You are on page 1of 156

Applied Mathematical Sciences I Volume 35

Jack Carr

Applications of
Centre Manifold Theory

Springer-Verlag
New York Heidelberg Berlin
Jack Carr
Department of Mathematics
Heriot-Watt University
Riccarton, Currie
Edinburgh EH14 4AS
Scotland

Library of Congress Cataloging in Publication Data

Carr, Jack.
Applications of centre manifold theory.

(Applied mathematical sciences; v. 35)


"Based on a series of lectures given in the
Lefschetz Center for Dynamical Systems in the
Division of Applied Mathematics at Brown University
during the academic year 1978-79"-Pref.
1. Manifolds (Mathematics) 2. Bifurcation
theory. I. Title. II. Series: Applied mathematical
sciences (Springer-Verlag New York Inc.); v. 35.
QA1.A647 vol. 35 [QA613] 510s [516'.07] 81-4431
AACR2

All rights reserved.

No part of this book may be translated or reproduced in any


form without written permission from Springer-Verlag.

The use of general descriptive names, trade names, trademarks,


etc. in this publication, even if the former are not especially
identified, is not to be taken as a sign that such names, as
understood by the Trade Marks and Merchandise Marks Act, may
accordingly be used freely by anyone.

© 1981 by Springer-Verlag New York Inc.

Printed in the United States of America

9 8 7 6 5 4 3 2 1
To my parents
PREFACE

These notes are based on a series of lectures given in

the Lefschetz Center for Dynamical Systems in the Division of


Applied Mathematics at Brown University during the academic
year 1978-79.

The purpose of the lectures was to give an introduction

to the applications of centre manifold theory to differential


equations. Most of the material is presented in an informal
fashion, by means of worked examples in the hope that this
clarifies the use of centre manifold theory.

The main application of centre manifold theory given


in these notes is to dynamic bifurcation theory. Dynamic

bifurcation theory is concerned with topological changes in


the nature of the solutions of differential equations as para-
meters are varied. Such an example is the creation of periodic
orbits from an equilibrium point as a parameter crosses a

critical value. In certain circumstances, the application of

centre manifold theory reduces the dimension of the system


under investigation. In this respect the centre manifold

theory plays the same role for dynamic problems as the

Liapunov-Schmitt procedure plays for the analysis of static


solutions. Our use of centre manifold theory in bifurcation
problems follows that of Ruelle and Takens [57] and of Marsden
and McCracken [51].

In order to make these notes more widely accessible,


we give a full account of centre manifold theory for finite
dimensional systems. Indeed, the first five chapters are de-

voted to this. Once the finite dimensional case is under-


stood, the step up to infinite dimensional problems is
essentially technical. Throughout these notes we give the

simplest such theory, for example our equations are autono-

mous. Once the core of an idea has been understood in a

simple setting, generalizations to more complicated situations

are much more readily understood.

In Chapter 1, we state the main results of centre mani-


fold theory for finite dimensional systems and we illustrate

their use by a few simple examples. In Chapter 2, we prove

the theorems which were stated in Chapter 1, and Chapter 3


contains further examples. In Section 2 of Chapter 3 we out-

line Hopf bifurcation theory for 2-dimensional systems. In

Section 3 of Chapter 3 we apply this theory to a singular per-


turbation problem which arises in biology. In Example 3 of

Chapter 6 we apply the same theory to a system of partial dif-

ferential equations. In Chapter 4 we study a dynamic bifurca-

tion problem in the plane with two parameters. Some of the

results in this chapter are new and, in particular, they con-

firm a conjecture of Takens [64]. Chapter 4 can be read in-

dependently of the rest of the notes. In Chapter 5, we apply

the theory of Chapter 4 to a 4-dimensional system. In Chap-

ter 6, we extend the centre manifold theory given in Chapter

2 to a simple class of infinite dimensional problems. Fin-

ally, we illustrate their use in partial differential equa-

tions by means of some simple examples.

I first became interested in centre manifold theory

through reading Dan Henry's Lecture Notes [34]. My debt to

these notes is enormous. I would like to thank Jack K. Hale,

Dan Henry and John Mallet-Paret for many valuable discussions

during the gestation period of these notes.


This work was done with the financial support of the
United States Army, Durham, under AROD DAAG 29-76-G0294.

Jack Carr
December 1980
TABLE OF CONTENTS
Page

CHAPTER 1. INTRODUCTION TO CENTRE MANIFOLD THEORY . . . 1

1.1. Introduction . . . . . . . . . . . . . . . . 1
1.2. Motivation . . . . . . . . . . . . . . 1
1.3. Centre Manifolds . . . . . . . . . . . . . . 3
1.4. Examples . . . . . . . . . . . . . . . 5
1.5. Bifurcation Theory . . . . . . . . . . . . . 11
1.6. Comments on the Literature . . . . . . . . . 13

CHAPTER 2. PROOFS OF THEOREMS . . . . . . . . . . . . . 14

2.1. Introduction . . . . . . . . . . . . . . 14
2.2. A Simple Example . . . . . . . . 14
2.3. Existence of Centre Manifolds . . . . . . . 16
2.4. Reduction Principle. . . . . . . . 19
2.5. Approximation of the Centre Manifold . . . . 25
2.6. Properties of Centre Manifolds . . . . 28
2.7. Global Invariant Manifolds for Singular
Perturbation Problems. . . . . . . . . 30
2.8. Centre Manifold Theorems for Maps. . . . . . 33

CHAPTER 3. EXAMPLES . . . . . . . . . . . . . . . . . . 37

3.1. Rate of Decay Estimates in Critical Cases. . 37


3.2. Hopf Bifurcation . . . . 39
3.3. Hopf Bifurcation in a Singular Perturbation
Problem. . . . . . . . . . . . . . . . 44
3.4. Bifurcation of Maps . . . . . . . . . . . . . 50

CHAPTER 4. BIFURCATIONS WITH TWO PARAMETERS IN TWO


SPACE DIMENSIONS . . . . . . . . . . . . . . 54

4.1. Introduction . . . . . . . . . . . . . . . . 54
4.2. Preliminaries . . . . . . . . . . . . . . . . 57
4.3. Scaling . . . . . . . . . . . . . . . . . . . 64
4.4. The Case e >0 . . . . . . . . . . . . . . 64
4.5. The Case c < 0 . . . . . . . . . . . . . . 77
4.6. More Scalin . . . . . . 78
4.7. Completion of the Phase Portraits. . . . . . 80
4.8. Remarks and Exercises. . . . . . . . . . . . 81
4.9. Quadratic Nonlinearities . . . . . . . . . . 83

CHAPTER S. APPLICATION TO A PANEL FLUTTER PROBLEM . . . 88

5.1. Introduction 88
5.2. Reduction to a Second Order Equation . . . 89
5.3. Calculation of Linear Terms. . . . . . . . . 93
5.4. Calculation of the Nonlinear Terms . . . . . 95
Page

CHAPTER 6. INFINITE DIMENSIONAL PROBLEMS. . . . . . . . 97

6. 1 . Introduction . . . . . . . . . . . . . . . . 97
6.2. Semigroup Theory . . . . . . . . . . . . . . 97
6.3. Centre Manifolds . . . . . . . . . . . . . . 117
6.4. Examples . . . . . . . . . . . . . . . . . . 120
REFERENCES . . . . . . . . . . . . . . . . . . . . . . . 136
INDEX . . . . . . . . . . . . . . . . . . . . . . . . . . 141
CHAPTER 1
INTRODUCTION TO CENTRE MANIFOLD THEORY

1.1. Introduction

In this chapter we state the main results of centre

manifold theory for finite dimensional systems and give some


simple examples to illustrate their application.

1.2. Motivation

To motivate the study of centre manifolds we first


look at a simple example. Consider the system

x = ax3, y = -y + y2, (1.2.1)

where a is a constant. Since the equations are uncoupled


we can easily show that the zero solution of (1.2.1) is
asymptotically stable if and only if a < 0. Suppose now that

x = ax 3 + x2y
(1.2.2)
y = -y + y + xy - x3.

Since the equations are coupled we cannot immediately decide

if the zero solution of (1.2.2) is asymptotically stable, but

we might suspect that it is if a < 0. The key to understand-


ing the relation of equation (1.2.2) to equation (1.2.1) is
t
2 1. INTRODUCTION TO CENTRE MANIFOLD THEORY

an abstraction of the idea of uncoupled equations.

A curve, say y = h(x), defined for lx1 small, is

said to be an invariant manifold for the system of differen-

tial equations

x = f(x,Y), Y = g(x,Y), (1.2.3)

if the solution (x(t),y(t)) of (1.2.3) through (x0,h(x0))


lies on the curve y = h(x) for small t, i.e., y(t) =

h(x(t)). Thus, for equation (1.2.1), y - 0 is an invariant


manifold. Note that in deciding upon the stability of the

zero solution of (1.2.1), the only important equation is

x = ax3, that is, we only need study a first order equation


on a particular invariant manifold.

The theory that we develop tells us that equation


(1.2.1') has an invariant manifold y = h(x), jxj small, with
h(x) = 0(x2) as x + 0. Furthermore, the asymptotic stability
of the zero solution of (1.2.2) can be proved by studying a
first order equation. This equation is given by

u = au + u2h(u) = au3 + 0(u4), (1.2.4)

and we see that the zero solution of (1.2.4) is asymptotically

stable if a < 0 and unstable if a > 0. This tells us that


the zero solution of (1.2.2) is asymptotically stable if

a < 0 and unstable if a > 0 as we expected.

We are also able to use this method to obtain estimates


for the rate of decay of solutions of (1.2.2) in the case
a < 0. For example, if (x(t),y(t)) is a solution of (1.2.2)

with (x(0),y(0)) small, we prove that there is a solution

u(t) of (1.2.4) such that x(t) u(t)(1+o(1)),

y(t) - h(u(t))(1+o(1)) as t -
1.3. Centre Manifolds 3

1.3. Centre Manifolds

We first recall the definition of an invariant mani-


fold for the equation

x - N(x) (1.3.1)

where x E1Rn. A set S d1Rn is said to be a local invari-

ant manifold for (1.3.1) if for x0 E S, the solution x(t)

of (1.3.1) with x(O) = x0 is in S for Iti < T where

T > 0. If we can always choose T = then we say that S

is an invariant manifold.

Consider the system

Ax + f(x,y)
(1.3.2)
By + g(x,y)

where x E1Rn, y E1Rm and A and B are constant matrices

such that all the eigenvalues of A have zero real parts

while all the eigenvalues of B have negative real parts.


The functions f and g are C2 with f(0,0) - 0,

f'(0,0) = 0, g(0,0) = 0, g'(0,0) = 0. (Here, f' is the

Jacobian matrix of f.)

If f and g are identically zero then (1.3.2) has


two obvious invariant manifolds, namely x - 0 and y = 0.

The invariant manifold x = 0 is called the stable manifold,


since if we restrict initial data to x = 0, all solutions
tend to zero. The invariant manifold y = 0 is called the

centre manifold.

In general, if y - h(x) is an invariant manifold for

(1.3.2) and h is smooth, then it is called a centre manifold


if h(0) - 0, h'(0) - 0. We use the term centre manifold in
place of local centre manifold if the meaning is clear.
4 1. INTRODUCTION TO CENTRE MANIFOLD THEORY

If f and g are identically zero, then all solutions

of (1.3.2) tend exponentially fast, as t to solutions

of

x - Ax. (1.3.3)

That is, the equation on the centre manifold determines the

asymptotic behavior of solutions of the full equation modulo

exponentially decaying terms. We now give the analogue of

these results when f and g are non-zero. These results

are proved in Chapter 2.

Theorem 1. There exists a centre manifold for (1.3.2),

y - h(x), lxj < 6, where h is C2.

The flow on the centre manifold is governed by the


n-dimensional system

u = Au + f(u,h(u)) (1.3.4)

which generalizes the corresponding problem (1.3.3) for the

linear case. The next theorem tells us that (1.3.4) contains


all the necessary information needed to determine the asymp-

totic behavior of small solutions of (1.3.2).

Theorem 2. (a) Suppose that the zero solution of (1.3.4) is


stable (asymptotically stable) (unstable). Then the zero solu-

tion of (1.3.2) is stable (asymptotically stable) (unstable).

(b) Suppose that the zero solution of (1.3.4) is

stable. Let (x(t),y(t)) be a solution of (1.3.2) with

(x(O),y(O)) sufficiently small. Then there exists a solution

u(t) of (1.3.4) such that as t i m,

x(t) - u(t) + O(e- Yt)


11.3.5)
yt)
y(t) - h(u(t)) + O(e
1.4. Examples 5

where y > 0 is a constant.

If we substitute y(t) = h(x(t)) into the second equa-


tion in (1.3.2) we obtain

h'(x)[Ax + f(x,h(x))] = Bh(x) + g(x,h(x)). (1.3.6)

Equation (1.3.6) together with the conditions h(0) = 0,

h'(0) = 0 is the system to be solved for the centre manifold.


This is impossible, in general, since it is equivalent to

solving (1.3.2). The next result, however, shows that, in prin-


ciple, the centre manifold can be approximated to any degree
of accuracy.

For functions $:]Rn +]Rm which are CI in a neigh-


borhood of the origin define

(MO) (X) _ O'(x)[Ax + f(x,h(x))] - B$(x) - g(x,h(x))


Note that by (1.3.6), (Mh)(x) = 0.

C1
Theorem 3. Let $ be a mapping of a neighborhood of the
origin in ]Rn into ]Rm with $(0) = 0 and $'(0) = 0.

Suppose that as x _ 0, (M$)(x) - O(jxjq) where q > 1. Then


as x - 0, jh(x) - $(x)l = O(Ixl q)

1.4. Examples

We now consider a few simple examples to illustrate the

use of the above results.

