of the time-dependent Ginzburg-Landau equation on
the line, J. Nonlinear Sci. 8 (1998), no. 1, 1–15.
Bibliography 557
[MU16] K. Matthies and H. Uecker, Low regularity justification results for envelope ap-
proximations of nonlinear wave packets in periodic media, Asymptotic Anal.
99 (2016), no. 1-2, 53–65.
[Mur89] J. D. Murray, Mathematical Biology, Springer, Berlin, 1989.
[MY60] L. Markus and H. Yamabe, Global stability criteria for differential systems,
Osaka Math. J. 12 (1960), 305–317.
[Nal74] V. I. Nalimov, The Cauchy-Poisson problem, Dinamika Splošn. Sredy 254
(1974), no. Vyp. 18 Dinamika Zidkost. so Svobod. Granicami, 104–210.
[New85] A. C. Newell, Solitons in mathematics and physics, vol. 48, CBMS-NSF Reg.
Conf. Ser. Appl. Math., 1985.
[New16] , Short pulse evolution equation, Laser filamentation, CRM Ser. Math.
Phys., Springer, Cham, 2016, pp. 1–17.
[Nis02] Y. Nishiura, Far-from-equilibrium dynamics, Translations of Mathematical
Monographs, vol. 209, American Mathematical Society, Providence, RI, 2002.
[Nti76] A. A. Ntinos, Lengths of instability intervals of second order periodic differ-
ential equations, Quart. J. Math. Oxford (2) 27 (1976), no. 107, 387–394.
[NTYU07] Y. Nishiura, T. Teramoto, X. Yuan, and Kei-Ichi Ueda, Dynamics of traveling
pulses in heterogeneous media, Chaos 17 (2007), no. 3, 037104, 21.
[NW69] A. C. Newell and J. A. Whitehead, Finite bandwidth, finite amplitude con-
vection, J. Fluid Mech. 38 (1969), 279–303.
[OD08] I. Oprea and G. Dangelmayr, Dynamics and bifurcations in the weak elec-
trolyte model for electroconvection of nematic liquid crystals: a Ginzburg-
Landau approach, Eur. J. Mech., B, Fluids 27 (2008), no. 6, 726–749.
[Olv14] P. J. Olver, Introduction to partial differential equations, Undergraduate Texts
in Mathematics, Springer, Cham, 2014.
[Osb10] A. R. Osborne, Nonlinear ocean waves and the inverse scattering transform,
International Geophysics Series 97. Amsterdam: Elsevier/Academic Press.
xxvi, 2010.
[Paz83] A. Pazy, Semigroups of linear operators and applications to partial differen-
tial equations, Applied Mathematical Sciences, vol. 44, Springer-Verlag, New
York, 1983.
[Pel11] D. E. Pelinovsky, Localization in periodic potentials. From Schrödinger oper-
ators to the Gross-Pitaevskii equation, London Mathematical Society Lecture
Note Series 390. Cambridge: Cambridge University Press. x, 398 p., 2011.
[Pis06] L. M. Pismen, Patterns and interfaces in dissipative dynamics, Springer Series
in Synergetics, Springer-Verlag, Berlin, 2006.
[PJS92] H.-O. Peitgen, H. Jürgens, and D. Saupe, Fractals for the classroom. Part 2:
Complex systems and the Mandelbrot set, Springer, 1992.
[PL68] I. Prigogine and R. Lefever, Symmetry breaking bifurcations in dissipative
systems II, J. Chem. Phys. 48 (1968), 1665–1700.
[Poi57] H. Poincaré, Les methodes nouvelles de la mecanique celeste, Dover Publica-
tions, 1957.
[Pom86] Y. Pomeau, Front motion, metastability and subcritical bifurcations in hydro-
dynamics, Physica D 23 (1986), 3–11.
[PS07] D. Pelinovsky and G. Schneider, Justification of the coupled-mode approxi-
mation for a nonlinear elliptic problem with a periodic potential, Appl. Anal.
86 (2007), no. 8, 1017–1036.
Bibliography 559
[Sch11] , Justification of the NLS approximation for the KdV equation using
the Miura transformation, Adv. Math. Phys. (2011), Art. ID 854719, 4.
[Sch15] R. Schnaubelt, Center manifolds and attractivity for quasilinear parabolic
problems with fully nonlinear dynamical boundary conditions, Discrete Con-
tin. Dyn. Syst. 35 (2015), no. 3, 1193–1230.
[SD98] R. Schielen and A. Doelman, Modulation equations for spatially periodic sys-
tems: derivation and solutions, SIAM J. Appl. Math. 58 (1998), no. 6, 1901–
1930.
[Seg69] L. A. Segel, Distant side-walls cause slow amplitude modulation of cellular
convection, Journal of Fluid Mechanics 38(1) (1969), 203–224.
[Ser99] D. Serre, Systems of conservation laws. I: Hyperbolicity, entropies, shock
waves, Cambridge: Cambridge University Press. xxii, 263 p., 1999.