Example 1. Consider the system

x - xy + ax3 + by2x

Y - -Y + cx2 + dx2Y.
6 1. INTRODUCTION TO CENTRE MANIFOLD THEORY

By Theorem 1, equation (1.4.1) has a centre manifold y = h(x).

To approximate h we set

(M0)(x) = m'(x)[xm(x) + ax3 + bxm2(x)l + m(x) - cx2 - dx2m(x)

If '(x) = 0(x2) then (MO) (x) = 4(x) - cx2 + 0(x4). Hence,

if m(x) = cx2, No) (x) = 0(x4), so by Theorem 3, h(x) =

cx2 + 0(x4). By Theorem 2, the equation which determines the

stability of the zero solution of (1.4.1) is

u = uh(u) + au3 + buh2(u) = (a+c)u3 + 0(u5).

Thus the zero solution of (1.4.1) is asymptotically stable if

a + c < 0 and unstable if a + c > 0. If a + c = 0 then

we have to obtain a better approximation to h.

Suppose that a + c = 0. Let 4(x) = cx2 + c(x) where

fi(x) = 0(x4). Then (M4)(x) = c(x) - cdx4 + 0(x6). Thus, if

m(x) = cx2 + cdx4 then (M4)(x) = 0(x6) so by Theorem 3,


2 4 6
h(x) = cx + cdx + 0(x ). The equation that governs the sta-

bility of the zero solution of (1.4.1) is

u = uh(u) + au3 + buh2(u) = (cd+bc2)u5 + 0(u7).

Hence, if a + c = 0, then the zero solution of (1.4.1) is

asymptotically stable if cd + bc2 < 0 and unstable if

cd + bc2 > 0. If cd + bc2 = 0 then we have to obtain a

better approximation to h (see Exercise 1).

Exercise 1. Suppose that a + c = cd + bc2 = 0 in Example 1.

Show that the equation which governs the stability of the

zero solution of (1.4.1) is u = -cd2u7 + 0(u9).

Exercise 2. Show that the zero solution of (1.2.2) 1% asymp-

totically stable if a < 0 and unstable if a n.


1.4. Examples 7

Exercise 3. Suppose that in equation (1.3.2), n = 1 so that


IyIq)
A = 0. Suppose also that f(x,y) = axp + O(IxIp+l +

where 2q > p + 1 and a is non-zero. Show that the zero

solution of (1.3.2) is asymptotically stable if a < 0 and

p is odd, and unstable otherwise.

Example 2. Consider the system

x = ex - x3 + xy
(1.4.2)
Y = -Y + Y2 - x2

where c is a real parameter. The object is to study small

solutions of (1.4.2) for small lei.

The linearized problem corresponding to (1.4.2) has

eigenvalues -1 and c. This means that the results given

in Section 3 do not apply directly. However, we can write

(1.4.2) in the equivalent form

3
x = ex - x + xy
y= -y+y 2
- x
2
(1.4.3)

When considered as an equation on ]R3 the ex term in

(1.4.3) is nonlinear. Thus the linearized problem correspond-

ing to (1.4.3) has eigenvalues -1,0,0. The theory given in


Section 3 now applies so that by Theorem 1, (1.4.3) has a two

dimensional centre manifold y = h(x,c), Ixl < 61, Iel < a2'
To find an approximation to h set

(M4)(x,c) - mx(x,e)[ex-x3+xm(x,e)] + m(x,e) + x2 - 02(x,e).

'then, if 4(x,c) - -x2, (M4)(x,e) = O(C(x,c)) where C is a

homogeneous cubic in x and c. By Theorem 3,


8 1. INTRODUCTION TO CENTRE MANIFOLD THEORY

h(x,c) - -x2 + O(C(x,c)) . Note also that h(O,c) = 0 (see

Section 2.6). By Theorem 2 the equation which governs small

solutions of (1.4.3) is

u = cu - 2u3 + O(IuIC(u,e))
(1.4.4)

The zero solution (u,c) = (0,0) of (1.4.4) is stable so

the representation of solutions given by Theorem 2 applies


here. For -62 < e < 0 the solution u = 0 of the first

equation in (1.4.4) is asymptotically stable and so by Theorem

2 the zero solution of (1.4.2) is asymptotically stable.

For 0 < c < 62, solutions of the first equation in

(1.4.4) consist of two orbits connecting the origin to two

small fixed points. Hence, for 0 < c < 62 the stable mani-

fold of the origin for (1.4.2) forms a separatrix, the unstable

manifold consisting of two stable orbits connecting the origin


to the fixed points.

Exercise 4. Study the behavior of all small solutions of


w + + ew + w3 = 0 for small C.

Example 3. Consider the equations

y = -y + (y+c)z
(1.4.5)
ei = y- (Y+1) z

where c > 0 is small and 0 < c < 1. The above equations

arise from a model of the kinetics of enzyme reactions [33].


If e = 0, then (1.4.5) degenerates into one algebraic equa-

tion and one differential equation. Solving the algebraic

equation we obtain
1.4. Examples 9

z= (1.4.6)
Y Y
and substituting this into the first equation in (1.4.5) leads

to the equation
_
(1.4.7)
y 1

where A = 1 - c.

Using singular perturbation techniques, it was shown

in [331 that for c sufficiently small, under certain condi-

tions, solutions of (1.4.5) are close to solutions of the de-


generate system (1.4.6), (1.4.7). We shall show how centre
manifolds can be used to obtain a similar result.

Let t = CT. We denote differentiation with respect

to t by ' and differentiation with respect to T by '.

Equation (1.4.5) can be rewritten in the equivalent form

Y' Ef(Y,w)
w' _ -w + y2 - yw + ef(Y,w) (1.4.8)

where f(y,w) = -y + (y+c)(y-w) and w = y - z. By Theorem

1, (1.4.8) has a centre manifold w = h(y,c). To find an ap-

proximation to h set

(MO)(Y,E) = Emy(Y,E)f(Y,m) + 0(Y,E) - y2 + y (Y, e) - ef(Y, m) -

If 0 (Y,e) = y2 - Acy then (Mm)(Y,c) O(1Y13 + IcI3) so

that by Theorem 3,

h(Y,e) - y2 - aey + O(IY13 + 1el3).

By Theorem 2, the equation which determines the asymptotic


behavior of small solutions of (1.4.8) is
10 1. INTRODUCTION TO CENTRE MANIFOLD THEORY

u' = ef(u,h(u,e))

or in terms of the original time scale

u = f(u,h(u,c)) _ -A(u-u2) + O(Ieul + lul3). (1.4.9)

Again, by Theorem 2, if a is sufficiently small and y(0),

z(0) are sufficiently small, then there is a solution u(t)

of (1.4.9) such that for some y > 0,

y(t) = u(t) + 0(e Yt/E)


(1.4.10)
z(t) = y(t) - h(y(t),c) + O(e yt/e).

Note that equation (1.4.7) is an approximation to the equation


on the centre manifold. Also, from (1.4.10), z(t) = y(t) -

y2(t), which shows that (1.4.6) is approximately true.

The above results are not satisfactory since we have to

assume that the initial data is small. In Chapter 2, we show

how we can deal with more general initial data. Here we

briefly indicate the procedure involved there. If y0 # -1,

then

(y,w,e) _ (y0,y0(l+y0) 1,0)

is a curve of equilibrium points for (1.4.8). Thus, we ex-

pect that there is an invariant manifold w = h(y,c) for

(1.4.8) defined for a small and 0 < y < m (m = 0(1)),

and with h(y,c) close to the curve

w = y2 (1+y)- 1. (1.4.11)

For initial data close to the curve given by (1.4.11), the

stability properties of (1.4.8) are the same as the stability

properties of the reduced equation

6 - f(u.h(u,e)).
1.5. Bifurcation Theory 11

1.5. Bifurcation Theory

Consider the system of ordinary differential equations

w = F (w, c)
F(O,c) s 0

where w EIRn+m and c is a p-dimensional parameter. We

say that c = 0 is a bifurcation point for (1.5.1) if the

qualitative nature of the flow changes at c = 0; that is, if

in any neighborhood of c = 0 there exist points and


c1
e2 such that the local phase portraits of (1.5.1) for

E = C1 and c = c2 are not topologically equivalent.

Suppose that the linearization of (1.5.1) about w = 0

is

w = C(e)w. (1.S.2)

If the eigenvalues of C(0) all have non-zero real parts


then, for small 10, small solutions of (1.5.1) behave like
solutions of (1.5.2) so that c = 0 is not a bifurcation
point. Thus, from the point of view of local bifurcation
theory the only interesting situation is when C(0) has eigen-

values with zero real parts.

Suppose that C(0) has n eigenvalues with zero real

parts and m eigenvalues whose real parts are negative. We

are assuming that C(0) does not have any positive eigen-

values since we are interested in the bifurcation of stable


phenomena.

Because of our hypothesis about the eigenvalues of

(:(0) we can rewrite (1.5.1) as


12 1. INTRODUCTION TO CENTRE MANIFOLD THEORY

X = Ax + f(x,y,c)

By + g(x,y,c) (1.5.3)

E= 0

where x EIRn, y EIRm, A is an n x n matrix whose eigen-

values all have zero real parts, B is an m x m matrix


whose eigenvalues all have negative real parts, and f and

g vanish together with each of their derivatives at

(x,y,E) _ (0,0,0).

By Theorem 1, (1.5.3) has a centre manifold y = h(x.E),

1xI < 61, IEI < d2. By Theorem 2 the behavior of small solu-

tions of (1.5.3) is governed by the equation

u - Au + f(u,h(u,c),c)
(1.5.4)
E = 0.

In applications n is frequently 1 or 2 so this is a

very useful reduction. The reduction to a lower dimensional

problem is analogous to the use of the Liapunov-Schmidt pro-


cedure in the analysis of static problems.

Our use of centre manifold theory in bifurcation prob-


lems follows that of Ruelle and Takens [57] and of Marsden

and McCracken [51]. For the relationship between centre

manifold theory and other perturbation techniques such as

amplitude expansions see [14].

We emphasize that the above analysis is local. In

general, given a parameter dependent differential equation it

is difficult to classify all the possible phase portraits.


For an example of how complicated such an analysis can be, see

[661 where a model of the dynamic behavior of a continuous

stirred tank reactor is studied. The model con%i%tt of is


1.6. Comments on the Literature 13

parameter dependent second order system of ordinary differ-

ential equations. The authors show that there are 3S pos-

sible phase portraits!

1.6. Comments on the Literature

Theorems 1-3 are the simplest such results in centre

manifold theory and we briefly mention some of the possible

generalizations.

(1) The assumption that the eigenvalues of the lin-


earized problem all have non-positive real parts is not

necessary.

(2) The equations need not be autonomous.


(3) In certain circumstances we can replace 'equilibrium

point' by 'invariant set'.

(4) Similar results can be obtained for certain classes

of infinite-dimensional evolution equations, such as partial

differential equations.

There is a vast literature on invariant manifold theory

[1,8,22,23,27,28,30,32,34,35,42,44,45,48,51]. For applications

of invariant manifold theory to bifurcation theory see [1,14,

17,18,19,24,31,34,36,37,38,47,48,49,51,57,65). For a simple

discussion of stable and unstable manifolds see (22, Chapter


13) or (27, Chapter 3].

In Chapter 2 we prove Theorems 1-3. Our proofs of


Theorems 1 and 2 are modeled on Kelly (44,45). Theorem 3 is

a special case of a result of Henry [34] and our proof follows


his. The method of approximating the centre manifold in
Theorem 3 was essentially used by Hausrath (32) in his work on

stability in critical cases for neutral functional differen-


tial equations. Throughout Chapter 2 we use methods that
generalize to infinite dimensional problems in an obvious way.
CHAPTER 2

PROOFS OF THEOREMS

2.1. Introduction

In this chapter we give proofs of the three main


theorems stated in Chapter 1. The proofs are essentially

applications of the contraction mapping principle. The pro-

cedure used for defining the mappings is rather involved, so

we first give a simple example to help clarify the technique.


The proofs that we give can easily be extended to the corres-

ponding infinite dimensional case; indeed essentially all we

have to do is to replace the norm 1.1 in finite dimensional

space by the norm 11.11 in a Banach space.

2.2. A Simple Example

We consider a simple example to illustrate the method

that we use to prove the existence of centre manifolds.

Consider the system

x1 = x2, x2 = 0, k = -Y + g(xl,x2), (2.2.1)

where g is smooth and g(xl,x2) = 0(x2+x2) as (xl,x2) +

(0,0). We prove that (2.2.1) has a local centre manifold.

1 e
2.2. A Simple Example is

Let *:IR2 IR be a C function with compact support


such that p(xl,x2) = 1 for (xl,x2) in a neighborhood of
the origin. Define G by G(xl,x2) = (xl,x2)g(xl,x2). We

prove that the system of equations

xl = x2, x2 = 0, Y = -y + G(xl,x2), (2.2.2)

has a centre manifold y = h(xl,x2), (xl,x2) E IR2. Since

G(xl,x2) = g(xl,x2) in a neighborhood of the origin, this

proves that y = h(xl,x2), x2 + x2 < 6 for some 6, is a

local centre manifold for (2.2.1).

The solution of the first two equations in (2.2.2) is

xl(t) = z1 + z2t, x2(t) = z2, where xi(0) = zi. If

y(t) = h(xl(t),x2(t)) is a solution of the third equation in

(2.2.2) then

ft-h(z = -h(zI+z2t,z2) + G(zl+z2t,z2). (2.2.3)

To determine a centre manifold for (2.2.2) we must single out


a special solution of (2.2.3). Since G(xl,x2) is small for
,ill xl and x2, solutions of the third equation in (2.2.2)
behave like solutions of the linearized equation y = -y.

l'he general solution of (2.2.2) therefore contains a term


t.
like e- As t + -, this component approaches the origin

perpendicular to the zl,z2 plane. Since the centre manifold

I,; tangent to the zl,z2 plane at the origin we must elim-

mate the a-t component, that is we must eliminate the com-


l-onent that approaches the origin along the stable manifold

.is t - To do this we solve (2.2.3) together with the


kondition

lim h(zl+z2t,z2)et - 0. (2.2.4)


t y -.
16 2. PROOFS OF THEOREMS

Integrating (2.2.3) between and 0 and using

(2.2.4) we obtain
0
h(zl,z2) = f- esG(zl+z2s,z2)ds.
1 m

By construction, y = h(zl,z2) is an invariant manifold for


(2.2.2). Using the fact that G has compact support and

that G(xl,x2) has a second order zero at the origin it

follows that h(zl,z2) has a second order zero at the origin;


that is, h is a centre manifold.

2.3. Existence of Centre Manifolds

In this section we prove that the system

x = Ax + f (x, y)
(2.3.1)
y = By + g(x,y)

has a centre manifold. As before x EIRn, y EIRm, the eigen-


values of A have zero real parts, the eigenvalues of B

have negative real parts and f and g are C2 functions


which vanish together with their derivatives at the origin.

Theorem 1. Equation (2.3.1) has a local centre manifold

y = h(x), lxi < 6, where h is C2.

Proof: As in the example given in the previous section, we

prove the existence of a centre manifold for a modified equa-

tion. Let p:IRn + [0,11 be a Cm function with ,y(x) = 1

when lxi < 1 and y,(x) = 0 when lxi > 2. For c > 0

define F and G by

F(x,y) = f(x*(E),y), G(x,y) - g(xP(E),y)

The reason that the cut-off function p is only a function of


2.3. Existence of Centre Manifolds 17

x is that the proof of the existence of a centre manifold


generalizes in an obvious way to infinite dimensional prob-

lems.

We prove that the system

x = Ax + F(x,y)
(2.3.2)
By + G(x,y)

has a centre manifold y = h(x) , x EIRn, for small enough c.

Since F and G agree with f and g in a neighborhood of

the origin, this proves the existence of a local centre mani-

fold for (2.3.1).

For p > 0 and pl > 0 let X be the set of Lipschitz


functions h:IRn +]Rm with Lipschitz constant pl, Ih(x)I < p

for xEIRn and h(0) = 0. With the supremum norm 11-11, X


is a complete space.
For h E X and x0 EIRn, let x(t,x0,h) be the solu-

tion of

x = Ax + F(x,h(x)), x(O,xO,h) = x0. (2.3.3)

The bounds on F and h ensure that the solution of (2.3.3)

exists for all t. We now define a new function Th by

r 0
(Th)(x0) = 1- e- Bs G(x(s,x0,h),h(x(s,x0,h)))ds. (2.3.4)

If h is a fixed point of (2.3.4) then h is a centre mani-

fold for (2.3.2). We prove that for p,pl, and c small

enough, T is a contraction on X.

Using the definitions of F and G, there is a con-

tinuous function k(c) with k(O) = 0 such that


18 2. PROOFS OF THEOREMS

IF(x,y)I + IG(x,y)I ck(c),

IF(x,y) - F(x',y')I c k(c)[Ix-x'I + Iy-y'II, (2.3.5)

IG(x,y) - G(x',y')I c k(c)[Ix-x'I + Iy-y'I],

for all x, x' E]Rn and all y, y' E]Rm with IyI, Iy'I < c.

Since the eigenvalues of B all have negative real

parts, there exist positive constants S,C such that for

s < 0 and y E IRm,

le- Bsyl
< CesslyI. (2.3.6)

Since the eigenvalues of A all have zero real parts, for

each r > 0 there is a constant M(r) such that for x E IRn

and s E IR,
IeAsxl
< M(r)erlsllxl. (2.3.7)

Note that in general, M(r) + m as r + 0.

If p < c, then we can use (2.3.5) to estimate terms

involving G(x(s,x0,h),h(x(s,x0,h))) and similar terms. We

shall suppose that p < c from now on.

If x0 E]Rn, then using (2.3.6) and the estimates on

G and h, we have from (2.3.4) that

ITh(x0)I < CS lck(c). (2.3.8)

Now let x0, x1 E]Rn. Using (2.3.7) and the estimates


on F and h, we have from (2.3.3) that for r > 0 and

t < 0,
M(r)e-rtlx0-x1I
Ix(t,x0,h) - x(t,xl,h)I <

0 r(s-t)Ix(s,x0,h)
+ (l+pl)M(r)k(c) I e - x(s,xl,h)Ids.
it

By Gronwall's inequality, for t < 0,


2.4. Reduction Principle 19

M(r)Ixl-x0Ie-Yt,

Ix(t,x0,h) - x(t,x1,h)I < (2.3.9)

where y = r + (l+pl)M(r)k(c). Using (2.3.9) and the bounds

on G and h, we obtain from (2.3.4)

1Jx0-x1I
ITh(x0) - Th(xl)I < C(M(r) +pl)k(c)(s-y) (2.3.10)

if c and r are small enough so that 0 > y.

Similarly, if hl, h2 E X and x0 E]Rn, we obtain

+(1+pi)M(r)k(c)r-1(S-y) 1]
ITh1(x0)-Th2(x0)I < Ck(c)(S
(2.3.11)
IIh1-h2II

By a suitable choice of p,p1,c and r, we see from

(2.3.8), (2.3.10) and (2.3.11) that T is a contraction on

X. This proves the existence of a Lipschitz centre manifold

for (2.3.2). To prove that h is C1 we show that T is a

contraction on a subset of X consisting of Lipschitz dif-

ferentiable functions. The details are similar to the proof


given above so we omit the details. To prove that h is C2

we imitate the proof of Theorem 4.2 on page 333 of (22].

2.4. Reduction Principle

The flow on the centre manifold is governed by the

n-dimensional system

u = Au + f(u,h(u)). (2.4.1)

In this section we prove a theorem which enables us to relate

the asymptotic behavior of small solutions of (2.3.1) to solu-


tions of (2.4.1).

We first prove a lemma which describes the stability


properties of the centre manifold.
20 2. PROOFS OF THEOREMS

Lemma 1. Let (x(t),y(t)) be a solution of (2.3.2) with

I(x(0),y(0))I sufficiently small. Then there exist positive


C1 and u such that

Iy(t) - h(x(t))I < C1e-utly(0) - h(x(0))I

for all t > 0.

Proof: Let (x(t),y(t)) be a solution of (2.3.2) with

(x(0),y(0)) sufficiently small. Let z(t) - y(t) - h(x(t)),

then by an easy computation

i = Bz + N(x,z) (2.4.2)

where

N(x,z) = h'(x)[F(x,h(x)) - F(x,z+h(x))] + G(x,z + h(x))

- G(x,h(x)).

Using the definitions of F and G and the bounds on h,

there is a continuous function 6(c) with 6(0) = 0 such

that IN(x,z)I < 6(c)Izi if IzI < c. Using (2.3.6) we ob-

tain, from (2.4.2),


t
e-'(t-s)Iz(s)Ids
Iz(t)I < Ce-stlz(0)I + C6(c)I0

and the result follows from Gronwall's inequality.

Before giving the main result in this section we make

some remarks about the matrix A. Since the eigenvalues of

A all have zero real parts, by a change of basis we can put

A in the form A = A + A2 where A2 is nilpotent and


1

A t
le 1 xI - IxI. (2.4.3)

Since A2 is nilpotent, we can choose the basis such that


2.4. Reduction Principle 21

IA2xI < (S/4)IxI, (2.4.4)

where S is defined by (2.3.6).


We assume for the rest of this section that a basis

has been chosen so that (2.4.3) and (2.4.4) hold.

Theorem 2. (a) Suppose that the zero solution of (2.4.1) is


stable (asymptotically stable) (unstable). Then the zero

solution of (2.3.1) is stable (asymptotically stable) (un-


stable).

(b) Suppose that the zero solution of (2.3.1) is

stable. Let (x(t),y(t)) be a solution of (2.3.1) with


(x(O),y(O)) sufficiently small. Then there exists a solu-
tion u(t) of (2.4.1) such that as t m,

x(t) - u(t) + 0(e-Yt)


0(e-Yt)
(2.4.5)
y(t) = h(u(t)) +

where y > 0 is a constant depending only on B.

Proof: If the zero solution of (2.4.1) is unstable then by


invariance, the zero solution of (2.3.1) is unstable. From

now on we assume that the zero solution of (2.3.1) is stable.


We prove that (2.4.5) holds where (x(t),y(t)) is a solution

of (2.3.2) with I(x(0),y(0))I sufficiently small. Since F

and G are equal to f and g in a neighborhood of the

origin this proves Theorem 2. We divide the proof into two


steps.

I. Let u0 E]Rn and z0 E]Rm with I(u0,z0)I suf-

ficiently small. Let u(t) be the solution of (2.4.1) with


u(0) - u0. We prove that there exists a solution (x(t),y(t))

of (2.3.2) with y(0) - h(x(0)) - z0 and x(t) - u(t),


22 2. PROOFS OF THEOREMS

y(t) - h(u(t)) exponentially small as t + m

II. By Step I we can define a mapping S from a

neighborhood of the origin in IRn+m into IRn+m by

S(u0,z0) _ (x0)z0) where x0 = x(0). For I(x0,z0)l suf-

ficiently small, we prove that (x0)20) is in the range of S.

I. Let (x(t),y(t)) be a solution of (2.3.2) and

u(t) a solution of (2.4.1). Note that if u(O) is suffici-

ently small,

u - Au + F(u,h(u)) (2.4.6)

since solutions of (2.4.1) are stable. Let z(t) - y(t) -

h(x(t)), $(t) - x(t) - u(t). Then by an easy computation

t - Bz + N($+u,z) (2.4.7)

AO + R($,z) (2.4.8)

where N is defined in the proof of Lemma 1 and

R($,z) - F(u+$,z+h(u+$)) - F(u,h(u)).

We now formulate (2.4.7), (2.4.8) as a fixed point


problem. For a > 0, K > 0, let X be the set of continuous
functions $: [0,m) +IRn with I,(t)eatl < K for all t > 0.

If we define II$II - sup{14(t)eatl: t > 0}, then X is a com-


plete space. Let (u0,z0) be sufficiently small and let u(t)
be the solution of (2.4.6) with u(0) - u0. Given $ E X
let z(t) be the solution of (2.4.7) with z(0) - z0.
Define T$ by

A1(t-s)
(T$)(t) - -
e [A20(s) + R(0(s),z(s))1ds. (2.4.9)
t

We solve (2.4.9) by means of the contraction mapping prin-

ciple. If $ is a fixed point of T, then retracing our


2.4. Reduction Principle 23

steps we find that x(t) = u(t) + 4(t), y(t) - z(t) + h(x(t))

is a solution of (2.3.2). We can take a to be as close to

S as we please at the cost of increasing K and shrinking

the neighborhood on which the result is valid. For simplicity

however, we take K - 1 and 2a - S where 0 is defined by

(2.3.6).

Using the bounds on F,G,h and the fact that


N($,0) - 0, there is a continuous function k(c) with
]Rn
k(0) - 0 such that if 01,02 E and zl,z2 EIRm with

Izil < c, then

IN($1,z1) - N($2,z2)I k(c)[Iz11101-021 + Izl-z21]


(2.4.10)
IR($1)z1) - R($2,z2)I k(c)[1z1-z2l + 141-$21]

From (2.4.7),
t
<
CIzOIe-Ot
+ Ck(e)j e-O(t-s)Iz(s)Ids
Iz(t)I
0

where we have used (2.3.6) and (2.4.10). By Gronwall's in-

equality
sl t
Iz(t)I CIz0Ie (2.4.11)

where S1 = 0 - Ck(e). From (2.4.9), if c is sufficiently

small,

e-at + CIzOIe- sls)ds


< k(e)1W(e-as + at
IT,(t)I < e-
it
where we have used (2.4.3), (2.4.4), (2.4.10) and (2.4.11).

Hence T maps X into X.

Now let 01,02 E X and let zl,z2 be the correspond-

ing solutions of (2.4.7) with zi(0) - z0. We first estimate


w(t) - zl(t) - z2(t). From (2.4.7) and (2.4.10),
24 2. PROOFS OF THEOREMS

t t s) [Izl(s)II01(s)-02(s)I
e
I w ( t )I < C k ( c ) I + Iw(s)lids.
0

Using (2.4.11),

t
e-S(t-s)Iw(s)Ids
Iw(t) I < Clk(c)IIO1-02IIe-St + Ck(c)IO

where C1 is a constant, so that by Gronwall's inequality

Iw(t) I = Iz1(t)-z2(t)I < C1k(c)II01-0211 e Slt. (2.4.12)

Using (2.4.4) and (2.4.12), for c sufficiently small,

ITO1(t)-T02(t)I < 21101-0211 + k(c)fm(I01(s) - 02(s)I

+ Izl(s) - z2(s)I)ds < a101-0211

where a < 1.

The above analysis proves that for each (u0,z0) suf-

ficiently small, T has a unique fixed point. If U is a

neighborhood of the origin in ]Rn+m then it is easy to re-

peat the above analysis to show that T: X x U i X is a con-

tinuous uniform contraction. This proves that the fixed point

depends continuously on u0 and z0.