[Ser00] , Systems of conservation laws 2. Geometric structures, oscillations,
and initial-boundary value problems, Cambridge: Cambridge University
Press. xi, 269 p., 2000.
[Ser15] S. Serfaty, Coulomb gases and Ginzburg-Landau vortices, Zürich: European
Mathematical Society (EMS), 2015.
[Sey10] R. Seydel, Practical bifurcation and stability analysis. 3rd ed., Springer, 2010.
[SGKM95] A. A. Samarskii, V. Galaktionov, S. Kurdyumov, and A. Mikhailov, Blow-up
in quasilinear parabolic equations, De Gruyter, 1995.
[SH77] J. Swift and P. C. Hohenberg, Hydrodynamic fluctuations at the convective
instability, Physical Review A 15 (1977), no. 1, 319–328.
[SH94a] P. J. Schmid and D. S. Henningson, Optimal energy density growth in Hagen-
Poiseuille flow, J. Fluid Mech. 277 (1994), 197–225.
[SH94b] Y. Shi and J. E. Hearst, The Kirchhoff elastic rod, the nonlinear Schrödinger
equation, and DNA supercoiling, J. Chem. Phys. 101 (1994), no. 6, 5186–
5200.
[SH96] A. M. Stuart and A. R. Humphries, Dynamical systems and numerical anal-
ysis, Cambridge: Cambridge University Press, 1996.
[Sha85] J. Shatah, Normal forms and quadratic nonlinear Klein-Gordon equations,
Comm. Pure Appl. Math. 38 (1985), no. 5, 685–696.
[Sha10] J. Shatah, Space-time resonances, Q. Appl. Math. 68 (2010), no. 1, 161–167.
[She97] A. Shepeleva, On the validity of the degenerate Ginzburg-Landau equation,
Math. Methods Appl. Sci. 20 (1997), no. 14, 1239–1256.
[Shi76] M. Shinbrot, The initial value problem for surface waves under gravity. I. The
simplest case, Indiana Univ. Math. J. 25 (1976), no. 3, 281–300.
[Sil65] L. P. Sil’nikov, A case of the existence of a denumerable set of periodic mo-
tions, Dokl. Akad. Nauk SSSR 160 (1965), 558–561.
[Siv77] G. Sivashinsky, Nonlinear analysis of hydrodynamic instability in laminar
flames. I - Derivation of basic equations, Acta Astronautica 4 (1977), 1177–
1206.
[Smo94] J. Smoller, Shock waves and reaction-diffusion equations, Springer, New York,
1994.
[Spa82] C. Sparrow, The Lorenz equations: bifurcations, chaos, and strange attractors,
Applied Mathematical Sciences, 41. Springer-Verlag, 1982.
[SR89] X. Saint Raymond, A simple Nash-Moser implicit function theorem, Enseign.
Math. (2) 35 (1989), no. 3-4, 217–226.
562 Bibliography
[SR10] P. L. Sachdev and Ch. Srinivasa Rao, Large time asymptotics for solutions of
nonlinear partial differential equations, Springer Monographs in Mathematics,
Springer, New York, 2010.
[SS99a] B. Sandstede and A. Scheel, Essential instability of pulses and bifurcations to
modulated travelling waves, Proc. R. Soc. Edinb., Sect. A, Math. 129 (1999),
no. 6, 1263–1290.
[SS99b] C. Sulem and P.-L. Sulem, The nonlinear Schrödinger equation, Springer-
Verlag, New York, 1999.
[SS00] B. Sandstede and A. Scheel, Absolute and convective instabilities of waves on
unbounded and large bounded domains, Phys. D 145 (2000), no. 3-4, 233–277.
[SS07] E. Sandier and S. Serfaty, Vortices in the magnetic Ginzburg-Landau model,
Basel: Birkhäuser, 2007.
[SSSU12] B. Sandstede, A. Scheel, G. Schneider, and H. Uecker, Diffusive mixing of
periodic wave trains in reaction-diffusion systems, J. Differ. Equations 252
(2012), no. 5, 3541–3574.
[SSW99] B. Sandstede, A. Scheel, and C. Wulff, Bifurcations and dynamics of spiral
waves, J. Nonlinear Sci. 9 (1999), no. 4, 439–478.
[SSZ15] G. Schneider, Danish Ali Sunny, and D. Zimmermann, The NLS approxi-
mation makes wrong predictions for the water wave problem in case of small
surface tension and spatially periodic boundary conditions, J. Dyn. Differ.
Equations 27 (2015), no. 3-4, 1077–1099.
[Ste93] E. M. Stein, Harmonic analysis: Real-variable methods, orthogonality, and
oscillatory integrals. With the assistance of Timothy S. Murphy, Princeton,
NJ: Princeton University Press, 1993.
[STE+ 10] I. Staude, M. Thiel, S. Essig, C. Wolff, K. Busch, G. Von Freymann, and
M. Wegener, Fabrication and characterization of silicon woodpile photonic
crystals with a complete bandgap at telecom wavelengths, Optics letters 35
(2010), no. 7, 1094–1096.