II. Define S by S(u0,z0) (x0,20) where

x0 = u0 + 0(0). Since 0 depends continuously on u0 and

z0, S is continuous. We prove that S is one-to-one, so

that by the Invariance of Domain Theorem (see [11] or [601)

S is an open mapping. Since S(0,0) = 0, this proves that

the range of S is a full neighborhood of the origin in ]Rn+m

Proving that S is one-to-one is clearly equivalent

to proving that if u0 + 00(0) = ul + 01(0) then u0 = u1

and 00(0) = 01(0). If u0 + 00(0) - ul + 41(0) then the

initial values for x and y are the same, so that by


2.5. Approximation of the Centre Manifold 2S

uniqueness of solution of (2.3.2), u0(t) + 00(t) = ul(t) +

01(t) for all t > 0, where ui(t) is the solution of


(2.4.6) with ui(0) = ui. Hence, for t > 0,

u0(t) - ul(t) = O1 (t) - 00(t). (2.4.13)

Since the real parts of the eigenvalues of A are all zero,


limlul(t) - u0(t)lect = m for any c > 0 unless u1(0) _
tam
at
u0(0). Also, 10i(t)l < e- for all t > 0. It now follows
from (2.4.13) that S is one-to-one and this completes the
proof of the theorem.

2.5. Approximation of the Centre Manifold

For functions 0:IRn +IRm which are C1 in a neigh-

borhood of the origin define

(MO) (x) = O'(x)[Ax + f(x,0(x))] - BO (x) - g(x,0(x))

Theorem 3. Suppose that 0(0) = 0, 0'(0) = 0 and that

(Mo)(x) = 0(1x1q) as x + 0 where q > 1. Then as x + 0,

lh(x) - O(x)l = 0(Ixlq)

Proof: Let 6: Rn Rm be a continuously differentiable


function with compact support such that 0(x) = O(x) for

lxj small. Set

N(x) = 6' (x) [Ax + F(x,6(x))] - BO(x) - G(x,6(x)), (2.5.1)

where F and G are defined in Theorem 1. Note that


N(x) = O(1xjq) as x + 0.

In Theorem 1, we proved that h was the fixed point of


a contraction mapping T: X + X. Define a mapping S by

Sz - T(z+e) - 0; the domain of S being a closed subset


26 2. PROOFS OF THEOREMS

Y c X. Since T is a contraction mapping on X, S is a

contraction mapping on Y. For K > 0 let

Y - { z E X : l z ( x ) I < K j x j q for all x E I R n}.

If we can find a K such that S maps Y into Y then we

will have proved the theorem.

We first find an alternative formulation of the map S.

For z E Y let x(t,x0) be the solution of

ii = Ax + F(x,z(x) + 6(x)), x(O,xO) - x(0). (2.5.2)

From (2.3.4)

0
(T(z+6))(x0) - - e-BsG(x(s,x0),z(x(s,x0)) + O(x(s,x0)))ds.
f

r 0
d[e-BsO(x(s,x0))]ds
-6(x0) -I-
1 1
u-s

0
_ (e-Bs[BO(x(s,x0)) O(x(s,x0))]ds.

- as

Writing x for x(s,x0) etc., from (2.5.1) and (2.5.2)

B6 (x) - ds (x) = B6 (x) - 6' (x) [Ax + F (x, z (x) + 6 (x)) ]

_ -N(x) - G(x,6(x)) + 6' (x)[F(x,6)

- F(x,z(x) + 6(x))].

Using Sz - T(z+6) - 6 and the above calculations

0
(Sz)(x0) = r- e-BsQ(x(s,x0),
z(x(s,x0)))ds (2.5.3)

where x(s,x0) is the solution of (2.5.2) and

Q(x,z) - G(x,6+z) - G(x,6) - N(x) + 6'(x)[F(x,6)


(2.5.4)
F(x,6+z)].
2.5. Approximation of the Centre Manifold 27

We now show that S maps Y into Y for some K > 0.

By choosing 6 suitably, we may assume that I0(x)I < E

for all x E]Rn. Since N(x) = O(Ixlq) as x + 0,

IN(x)I < Cllxlq, x E]Rn (2. S. S)

where C1 is a constant. Now

IQ(x,z)I_

IQ(x,o)I + IQ(x,z) - Q(x,o)I


(2.5.6)
= IN(x)I + IQ(x,z) - Q(x,0)I.

We can estimate IQ(x,z) - Q(x,0)I in terms of the Lipschitz

constants of F and G. Using (2.3.5), there is a continuous

function k(c) with k(0) - 0, such that

IQ(x,z) - Q(x,0)I < k(c)IzI (2.5.7)

for IzI < c. Using (2.S.S), (2.5.6), (2.5.7), for z E Y

and x E fn, we have that

IQ(x,z)I < C11xIq + k(c)Iz(x)I


(2. S.8)
< (C1 + Kk(c)) IxIq.

Using the same calculations as in the proof of Theorem 1, if

x(t,x0) is the solution of (2.5.2), then for each r > 0,

there is a constant M(r) such that

Ix(t,x0)I < M(r)Ix0Ie Yt, t < 0 (2.5.9)

where y - r + 2M(r)k(c).

Using (2.3.6), (2.5.8) and (2.5.9), if z E Y,

C(C1+Kk(c))(M(r))q(S-gy)-llxOIq

-
C2Ix0Iq
I($z)(x0)I <

provided c and r are small enough so that 0 - qy > 0.


28 2. PROOFS OF THEOREMS

By choosing K large enough and c small enough, we have

that C2 < K and this completes the proof of the theorem.

2.6. Properties of Centre Manifolds

(1) In general (2.3.1) does not have a unique centre

manifold. For example, the system x - -x3, y = -y, has the


two parameter family of centre manifolds y - h(x,cl,c2)

where

c
1
exp(
1
x
-2
), x > 0
h(x,cl,c2) = 0 x - 0
1 - 2
c2 exp(- x ), x < 0.

However, if h and h1 are two centre manifolds for

(2.3.1), then by Theorem 3, h(x) - h1(x) - O(jxjq) as x + 0


for all q > 1.

(2) If f and g are Ck, (k > 2), then h is Ck

(441. If f and g are analytic, then in general (2.3.1)

does not have an analytic centre manifold, for example, it

is easy to show that the system

x - -x3, y = -y + x2 (2.6.1)

does not have an analytic centre manifold (see exercise (1)).

(3) Centre manifolds need not be unique but there are

some points which must always be on any centre manifold. For

example, suppose that (x0,y0) is a small equilibrium point

of (2.3.1) and let y - h(x) be any centre manifold for

(2.3.1). Then by Lemma 1 we must have y0 - h(x0). Similarly,

if r is a small periodic orbit of (2.3.1), then r must lie


on all centre manifolds.
2.6. Properties of Centre Manifolds 29

(4) Suppose that (x(t),y(t)) is a solution of (2.3.1)


which remains in a neighborhood of the origin for all t > 0.

An examination of the proof of Theorem 2, shows that there is

a solution u(t) of (2.4.1) such that the representation

(2.4.5) holds.

(5) In many problems the initial data is not arbit-


rary, for example, some of the components might always be
nonnegative. Suppose S c7Rn+m with 0 E S and that (2.3.1)

defines a local dynamical system on S. It is easy to check,

that with the obvious modifications, Theorem 2 is valid when


(2.3.1) is studied on S.

Exercise 1. Consider

X -x 3 y -y + x2. (2.6.1)

Suppose that (2.6.1) has a centre manifold y - h(x), where

h is analytic at x = 0. Then

h(x) - E anxn
n-2
for small x. Show that a2n+1 ' 0 for all n and that

an+2 = nan for n - 2,4,..., with a2 = 1. Deduce that

(2.6.1) does not have an analytic centre manifold.

Exercise 2 (Modification of an example due to S. J. van

Strien [63]). If f and g are Cm functions, then for

each r, (2.3.1) has a Cr centre manifold. However, the

size of the neighborhood on which the centre manifold is de-

fined depends on r. The following example shows that in

general (2.3.1) does not have a Cm centre manifold, even if


f and g are analytic.
30 2. PROOFS OF THEOREMS

Consider

x -ex - x3, Y = -Y + x2, 0. (2.6.2)

Suppose that (2.6.2) has a C' centre manifold y = h(x,c)

for lxi < 6, jej < 6. Choose n > 6-l. Then since

h(x,(2n)-1) is Cw in x, there exist constants

al,a2.... ,a2n such that

2n
1) -
h(x,(2n) alxi + O(x2n+1)
i=1

for lxi small enough. Show that ai - 0 for odd i and

that if n > 1,

(2i)(2n)-1)a2i
(1 - = (2i-2)a2i-2' i - 2,...,n
(2.6.3)
a2 0.

Obtain a contradiction from (2.6.3) and deduce that (2.6.2)


does not have a C centre manifold.

Exercise 3. Suppose that the nonlinearities in (2.3.1) are

odd, that is f(x,y) = -f(-x,-y), g(x,y) - -g(-x,-y). Prove

that (2.3.1) has a centre manifold y = h(x) with h(x) -

-h(-x). [The example -x3, y = -y, shows that if h is

any centre manifold for (2.3.1) then h(x) -h(-x) in

general.]

2.7. Global Invariant Manifolds for Singular Perturbation


Problems

To motivate the results in this section we reconsider

Example 3 in Chapter 1. In that example we applied centre

manifold theory to a system of the form


31
2.7. Global Invariant Manifolds

Y' - Ef(Y,w)

w' = -w + y2 - yw + Ef(Y,w) (2.7.1)

where f(0,0) = 0. Because of the local nature of our results


on centre manifolds, we only obtained a result concerning

small initial data. Let v - -w(l+y) + y2, s S(T), where


s'(T) = 1 + y(T); then we obtain a system of the form

Y' = Egl(Y,v)
v' = -v + Eg2(Y,v) (2.7.2)

E, -0

where gi(0,0) - 0, i = 1,2. Note that if y j -1, then


(y,0,0) is always an equilibrium point for (2.7.2) so we ex-
pect that (2.7.2) has an invariant manifold v - h(y,c) de-

fined for -1 < y < m, say, and c sufficiently small.

Theorem 4. Consider the system

x' - Ax + Ef (x,y, E)
Y' - By + Eg(x,Y,E) (2.7.3)

E' = 0

where x E]Rn, y E]Rm and A,B are as in Theorem 1. Suppose


also that f,g are C2 with f(0,0,0) - 0, g(0,0,0) - 0.
Let m > 0. Then there is a 6 > 0 such that (2.7.3) has an
invariant manifold y - h(x,E), jxj < m, Ic < 6, with
Ih(x,E)l < Cici, where C is a constant which depends on
m,A,B,f and g.

Proof: Let p:]Rn + [0,1) be a C_ function with p(x) - 1

if lxi < m and y(x) - 0 if lxi > m + 1. Define F and

C by
32 2. PROOFS OF THEOREMS

F(x,y,c) - ef(xip(x),y,e), G(x,y,c) = eg(xi(x),y,e).

We can then prove that the system

x' = Ax + F(x,y,c)
(2.7.4)
y' = By + G(x,y,c)

has an invariant manifold y = h(x,e), x E]Rn, for IcI suf-

ficiently small. The proof is essentially the same as that

given in the proof of Theorem 1 so we omit the details.

Remark. If x = (x1,x2,...9xn) then we can similarly prove

the existence of h(x,c) for mi < xi < mi.

The flow on the invariant manifold is given by the

equation

u' - Au + cf(u,h(u,c)). (2.7.5)

With the obvious modifications it is easy to show that the


stability of solutions of (2.7.3) is determined by equation

(2.7.5) and that the representation of solutions given in


(2.4.5) holds.

Finally, we state an approximation result.

Theorem 5. Let $:]Rn+l i]Rm satisfy $(0,0) = 0 and

J(M$)(x,e)1 < Cep for jxl < m where p is a positive

integer, C is a constant and

(M$)(x,e) - Dx$(x,e)[Ax + ef(x,$(x,e))] - B$(x,e)

- eg(x,$(x,e)).

Then, for 1xI < m,

jh(x,e) - $(x,e)l < C1ep

for some constant C1.


2.8. Centre Manifold Theorems for Maps 33

Theorem 5 is proved in exactly the same way as Theorem

3 so we omit the proof.

For further information on the application of centre


manifold theory to singular perturbation problems see Fenichel
[24] and Henry [34].

2.8. Centre Manifold Theorems for Maps

In this section we briefly indicate some results on


centre manifolds for maps. We first indicate how the study
of maps arises naturally in studying periodic solutions of

differential equations.

Consider the following equation in ]Rn

A = f(x,A) (2.8.1)

where f is smooth and A is a real parameter. Suppose that


for A = A0, (2.8.1) has a periodic solution y with period

1'. One way to study solutions of (2.8.1) near y for

A - A01 small is to consider the Poincare map P(A). To

define P(A) let y be some point on y, let U be a local

cross section of y through y and let U1 be an open


neighborhood of y in U. Then P(A): U1 i U is defined by
I'(a)(z) - x(s), where x(t) is the solution of (2.8.1) with
x(0) = z and s > 0 is the first time x(t) hits U.

(See [51] for the details).

If P(A) has a fixed point then (2.8.1) has a periodic


,,rhit with period close to T. If P(A) has a periodic
point of order n,(P(A)kz / z for 1 < k < n - 1 and

I'r'(a)n - z) then (2.8.1) has a periodic solution with period

,lose to nT. If P(a) preserves orientation and there is a


,losed curve which is invariant under P(a) then there exists
34 2. PROOFS OF THEOREMS

an invariant torus for (2.8.1).

If none of the eigenvalues of the linearized map

P'(A0) lie on the unit circle then it can be shown that

P(a) has essentially the same behavior as P(A0) for

Ix - small. Hence in this case, for IA - a0I small,


X01
solutions of (2.8.1) near y have the same behavior as when
A - a0. If some of the eigenvalues of P'(A0) lie on the

unit circle then there is the possibility of bifurcations

taking place. In this case centre manifold theory reduces

the dimension of the problem. As in ordinary differential

equations we only discuss the stable case, that is, none of

the eigenvalues of the linearized problem lie outside the


unit circle.
n+m + n+m
Let T: IR IR have the following form:

T(x.Y) - (Ax + f(x,y), By + g(x,y)) (2.8.2)

where x E]Rn, y E]Rm, A and B are square matrices such

that each eigenvalue of A has modulus 1 and each eigen-

value of B has modulus less than 1, f and g are C2

and f,g and their first order derivatives are zero at the

origin.