[Str26] D. J. Struik, Détermination rigoureuse des ondes irrotationelles périodiques
dans un canal à profondeur finie, Math. Ann. 95 (1926), 595–634 (French).
[Str89] W. A. Strauss, Nonlinear wave equations. Expository lectures from the CBMS
regional conference held at George Mason University, January 16-22, 1989,
Providence, RI: American Mathematical Society, 1989.
[Str92] W. A. Strauss, Partial differential equations - an introduction, John Wiley &
Sons Inc., New York, 1992.
[Str94] S. H. Strogatz, Nonlinear dynamics and chaos, Perseus, New York, 1994.
[Str04] B. Straughan, The energy method, stability, and nonlinear convection, second
ed., Applied Mathematical Sciences, vol. 91, Springer-Verlag, New York, 2004.
[Str08] , Stability and wave motion in porous media, Applied Mathematical
Sciences, vol. 165, Springer, New York, 2008.
[SU01] G. Schneider and H. Uecker, Nonlinear coupled mode dynamics in hyperbolic
and parabolic periodically structured spatially extended systems, Asymptotic
Anal. 28 (2001), no. 2, 163–180.
[SU03a] , Almost global existence and transient self similar decay for Poiseuille
flow at criticality over exponentially long times, Physica D 185 (2003), no. 3-
4, 209–226.
[SU03b] , Existence and stability of exact pulse solutions for Maxwell’s equa-
tions describing nonlinear optics, ZAMP 54 (2003), 677–712.
Bibliography 563
[Wu09] , Almost global wellposedness of the 2-D full water wave problem, In-
vent. Math. 177 (2009), no. 1, 45–135.
[Wu11] , Global wellposedness of the 3-D full water wave problem, Invent.
Math. 184 (2011), no. 1, 125–220.
[WW02] C. E. Wayne and J. D. Wright, Higher-order modulation equations for a
Boussinesq equation, SIAM J. Appl. Dyn. Syst. 1 (2002), no. 2, 271–302.
[WW15] C. E. Wayne and M. I. Weinstein, Dynamics of partial differential equa-
tions, Frontiers in Applied Dynamical Systems: Reviews and Tutorials, vol. 3,
Springer, Cham, 2015.
[Yan10] J. Yang, Nonlinear waves in integrable and nonintegrable systems, SIAM,
2010.
[YLWM13] S. Yanchuk, L. Lücken, M. Wolfrum, and A. Mielke, Spectrum and ampli-
tude equations for scalar delay-differential equations with large delay, Discrete
Contin. Dyn. Syst. 35 (2015), no. 1, 537–553.
[Yos71] K. Yosida, Functional analysis, Springer, New York, 1971.
[Yos82] H. Yosihara, Gravity waves on the free surface of an incompressible perfect
fluid of finite despth, Publ. Res. Inst. Math. Sci. 18 (1982), 49–96.
[Yos83] , Capillary-gravity waves for an incompressible ideal fluid, J. Math.
Kyoto Univ. 23 (1983), 649–694.
[Zak68] V. E. Zakharov, Stability of periodic waves of finite amplitude on the surface
of a deep fluid, Sov. Phys. J. Appl. Mech. Tech. Phys 4 (1968), 190–194.
[Zei71] E. Zeidler, Existenzbeweis für cnoidal waves unter Berücksichtigung der
Oberflächenspannung. (Existence proof for cnoidal waves under considera-
tion of the surface tension.), Arch. Ration. Mech. Anal. 41 (1971), 81–107
(German).
[ZF71] V. E. Zakharov and L. D. Fadeev, The Korteweg-de Vries equation is a fully
integrable Hamiltonian system, Functional Anal. Appl. 5 (1971), 280–287.
[Zim16] D. Zimmermann, Justification of the Ginzburg-Landau approximation for
quasilinear problems - including an application to Bénard-Marangoni con-
vection, Preprint Universität Stuttgart (2016), 89 pages.
[ZK65] N. J. Zabusky and M. D. Kruskal, Interactions of solitons in a collisionless
plasma and the recurrence of initial states, Phys. Rev. Lett. 15 (1965), 240–
243.
List of symbols
567
568 List of symbols
569
570 Index
theorem
Beppo-Levi, 226
center manifold, 86, 476
contraction mapping, 20
dominated convergence, 226
Floquet, 30
Fubini, 226
Hartman-Grobman, 43
Hausdorff-Young, 231
Heine Borel, 150
KAM, 126
Liouville, 111, 123
Lyapunov’s subcenter, 115
Nekhoroshov, 127
Paley-Wiener, 233
Picard-Lindelöf, 34
Poincaré-Bendixson, 55
Sarkovskii, 101
Smale-Birkhoff, 128
three wave interaction, 427, 448
traffic flow, 237
trapping region, 299
traveling wave
for reaction diffusion, 299
for the Burgers equation, 243
for the GL equation, 322
for the KdV equation, 261, 293
tsunami, 9
turbulence, 98
Turing instability, 307, 316, 347
Turing space, 310
Turing-Hopf, 375
www.ams.org
GSM/182