Theorem 6. There exists a centre manifold h:]Rn + ]Rm for

T. More precisely, for some c > 0 there exists a C2 func-

tion h: ]Rn + ]Rm with h(0) = 0, h'(0) - 0 such that

jxl < c and (xl,yl) - T(x,h(x)) implies yl - h(xl).

In order to determine h we have to solve the equation

(x1,h(xl)) - T(x,h(x))

By (2.8.2) this is equivalent to


2.8. Centre Manifold Theorems for Maps 35

h(Ax + f(x,h(x))) = Bh(x) + g(x,h(x)).

n
For functions $: IR i Dm define M$ by

(M$)(x) = $(Ax + f(x,h(x))) - B$(x) - g(x,h(x))

so that Mh - 0.

Theorem 7. Let $: IR
n
i
Dm be a C1 map with $(0) = 0,

01(0) = 0 and (M$)(x) = O(Ixlq) as x i 0 for some q > 1.

Then h(x) = $(x) + O(Ixlq) as x i 0.

We now study the difference equation

xr+l = Axr + f(xr''r)


(2.8.3)
yr+l = Byr +

As in the ordinary differential equation case, the asymptotic


behavior of small solutions of (2.8.3) is determined by the
flow on the centre manifold which is given by

ur+l = Aur + f(ur,h(ur)). (2.8.4)

Theorem 8. (a) Suppose that the zero solution of (2.8.4) is


stable (asymptotically stable) (unstable). Then the zero solu-

tion of (2.8.3) is stable (asymptotically stable) (unstable).


(b) Suppose that the zero solution of (2.8.3) is
stable. Let (xr,yr) be a solution of (2.8.3) with
(x1,yl)
sufficiently small. Then there is a solution ur of (2.8.4)

such that Ixr - url < KBr and Iyr - h(ur)I < KBr for all

r where K and B are positive constants with B < 1.

The proof of Theorem 6 and the stability claim of


Theorem 8 can be found in [30,40,51]. The rest of the asser-
tions are proved in the same way as the ordinary differential
36 2. PROOFS OF THEOREMS

equations case.

Exercise 4. Show that the zero solution of

xr+l = -xr + taxi + xryr

_ 1 2
yr+1 2 yr - axr + xryr

is asymptotically stable if a > 0.


CHAPTER 3
EXAMPLES

3.1. Rate of Decay Estimates in Critical Cases

In this section we study the decay to zero of solutions


of the equation

i + i + f(r) = 0
where f is a smooth function with

f(r) = r3 + ar5 + 0(r7) as r + 0, (3.1.2)

where a is a constant. By using a suitable Liapunov func-


tion it is easy to show that the zero solution of (3.1.1) is
asymptotically stable. However, because f'(0) = 0, the rate
of decay cannot be determined by linearization.

In [10] the rate of decay of solutions was given using


techniques which were special to second order equations. We

show how centre manifolds can be used to obtain similar re-


sults.

We first put (3.1.1) into canonical form. Let

x = r + i, y = i, then
x -f(x-y)
(3.1.3)
y - -y - f(x-y).
37
38 3. EXAMPLES

By Theorem 1 of Chapter 2, (3.1.3) has a centre manifold

y = h(x). By Theorem 2 of Chapter 2, the equation which

determines the asymptotic behavior of small solutions of

(3.1.3) is

u - -f(u - h(u)). (3.1.4)

Using (3.1.2) and h(u) - 0(u2),

u = -u3 + 0(u4). (3.1.5)

Without loss of generality we can suppose that the solution

u(t) of (3.1.5) is positive for all t > 0. Using L'Hopital's


rule,

ju(t)s-3ds.
-1 - lim - lim t-1
t- u J 1

Hence, if w(t) is the solution of

w - -w3, w(0) - 1, (3.1.6)

then u(t) - w(t+o(t)). Since

t-1/2
w(t) - 1 + Ct-3/2 + 0(t-5/2) (3.1.7)
7
where C is a constant, we have that

u(t) - t-1/2 + o(t-1/2). (3.1.8)

To obtain further terms in the asymptotic expansion of

u(t), we need an approximation to h(u). To do this, set

(M$)(x) - -$'(x)f(x-$(x)) + $(x) + f(x-$(x)).

If $(x) = -x3 then (M$)(x) - 0(x5) so that by

Theorem 3 of Chapter 2, h(x) - -x3 + 0(x5). Substituting this

into (3.1.4) we obtain


3.2. Hopf Bifurcation 39

u = -u3 - (a+3)u5 + 0(u7) (3.1.9)

Choose T so that u(T) a 1. Dividing (3.1.9) by u3, inte-


grating over [T,t] and using (3.1.8), we obtain

rt
w-1(u(t)) - t + constant + (3+a) ft u (3.1.10)
T

where w is the solution of (3.1.6). Using (3.1.8) and


(3.1.9),
ft
ftu2(s)ds -
u s ds + ft 0(u4(s))ds
T
u(s) T
T
(3.1.11)

- -ln t-1/2 + constant + 0(1).

Substituting (3.1.11) into (3.1.10) and using (3.1.7),

u(t) t-1/2 t-3/2[(a+3)ln t + C]


+ o(t-3/2) (3.1.12)
4/7
where C is a constant.

If (x(t),y(t)) is a solution of (3.1.3), it follows

from Theorem 2 of Chapter 2 that either (x(t),y(t)) tends to

zero exponentially fast or

x(t) = tu(t), y(t) = Tu3(t)


where u(t) is given by (3.1.12).

3.2. Hopf Bifurcation

There is an extensive literature on Hopf Bifurcation

[1,17,19,20,31,34,39,47,48,51,57,59] so we give only an out-


line of the theory. Our treatment is based on [19].
Consider the one-parameter family of ordinary dif-
ferential equations on ]R 2
40 3. EXAMPLES

z = f(x,a), xE IR2

such that f(0,a) = 0 for all sufficiently small a. Assume


that the linearized equation about x = 0 has eigenvalues

y(a) ± iw(a) where y(O) = 0, w(0) = w0 + 0. We also assume

that the eigenvalues cross the imaginary axis with nonzero

speed so that y'(0) + 0. Since y'(0) + 0, by the implicit


function theorem we can assume without loss of generality

that y(a) = a. By means of a change of basis the differen-

tial equation takes the form

i = A(a)x + F(x,a), (3.2.1)

where

r a -w(a)
A(a) =
1 w (a) a

F(x,a) = O(lx12).

Under the above conditions, there are periodic solu-

tions of (3.2.1) bifurcating from the zero solution. More

precisely, for a small there exists a unique one parameter

family of small amplitude periodic solutions of (3.2.1) in

exactly one of the cases (i) a < 0, (ii) a = 0, (iii) a > 0.

However, further conditions on the nonlinear terms are re-


quired to determine the specific type of bifurcation.

Exercise 1. Use polar co-ordinates for

xl = axl - wx2 + Kxl(x2 + x2)

x2 = wxl + ax2 + Kx2(x2 + x2)

to show that case (i) applies if K > 0 and case (iii) ap-

plies if K < 0.
3.2. Hopf Bifurcation 41

To find periodic solutions of (3.2.1) we make the sub-

stitution

x1 = Cr cos 0, x2 = er sin 0, a + ca, (3.2.2)

where c is a function of a. After substituting (3.2.2)

into (3.2.1) we obtain a system of the form

r = e[ar + r2C3(O,ac)I + e2r3C4(O,ac) + O(e3)


(3.2.3)
6 = m0 + O(e) .

We now look for periodic solutions of (3.2.3) with c - 0

and r near a constant r0. If C3 and C4 are independent

of 0 and the higher order terms are zero then the first

equation in (3.2.3) takes the form

c[ar + $r2] + e2Kr3. (3.2.4)

Periodic solutions are then the circles r = r0, where r0

is a zero of the right hand side of (3.2.4). We reduce the

first equation in (3.2.3) to the form (3.2.4) modulo higher

order terms by means of a certain transformation. It turns

out that the constant B is zero. Under the hypothesis K

is non-zero, it is straightforward to prove the existence of


periodic solutions by means of the implicit function theorem.
The specific type of bifurcation depends on the sign of K

so it is necessary to obtain a formula for K.

Let F(x,a) = [F1(xl,x2,a), F2(xl,x2,a)]T and let

F(xiIx2)a) = BZ(xl,x2,a) + B3(xl,x2,a) + 0(x4+x4) (3.2.5)

where Bi is a homogeneous polynomial of degree i in

(xl,x2). Substituting (3.2.2) into (3.2.1) and using (3.2.5)


we obtain (3.2.3) where for i - 3,4,
42 3. EXAMPLES

Ci(6,a) _ (cos 0)B1i _1(cos 6, sin 6,a)


(3.2.6)
+ (sin 0)Bi_1(cos 6, sin 6,a).

Lemma 1. There exists a coordinate change

r - r + cul(r,6,a,c) + c2u2(r,6,a,c)

which transforms (3.2.3) into the system

r = car + c2 r3K + O(c3)


(3.2.7)
6 - m0 + 0(c)

where the constant K is given by


2n
K - (1/2n)f [C4(6.0) - w01C3(6.0)D3(6.0)]dO (3.2.8)
0

where C3 and C4 are given by (3.2.6) and

D3(6,0) - (cos 6)BZ(cos 6, sin 6,0) (sin 6)B2(cos 6, sin 6,0).

The coordinate change is constructed via averaging. We

refer to [19] for a proof of the lemma.

If K = 0 then we must make further coordinate changes.


We assume that K + 0 from now on.

Recall that we are looking for periodic solutions of


(3.2.7) with c + 0 and r near a constant r0. This sug-

gests that we set a = -sgn(K)c and r0 = IKJ -1/2. The next


result gives the existence of periodic solutions of

r - c2[-sgn(K)r + r3K] + 0(c3)


(3.2.9)
e'W0+0(c)
with r - r0 small.

Lemma 2. Equation (3.2.9) has a unique periodic solution for


c small and r in a compact region either for c > 0 (when
3.2. Hopf Bifurcation 43

K< 0) or c < 0 (when K > 0) . Also

r = r0 + 0(c), 6(t,c) - mot + 0(c)

and the period of the solution i(c) is given by

t(c) - (21r/wo) + 0(c).

The periodic solution is stable if K < 0 and unstable if

K > 0.
Lemma 2 is proved by a simple application of the impli-

cit function theorem. We again refer to [19] for a proof.

Lemma 2 also proves the existence of a one parameter

family of periodic solutions of (3.2.1). We cannot immedi-


ately assert that this family is unique however, since we may

have lost some periodic solutions by the choice of scaling,


i.e. by scaling a + ac and by choosing c - -sgn(K)a.

In order to justify the scaling suppose that


x1 = R cos 6, x2 = R sin 6 is a periodic solution of (3.2.1)
bifurcating from x1 - x2 - 0. Then R satisfies

R = aR + 0(R2).

When R attains its maximum, = 0 so that a - 0(R). This

justifies the scaling a + ac. A similar argument applied to

periodic solutions of (3.2.7) justifies the choice of c.

Theorem. Suppose that the constant K defined by (3.2.8) is

nonzero. Then (3.2.1) has a unique periodic solution bifur-

cating from the origin, either for a > 0 (when K < 0) or

a < 0 (when K > 0). If x - R cos 6, y = R sin 6, then the

periodic solution has the form


44 3. EXAMPLES

R(t,a) _
JaK-lll/2 + O(JaI)

9(t,a) - w0t + 0(Jajl/2)

with period t(a) _ (2n/w0) + O(Iaj 1/2). The periodic solu-

tion is stable if K < 0 and unstable if K > 0.

Finally, we note that since the value of K depends

only on the nonlinear terms evaluated at a - 0, when apply-

ing the above theorem to (3.2.1) we only need assume that the

eigenvalues of A(a) cross the imaginary axis with non-zero

speed and that

A(0) =

with m0 non-zero.

3.3. Hopf Bifurcation in a Singular Perturbation Problem

In this section we study a singular perturbation prob-

lem which arises from a mathematical model of the immune re-

sponse to antigen [521. The equations are

ez [x3 + (a - )x + b -]
a m 6(1-x) - a - ylab (3.3.1)

ylab + y2b

where e,6,yl,y2 are positive parameters. In the above model


a and b represent certain concentrations so they must be
non-negative. Also, x measures the stimulation of the system
and it is scaled so that jxj < 1. The stimulation is assumed
to take place on a much faster time scale than the response
so that c is very small.
3.3. Hopf Bifurcation in a Singular Perturbation Problem 45

The above problem was studied in [52J and we briefly


outline the method used by Merrill to prove the existence of
periodic solutions of (3.3.1). Putting c = 0 in the first

equation in (3.3.1) we obtain

x3 + (a - 4)x + b - m 0. (3.3.2)

Solving (3.3.2) for x as a function of a and b we ob-

tain x = F(a,b) and substituting this into the second equa-


tion in (3.3.1) we obtain

6(1-F(a,b)) - a - ylab
2 (3.3.3)
-Ylab + b.
Y2

Using 6 as the parameter, it was shown that relative to a


certain equilibrium point (a0,b0) a Hopf bifurcation takes
place in (3.3.3). By appealing to a result in singular per-

turbation theory, it was concluded that for c sufficiently

small, (3.3.1) also has a periodic solution.

We use the theory given in Chapter 2 to obtain a simi-


lar result.

Let (x0,a0,b0) be a fixed point of (3.3.1). If

b0 + 0 then and x0,b0 satisfy


a0 = Y2/Y1
3
Y
1
x0 + (Y1 - 2)x0 + b0 - = 0,
Y
6(1-x 0) - Y2 - Y2b0 = 0.
2 1

Recall also that for the biological problem, we must have


b0 > 0 and jx0I < 1. We assume for the moment that x0 and
b0 satisfy (3.3.4) and these restrictions. The reality of
these solutions are considered later. We let a0 = Y2/Yl

for the rest of this section.


46 3. EXAMPLES

Let y = a - a0, z - b - b0, w - -i(x-x0) - x0y - z

where ip = 3x2 + a0 - . Then assuming ip is non-zero,

Ew = g(w,Y,z,E)

Y - f2(w,Y,z,E) (3.3.5)

i - f3(w,Y,z,E)

where

g(w,Y,z,E) - fl(w,Y,z,E) - Ex0f2(w,y,z,E) - Ef3(w,Y,z,E)

fl(w,Y,z,E) - -ipw + N(w+x0Y+z,Y)

i-1
f2(w,Y,z,E) - (2 -1x0 - 1 - ylb0)y + (. - y2)z

+
2
lw - ylyz
f3(w,Y,z,E) e -ylb0y - ylyz

-p-263
N(e,Y) - + 3i-1x062 - ye.

In order to apply centre manifold theory we change the

time scale by setting t = Es. We denote differentiation with


respect to s by ' and differentiation with respect to t

by Equation (3.3.5) can now be written in the form

w' - g(w,Y,z,E)
y' = Ef2(w,y,z,E)
(3.3.6)
z' = Ef3(w,y,z,E)

E' = 0.

Suppose that * > 0. Then the linearized system corresponding

to (3.3.6) has one negative eigenvalue and three zero eigen-


values. By Theorem 1 of Chapter 2, (3.3.6) has a centre mani-
fold w = h(y,z,c). By Theorem 2 of Chapter 2, the local be-

havior of solutions of (3.3.6) is determined by the equation


3.3. Hopf Bifurcation in a Singular Perturbation Problem 47

y' = Ef2(h(y,z,E),y,z,E)
(3.3.7)
z' = Ef3(h(y,z,E),y,z,E)

or in terms of the original time scale

y - f2(h(y,z,E),y,z,E)
(3.3.8)
i = f3(h(y,z,E),y,z,E).

We now apply the theory given in the previous section to show


that (3.3.8) has a periodic solution bifurcating from the

origin for certain values of the parameters.

The linearization of the vector field in (3.3.8) about

y - z = 0 irs given by
*-1
>y 1x0 - 1 - Ylb0
-
Y2
2 2
J(E) -
0
-Y1b0

If (3.3.8) is to have a Hopf bifurcation then we must have

trace(J(E)) - 0 and 2 *-I - Y2 > 0. From the previous analy-


sis, we must also have that x0,b0 are solutions of (3.3.4)
with 1x01 < 1, b0 > 0 and ip > 0. We do not attempt to ob-
tain the general conditions under which the above conditions
are satisfied, we only work out a special case.

Lemma. Let Y1 < 2Y2. Then for each E > 0, there exists

6(E), x0(E), b0(E) such that 0 < 2x0(E) < 1, b0(E) > 0,

> 0, 6(E)iy-1 - 2Y2 > 0, trace (J(E)) = 0 and (3.3.4) is

satisfied.

Proof: Fix Y1 and Y2 with Y1 < 2Y2. If x0,b0,6 sat-

isfy the second equation in (3.3.4) then

6
Y
trace(J(0)) - T[x0>V 1 - 1 (1-x0)J.
2Y
48 3. EXAMPLES

It is easy to show that there is a unique x0(0) E (0,-) that

satisfies trace(J(O)) = 0. Clearly i > 0 for this choice

of x0(0). We now obtain b0(0) and 6(0) as the unique

solution of (3.3.4) and an easy computation shows that

b0(0) > 0, 6(0) > 0 and 6(0)i-1 - 2Y2 > 0. By the implicit
function theorem, for c, x0 - x0(0), b0 - b0(0) sufficiently
small, there exists 6(e,x0,b0) = 6(0) + O(e) such that

trace(J(c)) = 2 i-1x0 - 1 - y1b0 + O(e) = 0.

After substituting 6 = 6(e,x0,b0) into (3.3.4), another


application of the implicit function theorem gives the re-

sult. This completes the proof of the lemma.

From now on we fix Y1 and Y2 with yl < 2Y2. Using


the same calculations as in the lemma, for each c and 6

with a and 6 - 6(c) sufficiently small there is a solution


x0(e,6), b0(e,6) of (3.3.4). Writing x0 = x0(e,6(e)) and
trace(J) as a function of 6, we have that

8
Y16(e)1-1 3
a6(trace(J(6)))
6=6(e) (f x0) [6y1x0

+ (2Y2-y1)x0 - 12y1x0] + O(e) < 0,

if c is sufficiently small. Hence, the eigenvalues of J(6)

cross the imaginary axis with non-zero speed at 6 - 6(c).

Let y1 = R(e)z, z1 - m(c)y where

R(e) - (y1b0(e))-1/2 + O(e), m(c) = [(6/2)x-1-Y2]1/2 + O(e).

For 6 = 6(c), (3.3.8) in these new coordinates becomes

Y1m- 1

y1 ` -w0z1 - (e)y1z1
(3.3.9)
1(e)y1,e)-Y1L-1(c)y1zI
z1 = w0y1 + m(e)(6/2)*-1h(m-1(e)z1,
3.3. Hopf Bifurcation in a Singular Perturbation Problem 49

where

w2(c) = Y1b0(c)[(6/2),y-1 + O(c).


Y2]

To apply the results of the last section, we need to

calculate the K(c) associated with (3.3.9). We shall show


how K(0) can be calculated; if K(0) is non-zero then K(c)

will be non-zero also. To calculate K(0) we need to know


the quadratic and cubic terms in (3.3.9) when c = 0. Thus,

we have to find h(y,z,0) modulo fourth order terms.

Let

(MO)(Y,z) ` -g(,(Y,z),Y,z,O) (3.3.10)

Then by Theorem 2 of Chapter 2, if we can find 0 such that

(MO)(Y,z) = O(y4+z4) then h(Y,z,0) = 0(Y,z) + O(y4+z4).


Suppose that

(3.3.11)
0 _ 02 + 03
where 0j is a homogeneous polynomial of degree j. Substi-

tuting (3.3.11) into (3.3.10) we obtain

(Mo)(Y,z) = *02(Y,z) - 3iy-1x0(x0y+z)2 + Y(XOY+z)


+ O(IYI3 + IzJ3).
Hence, if
1y(x0y+z)
02(Y,z) = 3iy-2x0(Xoy+z)2 - (3.3.12)

then (Mo)(y,z) = O(IyJ3 + jzJ3). Substituting (3.3.11) into

(3.3.10) with g2(y,z) given by (3.3.12), we obtain

(Mo)(Y,z) _ ip03(Y,z) + i-2(XOY+z)3 6iy-1x0(XOY+z)02(Y,z)

+ Y02(Y,z) + O(Y4+z4).

Hence,
50 3. EXAMPLES

h(y,z) 02(y,z) - i-3(x0y+z)3 + 6iy-2x0(xOY+z)02(Y,z)

*-1y02(Y,z)
+ O(Y4+z4)

where 02(y,z) is defined by (3.3.12). K(0) can now be cal-

culated as in the previous section. The sign of K(0) will

depend on yl and y2. If K(0) is non-zero then we can

apply Theorem 1 of Section 2 to prove the existence of

periodic solutions of (3.3.1). The stability of the periodic

solutions is determined by the sign of K(0). If K(0) is

zero then we have to calculate K(e).

3.4. Bifurcation of Maps

In this section we give a brief indication of some re-

sults on the bifurcation of maps. For more details and

references the reader is referred to the book by Iooss [40].

Let V be a neighborhood of the origin in ]Rm and

let F V -,]Rm be a smooth map depending on a parameter

u E]Rq with F(0) = 0 for all V. If bifurcation is to

take place then the linearized problem must be critical. We

assume that F6 has simple complex eigenvalues A(0), a(0)

with IA(0)I - 1, A(0) 0 ±1, while all other eigenvalues of

FO' are inside the unit circle. By centre manifold theory

all bifurcation phenomena take place on a two-dimensional


manifold so we can assume m = 2 without loss of generality.

We first consider the case when u is a one-dimen-

sional parameter and we assume that a(u) crosses the unit

circle, that is

A(u) 0 when u = 0.
U 1-1

If An 0 1 for n - 3 or 4 then in general Fu has an


3.4. Bifurcation of Maps 51

invariant closed curve bifurcating from 0 [57,58]. More-

over, if an = 1, n > 5, then in general Fu does not have

any period points of order n bifurcating from 0 [40].

However, if u is a two-dimensional parameter, then in gen-

eral Fu has periodic points of order n for u in a small

region [4,64]. The method of proof used in [4] and [64] re-

lies on approximating Fu by the time one map of a certain


differential equation; we outline a more direct proof.
Suppose that

Fu (w) - A(u)w + O(IwI2), w E1R2, u E IR2

where A(0) has eigenvalues a(u), a(u) with an(0) - 1,

n > 5. For definiteness we assume that n - 2p + 3 for some

positive integer p; the case n = 2p + 4 is treated in the

same way. Then by changing coordinates [40], Fu has the

normal form

2k
Fn(z) - An(u)z + + b(u)z2p+2 + R(z,z,u)
U
z
k-1 a k(u)l11

where z = x + iy, R(z,z,u) - O(jzj 2p+3) and ak(u), b(u)


are complex numbers. We assume that after a change of para-

meter, An(u) - 1 - u where u - ul + iu2. We also assume


that al(0) and b(0) are nonzero. If z = reie is a fixed

point of Fu then

h(r,O,u) = 0
where

ul + iu2 -
J1k+ b()r2eie + O(r22).

By redefining u + u(a1(u))-1, we can assume without loss of


generality that aI(u) - 1. Given 0 E [0,27T), ul > 0 and
52 3. EXAMPLES

u2 small enough, the equation

Re(h(r,e,u)) - 0

has a unique (small) solution r - r(ul'u2,O) with

r(ul,u2,0) - ul/2(1 + O(u)) (3.4.1)

Substituting this into the equation

Im(h(r,e,u)) = 0 (3.4.2)

we obtain

Im(ak(u))r2k + Im(b(u)e-in9)r2p+1 + O(r2P+2)


u2 =
k=2

Let v2(ul,u2,0 J2kIm(a(u))r2k. We now make the change


o f parameter c = el + ic = ul + i(u2 - u2). By (3.4.1),
2

the map u -+ c is one-to-one for u small enough. Equation

(3.4.2) now takes the form

e2 - Im(b(u)e-in9)e12p+l)/2(1 + O(el/2 + Ie2I)).

Since b(0) 0 0, by the implicit function theorem the above

equation has n distinct solutions 0r(e), r - 1,...,n, for


< Cei2p+1)/2 n
Ie2l where C is a constant. Thus Fu has
(n-2)/2
fixed points for u in a region of width ON
If A3(0) = 1, then for one parameter families of maps,

in general, there are bifurcating fixed points of order 3

and no invariant closed curves [40,41). The case A4(0) = 1

is more complicated. In this case the map can be put into

the normal form

Fu(z) - a(u)z + b(u)z2z + c(u)z3 + O(IzIS)


3.4. Bifurcation of Maps 53

where a(u) - i + O(u). If u E IR then in general either


fixed points of order 4 or an invariant closed curve bifur-
cate from 0; the actual details depend on a(u), b(u) and

c(u) [40,41,681. If u is a two-dimensional parameter then


Fu is studied by investigating bifurcations of the differen-
tial equation

A(u)z + B(u)z2z + C(u)z3 (3.4.3)

where A(u), B(u) and C(u) are chosen such that the time
one map T(u) of (3.4.3) satisfies

Fu (z) - iT(u) (z) - O(Iz15)

The bifurcation diagrams given in [41 for equation (3.4.3)

seem only to be conjectures. Bifurcations of (3.4.3) are


studied in [541 in the case A(u) - u E with bifurcation
parameters Re(u) and Re(B(u)).

We note that the above bifurcation results do not ap-


ply if the map has special properties. For example, F
U
could represent the phase flow of a periodic Hamiltonian sys-
tem so that Fu is a symplectic map. The following result
is due to Moser [531: if F is a smooth symplectic map with
F(0) - 0, then in general F has infinitely many periodic
points in every neighborhood of the origin if the linearized
map F'(0) has an eigenvalue on the unit circle. For an ap-

plication of this result to the existence of closed geodesics


on a manifold see [461.
CHAPTER 4

BIFURCATIONS WITH TWO PARAMETERS IN TWO DIMENSIONS

4.1. Introduction

In this chapter we consider an autonomous ordinary dif-


ferential equation in the plane depending on a two-dimensional

parameter c. We suppose that the origin x = 0 is a fixed

point for all c. More precisely, we consider

z f(x,E), x E ]R2, a El,E2) E ]R2,


(4.1.1)
f(O,E) E 0.

The linearized equation about x = 0 is

z = A(c)x,

and we suppose that A(0) has two zero eigenvalues. The ob-

ject is to study small solutions of (4.1.1) for (E1,E2) in

a full neighborhood of the origin. More specifically, we wish


to divide a neighborhood of c = 0 into distinct components,

such that if e,- are in the same component, then the phase

portraits of (4.1.1)E and (4.1.1)- are topologically equival-


ent. We also want to describe the behavior of solutions for

each component. The boundaries of the components correspond

S4
4.1. Introduction 55

to bifurcation points.

Since the eigenvalues of A(O) are both zero we have

that either (i) A(O) is the zero matrix, or (ii) A(O) has

a Jordan block,

There is a distinction between (i) and (ii) even for the study
of fixed points. Under generic assumptions, in case (ii),
equation (4.1.1) has exactly 2 fixed points in a neighborhood
of the origin. For case (i) the situation is much more com-
plicated [21,29].

Another distinction arises when we consider the eigen-


values of A(c). We would expect the nature of the eigen-
values of A(c) to determine (in part) the possible type of
bifurcation. If

then the eigenvalues of A(c) are always real so we do not

expect to obtain periodic orbits surrounding the origin. On

the other hand if

(4.1.2)

then the range of the eigenvalues of A(e) is a neighborhood


of the origin in , that is, if z is a small complex num-

ber then A(e) has an eigenvalue z for some C.

We shall assume from now on that A(e) is given by


(4.1.2).
56 4. BIFURCATIONS WITH TWO PARAMETERS

Takens [64,65] and Bodganov [see 3] have studied nor-

mal forms for local singularities. Takens shows, for example,

that any perturbation of the equation

2 2
xl = x2 + x1,

is topologically equivalent to

2 2
1 = x2 + xl, x2 = E1 + E x - xl
2 1

for some E1,E2. There are certain difficulties in applying

these results since we must transform our equation into nor-

mal form, modulo higher order terms.

In [47, p. 333-348], Kopell and Howard study (4.1.1)

under the assumptions that A(E) is given by (4.1.2) and that


2
a f
-2 2(0,0) + 0
ax1

where f2 is the second component of f. Their approach con-

sists of a systematic use of scaling and applications of the

implicit function theorem.

In this chapter, we use the same techniques as Kopell

and Howard to study (4.1.1) when the nonlinearities are cubic.

Our results confirm the conjecture made by Takens [64] on the

bifurcation set of (4.1.1).

Similar results are given in [4] together with a brief

outline of their derivation.

The results on quadratic nonlinearities are given in

Section 9 in the form of exercises. Most of these results can


be found in Kopell and Howard [47].
4.2. Preliminaries 57

4.2. Preliminaries

Consider equation (4.1.1) where A(e) is given by

(4.1.2). We also suppose that the linearization of f(x,e)

is A(e)x, and that

f(O,e) = 0, f(x,e) = -f(-x,e). (4.2.1)

The object is to study the behavior of all small solutions of


(4.1.1) for a in a full neighborhood of the origin. Equa-

tion (4.1.1) is still too general, however, so we shall make


some additional hypotheses on the nonlinear terms. Set

f = (fl,f2)T

3 3
- f2(0,0)' s = f2(0,0
axl axlx2

(Hl) a+ 0

(H2) $+ 0

3
f1(0,0) = 0.
(H3) ax3
1

(Hl) implies that for small c, (4.1.1) has either 1

or 3 fixed points. Under (Hl), it is easy to show that by


a change of co-ordinates in (4.1.1), we can assume (H3) (see
Remark 1). We assume (H3) in order to simplify the computa-
tions.

Under (Hl) we can prove the existence of families of


periodic orbits and homoclinic orbits. Under (Hl)-(H3) we can
say how many periodic orbits of (4.1.1) exist for fixed E.

The sign of $ will determine the direction of bifurcation

and the stability of the periodic orbits, among other things.


From now on we assume (Hl)-(H3).
58 4. BIFURCATIONS WITH TWO PARAMETERS

Level Curves of H(yl,y2) a (y2/2)


- (y2/2) + (y4/4).

Figure 1

E1

Bifurcation Set for the Case a < 0, 8 < 0.

Figure 2

The main results are given in Figures 2-5. Sections


3-8 of this chapter show how we obtain these pictures. The
pictures for $ > 0 are obtained by using the change of
variables x2 + -x2, E2 + -E2 and t + -t.
4.2. Preliminaries 59

E2

E1

Bifurcation Set for the Case a > 0, 8 < 0.

Figure 3

The cases a > 0 and a < 0 are geometrically dif-


ferent. The techniques involved in each case are the same
and we only do the more difficult case a < 0. The case
a > 0 is left for the reader as an exercise.
From now on we assume a < 0 in addition to (Hl)-(H3).
Note that this implies that locally, (4.1.1) has 1 fixed
point for E1 < 0 and 3 fixed points for E1 > 0.
60 4. BIFURCATIONS WITH TWO PARAMETERS

REGION 1 REGION 2

REGION 3

REGION 4

Phase Portraits for the Case a < 0, a < 0.

Figure 4
4.2. Preliminaries 61

ON L1

REGION 5

ON L2

Figure 4 (cont.)
62 4. BIFURCATIONS WITH TWO PARAMETERS

REGION 6

Figure 4 (cont.)
4.2. Preliminaries 63

REGION 1

REGION 2

REGION 3

REGION 4

Phase Portraits for the Case a > 0, $ < 0.

Figure S
64 4. BIFURCATIONS WITH TWO PARAMETERS

4.3. Scaling

We scale the variables in equation (4.1.1) so that the

first components of the non-zero fixed points are given by

±1 + 0(E). To do this we introduce parameters u,6, scaled

variables yl,y2 and a new time t by the relations

62l0111/2
6 = lEla-1 1/2, E2 = lall/26u, xl = 6yl, x2 = Y29

t = lal-1/26-1T.

For (u,6) in a neighborhood of the origin, (E1,E2) belongs


to a region of the form {(El,E2): 1Ell
< E0, El I
2
(constant)E 2}. The yi are assumed to lie in a bounded set,
say lyil < M. A further discussion of the scaling is given

in Section 6.

After scaling (4.1.1) becomes

Yl = Y2 + 62g1(u,6,Y)
(4.3.1)
y2 = sgn(E1)Y1 + - y1 + 6yyly2 + 62g2(u,6,Y)
uy2

where the dot means differentiation with respect to t,

y = slat-1/2 and gi(u,6,y) = 0(1). The size of the bounds


on the gi depends on M,u,6. We write t for T from now
on.

The cases E1 > 0 and E1 < 0 are treated separately.

4.4. The Case El > 0

With E1 > 0, equation (4.3.1) becomes

Yl = Y2 + 62g1(u,6,Y)
(4.4.1)
2
Y2 = y1 + zY2 - y1 + 6yyly2 + 6 92(u,6,Y)

Let H(yl,y2) (y2/2) Then along


- (y2/2) +
(y4/4).
4.4. The Case El > 0 65

solutions of (4.4.1),

H(Y1,Y2) = VY2 + 6yy2Y2 + 0(62). (4.4.2)

Note that for u = 6 = 0, H is a first integral of (4.4.1).


The level curves of H(yl,y2) = b consist of a figure
of eight if b = 0, and a single closed curve if b > 0 (see

Figure 1). For b > 0, the curve H(yl,y2) = b passes

through the point yl = 0, y2 = (2b)1'2. For b > 0 and 6

sufficiently small, we prove the existence of a function

u = u1(b,6) = -yP(b)6 + 0(62) such that for b > 0, (4.4.1)

with u = ul(b,6) has a periodic solution passing through the


point yl = 0, y2 = (2b)1'2, and with ul(0,6), (4.4.1)

has a figure of eight solutions.


For fixed u,6, the number of periodic solutions of

(4.4.1), surrounding all three fixed points, depends upon the

number of solutions of

u = ul(b,6) = -yP(b)6 + 0(62). (4.4.3)

We prove that P(b) - m as b - m and that there exists


b1 > 0 such that P'(b) < 0 for b < b1 and P'(b) > 0

for b > b1. These properties of P(b) determine the number


of solutions of (4.4.3).

Suppose, for simplicity of exposition, that ul(b,6)

-yP(b)6 and that y < 0. If 0 < b2 < bl, then there exists

b3 > b1 such that ul(b2,6) = ul(b3,6). Hence, if


ul(b216), then (4.4.1) has two periodic solutions, one
passing through yl = 0, y2 = (2b2)1'2, the other passing

through yl - 0, y2 = u > ul(0,6), then (4.4.1)


(2b3)1"2.

If

has one periodic solution surrounding all three fixed points.


66 4. BIFURCATIONS WITH TWO PARAMETERS

Finally, if u = u(b1,6), then the periodic solutions coincide.

In Figure 4, the periodic solutions surrounding all

three fixed points in regions 3-5 correspond to the periodic

solutions of (4.4.1) which are parametrized by u = ul(b,6),

b > b1. Similarly the "inner" periodic solutions in region 5

are parametrized by u = ul(b,6), 0 < b < bl. The curve L1

in (el,e2) space corresponds to the curve u = ul(0,6).

Similarly the curve L2 corresponds to the curve u = ul(bl,6)

(see Figure 2).

In general ul(b,6) is not identically equal to

-yP(b)6, but the results are qualitatively the same. For

example, we prove the existence of a function bl(6) = bl +

0(6), such that if u,6 satisfy u = ul(bl(6),6), then equa-

tion (4.4.3) has exactly one solution. The curve

u = ul(bl(6),6) which is mapped into the curve L2 in

(el,e2) space corresponds to the points where the two

periodic solutions coincide.

If b < 0, then the set of points for which

H(yl,y2) = b consists of two closed curves surrounding the

points (-1,0) and (1,0). For 0 < c < 1, we prove the

existence of a function u = u2(c,6) = -yQ(c)6 + 0(62) such

that for u = u2(c,6), (4.4.1) has a periodic solution surrounding

the point (1,0) and passing through yl = c, y2 = 0. Using


f(x) = -f(-x), this proves the existence of a periodic solu-

tion surrounding the point (-1,0) and passing through

yl = -c, y2 = 0.

We also prove that Q'(c) > 0 for 0 < c < 1. Let

6 > 0 and suppose u satisfies

u2(0,6) < -sgn(y)u < u2(1,6) (4.4.4)


4.4. The Case el > 0 67

Then the equation u = P2 (c,6) has exactly one solution.

Hence, for fixed u,6 satisfying (4.4.4), equation (4.4.1)

has exactly one periodic solution surrounding (1,0). The

region in (u,6) space corresponding to (4.4.4) is mapped

into region 4 in (el,e2) space (see Figure 2).

Lemma 1. For 6 sufficiently small, there exists a function

u = u(6) = -(4/5)6y + 0(62) such that when u(6), (4.4.1)

has a homoclinic orbit.

Proof: Let S(u,6), U(u,6) be the stable and unstable mani-


folds of the fixed point (0,0) in (4.4.1). These manifolds

exist, since for u,6 sufficiently small, (0,0) is a saddle

[271. Let H(u,6,+) be the value of H(yl,y2) when U(u,6)

hits y2 = 0, yl > 0. Similarly, H(u,6,-) is the value of


H(yl,y2) when S(u,6) hits y2 = 0, yl > 0. H(u,6,±) are

well defined since stable and unstable manifolds depend con-


tinuously on parameters.

Let I(u,6,+) denote the integral of H(yl,y2) over

the portion of U(u,6) with yl > 0, y2 > 0 from yl =

y2 = 0 to y2 = 0, yl > 0. Then

H(u,6,+) = I(u,6,+). (4.4.5)

Similarly, I(u,6,-) denotes the integral of H(yl,y2) over

the portion of S(u,6) with yl > 0, y2 < 0 from yl = y2 = 0

to y2 = 0, yl > 0, so that H(u,6,-) = I(u,6,-).

Equation (4.4.1) has a homoclinic orbit (with yl > 0)

if and only if

H(u,6,+) - H(u,6,-) = 0. (4.4.6)

We solve (4.4.6) by the implicit function theorem.


68 4. BIFURCATIONS WITH TWO PARAMETERS

Using (4.4.2) and (4.4.5),

H(u,6,+) f (iy2 + y6y2y2)dt + O(u2 + 62 (4.4.7)

where the above integral is taken over the portion of U(0,0)

from yl - y2 = 0 to y2 = 0, yl > 0. Similarly,

(1YZ + y6y2y2)dt + O(u2 + 62), (4.4.8)

where the integral is taken over the portion of S(0,0) from

0 to Y2 = 0, yl > 0. Using (4.4.7), (4.4.8) and


yl = Y2 =
U(0,0) = -S(0,0), we obtain

H(u,6,+) = -H(u,6,-) + O(u2 + 62) (4.4.9)

Using (4.4.7) and (4.4.9),

. (H1(0,0,+) - H1(0,0,-)) - 2f y2dt > 0,

so that by the implicit function theorem, we can solve (4.4.6)

for u as a function of 6, say u - u(6). We now show how

to get an approximate formula for u(6). We can write equa-

tion (4.4.6) in the form

of+y2dt + y6f+ y2Y2dt + O(u2 + 62) = 0.

4
Hence, using (4.4.1) and y2 = +[yl2 (yl/2)]1/2, we obtain

yly2dyl
y6f
U= + 0(6 2 ) = - 4 5 + 0(6 2 )
+ y2dy1

This completes the proof of Lemma 1.

Lemma 1 proves the existence of a homoclinic orbit of

(4.1.1) when (E1,E2) lies on the curve L1 given by

1
E2 - -(4/5)lal + O(Ei/2).
8E1
4.4. The Case el > 0 69

Using f(x,e) = -f(-x,e), when (el,e2) lies on L1, equa-

tion (4.1.1) has a figure of eight solutions.


We now prove the existence of periodic solutions of

(4.4.1) surrounding all three fixed points. In the introduc-

tion to this section, we stated that (4.4.1) has a periodic


solution passing through yl = 0, y2 = (2b) 1/2 for any b > 0.

In Lemma 2, we only prove this for "moderate" values of b.

The reason for this is that in (4.3.1) the gi are bounded

only for y in a bounded set. In Section 6, we show that by


a simple modification of the scaling, we can extend these re-

sults to all b > 0.

Lemma 2. Fix b > 0. Then for 0 < b < b and 6 suffici-

ently small, there exists a function u ul(b,6) -yP(b)6 +

0(62) such that if u = ul(b,6) in (4.4.1), then (4.4.1) has

a periodic solution passing through yl = 0, y2 = (2b)1/2.

As b - 0 the periodic solution tends to the figure of

eight solutions obtained in Lemma 1.

Proof: Let H(u,6,b,+) be the value of H(yl,y2) when the

orbit of (4.4.1) which starts at yl - 0, y2 = (2b) 1/2 inter-

sects y2 = 0. Similarly, H(u,6,b,-) is the value of

H(yl,y2) when the orbit of (4.4.1) which starts at yl = 0,


1/2
y2 = -(2b) is integrated backwards in time until it inter-
sects y2 = 0. Then (4.4.1) has a periodic solution passing

through yl = 0, y2 - (2b) 1/2 if and only if

H(u,6,b,+) - H(u,6,b,-) = 0. (4.4.10)

Let I(u,6,b,+) denote the integral of H(yl,y2) over

the portion of the orbit of (4.4.1) with y2 > 0, starting at

yl - 0, y, - (2b)1/2 and finishing at y2 - 0, yl > 0.


70 4. BIFURCATIONS WITH TWO PARAMETERS

Similarly, I(u,6,b,-) is defined by integrating backwards

in time. Thus,

H(u,6,b,±) = b + I(11,6,b,±). (4.4.11)

Using (4.4.2) and (4.4.11),

H(u,6,b,+) - b + J(uy2 + y6y2y2)dt + 0(u2+62) (4.4.12)

where the above integral is taken over the portion of the or-
1/2
bit of (4.4.1) with u = 6 = 0 from yl = 0, y2 = (2b)

to yl = c, y2 = 0 where

4b = c4 - 2c2. (4.4.13)

Similarly, I(u,6,b,-) = -I(u,6,b,+) + O(u2 + 62), so

that equation (4.4.10) may be written in the form

J(uy2 y6y2y2)dt + O(u2 + 62) = 0. (4.4.14)


+

Hence by the implicit function theorem, we can solve (4.4.14)

to obtain u = -yP(b)6 + 0(62) where

2
d
P(b) = yly2 yl (4.4.15)
y2dy1

In order to prove that the periodic solution tends to


the figure of eight solutions as b + 0, we prove that

H(u,6,b,±) - H(u,6,±) as b - 0. (4.4.16)

This does not follow from continuous dependence of solutions


on initial conditions, since as b + 0 the period of the

periodic solution tends to infinity. The same problem occurs

in Kopell and Howard [47, p. 339] and we outline their method.

For yl and y2 small, solutions of (4.4.1) behave like


4.4. The Case Cl > 0 71

solutions of the linearized equations. The proof of (4.4.16)

follows from the fact that the periodic solution stays close
to the solution of the linearized equation for the part of the

solution with (yl,y2) small and continuous dependence on


initial data for the rest of the solution.

Lemma 3. P(b) + - as b + m.

Proof: The integrals in (4.4.15) are taken over the curve

y2 In (y1/2) + 2b]1/2, from yl = 0 to yl = c where

c is defined by (4.4.13). Thus, JO(b)P(b) = Jl(b) where

Ji(b) = Jc w2i(w2 - (w4/2) + 2b)1/2dw. (4.4.17)


0

Substituting w = cz in (4.4.17) we obtain

r1
Ji(b) = c2 (cz)2ig(z)dz
J
0

where g(z) _ [(z2-1) + (c2/2)(l-z4)]1'2. Since g(z) <

g(c-1) for 0 < z < 1, we have that JO(b) < Dlc3 for some

positive constant D1. Similarly, there exists a positive


constant D2 such that Jl(b) > D2c5. The result now follows.

Lemma 4. There exists b1 > 0 such that P'(b) < 0 for

b < b1 and PI(b) > 0 for b > b1.

Proof: It is easy to show that P'(b) + as b + 0. Hence,

by Lemma 3 it is sufficient to show that if P'(bl) = 0 then

P"(bl) > 0.

Let r(w) _ [w2 - (w4/2) + 2b]1/2. Differentiating


(4.4.17) with respect to b we obtain
c 2i
dw. (4.4.18)
Ji J r w
0

Integrating by harts in .J0 we obtain


72 4. BIFURCATIONS WITH TWO PARAMETERS

c 4 2
JO = wr w dw. (4.4.19)
1
0
Also

1c r2 w w4 2 + 2b]
J = w = (c [w2 dw. (4.4.20)
0 0 r(w) r w
O

Similarly,

1c rc 2
rww
3J1 =
0
iw,6ww4 w = 3
0 rwj [w2 - (w4/2)+2b]dw. (4.4.21)

Using (4.4.18) - (4.4.21) we can express J0 and J1 in

terms of JI and J. A straightforward calculation yields

3J0 = 4bJ0 + J1
(4.4.22)
15J1 = 4bJ6 + (4+12b)Ji.

Suppose that P'(b1) = 0. Then JO(b1)P"(b1) _

J111(b1) - P(b1)J0"(b1). Using (4.4.22) we obtain

4b1(4b1+l)[J"(b1) - P(b1)J3(b1)]
(4.4. 23)
J6(b1) [P2(b1) + 8b1P(b1) - 4b1] .

Hence P"(b1) has the same sign as

P2(b1) + 8b1P(b1) - 4b1. (4.4.24)

Since P'(b1) = 0 we have that Ji(b1) = P(b1)J;(b1). Using


(4.4.22) we obtain

5P2(b1) + 8b1P(b1) - 4P(b1) - 4b1 = 0. (4.4.25)

Using b1 > 0, it is easy to show that (4.4.25) implies that

P(b1) < 1. Using (4.4.24) and (4.4.25), P"(b1) has the same

sign as P(b1) - P2(b1). This proves that P"(b1) > 0 as

required.
4.4. The Case e1 > 0 73

Lemma S. For 6 sufficiently small there exist b1(6) =

b1 + 0(6), b2(6) = b2 + 0(6), where P(0) = P(b2), with the


following properties:

Let 6 > 0.

(i) if u = u1(b1(6),6), then the equation

u = 'l(b,6) (4.4.26)

has exactly one solution.

(ii) If y < 0 and u1(b1(6),6) < u < ul(b2(6)P6)p


then equation (4.4.26) has exactly two solutions b3(6), b4(6)

with b3(6) = b3 + 0(6), b4(6) = b4 + 0(6), where b3 and

b4 are solutions of

ly 1.
P(b) = -ud (4.4.27)

A similar result holds if y > 0.

(iii) If y < 0 and u > ul(b2(6),6), then (4.4.26)


has exactly one solution b5(6) = b5 + 0(6), where b5 is

the unique solution of (4.4.27).

(iv) If y < 0 and u < u1(b1(6),6), then (4.4.27)

has no solutions. A similar result holds for y > 0.

Proof: By Lemma 4, there exists b2 > 0 such that P(b2) _

P(0) and P'(b2) > 0. Set g(z,6) = 6-1[111(b2+z,6) -

ul(0,6)], for 6 + 0 and g(z,O) = 0. Then g(z,6) _


P'(b2)z + 0(161 + z2). By the implicit function theorem there
exists z(6) = 0(6) such that g(z(6),6) = 0. Hence, if

b2(6) = b2 + z(6) then u1(b2(6),6) = u1(0,6). The existence

of b1(6) is proved in a similar way. The rest of the lemma

follows from the properties of P(b).

We now prove the existence of periodic solutions of


(4.4.1) currnunding a jingle fixed point.
74 4. BIFURCATIONS WITH TWO PARAMETERS

Lemma 6. For 0 < c < 1 and 6 sufficiently small, there

exists u = u2(c,6) _ -yQ(c)6 + 0(62) such that if

u - u2(c,6), then (4.4.1) has a periodic solution passing

through yl - c, y2 - 0. As c - 0 the periodic orbits tend

to the homoclinic orbit obtained in Lemma 1.

Proof: Let H(u,6,c,+) be the value of H(yl,y2) when the


orbit of (4.4.1) starting at yl - c, y2 - 0 intersects

y2 - 0, y2 > 0. Similarly, H(u,6,c,-) is the value of


H(yl,y2) when the orbit of (4.4.1) starting at yl = c,

y2 = 0 is integrated backwards in time until it intersects

y2 - 0. Equation (4.4.1) will have a periodic orbit passing


through yl - c, y2 - 0 if and only if

H(u,6,c,+) - H(u,6,c,-) - 0. (4.4.28)

Using the same method as in Lemma 2, we can rewrite

equation (4.4.28) as

G(u,6,c) - j(uy2 + y6y2y2)dt + O(u2 + 62) - 0 (4.4.29)

where the above integral is taken over the curve

y2 = r(yl) = [y1 - (y4l /2) + (c4/2)-c2]1/2, (4.4.30)

from yl = c, y2 - 0 to y2 = 0 again. By (4.4.29)

au G(0,0,c) - f y2dt > 0. (4.4.31)

Thus, for fixed c we can solve (4.4.29). We cannot solve


(4.4.29) uniformly in c however, since as c - 1 the right
hand side of (4.4.31) tends to 0. We use a method similar to
that used by Kopell and Howard in [47, p. 337-338] to obtain
u2(c,6) for 0 < c < 1.
4.4. The Case c1 > 0 75

Equation (4.4.1) has a fixed point at (yl,y2) -

(1,0) + O(IiiI,I61). For u,6 sufficiently small we make a


change of variables yj = hj(yl)y2,u,6) so that the fixed
point is transformed into (yl,y2) = (1,0). An easy calcula-
tion shows that if we make this transformation, then the only
change in (4.4.1) is in the functions gl and g2. We sup-

pose that the above change of variables has been made and we
write y, for y,.

The curves H(yl,y2) = H(c,0) can be written in the

form

4H(Y1,Y2)+1 - 2y2 + (y1-l) 2(yl+l)2 - (c-1)2(c+l)2. (4.4.32)

Thus, for c close to 1 the closed curves are approximately

y2 + 2(yl-l) 2 - 2(c-1) 2. (4.4.33)

So, instead of equation (4.4.29) we consider the equation

G(1,6,c) = (c-1)-2G(c,u,6) - 0. (4.4.34)

If we prove that G(u,6,c) is bounded for 0 < c < 1 and

that (3/3u)_(0,0,c) is bounded away from zero for 0 < c < 1,

then we can solve (4.4.34) uniformly in c.

Now

H(Y1,Y2) = VY2 + 6(yy12


y + 6Y2g2 + 6(y3l-yl)g1].

Since the fixed point is at (1,0), H(yl,y2) is bounded above


and below by quadratic forms in y2 and yl - 1. By (4.4.32),

4H(yl,y2) + 1 is bounded above and below by quadratic forms


in y2 and yl - 1. Hence there exist functions gi(u,6)

O(juj + 161) such that


76 4. BIFURCATIONS WITH TWO PARAMETERS

gl(u,6) < at ln(4H(yl,y2) + 1) g2(u,6)

Integrating the above inequality over the curve given by

(4.4.30) we obtain

exp(gl(11,6)T) < [4H(u,6,c,+)+1]/[4H(c,0)+1]


(4.4.35)
< exp(g2(u,6)T),

where T is a bound for the time taken to trace the orbit.

Using (4.4.29), (4.4.32) and (4.3.35) we see that 6(11,6,c)

is bounded. The fact that (3/3v)-(c,0,0) is bounded away


from zero follows easily from (4.4.32). This proves that for
6 sufficiently small and 0 < c < 1, we can solve (4.4.32)

to obtain u - u2(c,6) - -yQ(c)6 + 0(62) where

J0(c)Q(c) = J1(c),
d
Ji(c) - 1 w2ir(w)dw
c

and r(w) is defined by (4.4.30) and r(d) - 0, d > c.

The fact that the periodic solution tends to the homo-


clinic orbit is proved in the same way as the corresponding

result in Lemma 2.

Using f(x) = -f(-x), Lemma 6 proves the existence of


periodic solutions of (4.4.1) surrounding (-1,0).

Lemma 7. Q'(c) > 0 for 0 < c < 1.

Proof: We write Q,J0 and J1 as functions of b where

4b - c4 - 2c2. Since (db/dc) < 0 for 0 < c < 1, we must


prove that Q1(b) < 0 for -1 < 4b < 0. Following the same
procedure as in Lemma 4, we find that J0,J0,, J1,Ji satisfy
equation (4.4.22). Thus, if Q'(bl) - 0 then
4.5. The Case c1 < 0 77

5Q2(bl) + 8b1Q(bl) - 4Q(bl) - 4b1 = 0. (4.4.36)

Since -1 < 4b1 < 0, the roots of (4.4.36) are less than 1.

Hence, if Q'(bl) - 0 then Q(bl) < 1. Also, from (4.4.23),


if Q'(bl) - 0 then Q"(bl) has the same sign as 4b1 -

8b1Q(bl) - Q2(bl). Using (4.4.36) and Q(bl) < 1, this im-

plies that Q"(bl) < 0. Since Q(-1/4) - 1, Q(O) - 4/5, this


shows that Q1(b) < 0 for -1 < 4b < 0. This completes the

proof of the lemma.

4.5. The Case El < 0

With < 0, equation (4.3.1) becomes


El

yl = Y2 + 6g1(u,6,Y)
(4.5.1)
Y2 - -Yl + uY2 - yl + 6Yyly2 + 62g2(v,6,Y)

Let H1(yl'y2) = (y2/2) + (y1/2) + (y4l/4). Then along solu-

tions of (4.5.1), H1 - yZ + 6yy2y2 + 0(62). Using the same


methods as in the previous lemmas, we prove that (4.5.1) has
a periodic solution passing through y1 - c > 0, y2 = 0, if
and only if u = u3(c,6) - -yR(c)6 + 0(62), where

JO(c)R(c) = J1(c),

c
J1(c) - r w2ir(w)dw,
0

r(w) = [2b - (w4/2) - w211/2, 4b - c4 + 2c2.

In order to prove that for fixed u,6, equation (4.5.1)

has at most one periodic solution we prove that R is strictly

monotonic.

Lemma 8. R'(c) > 0 for c > 0.


78 4. BIFURCATIONS WITH TWO PARAMETERS

Proof: We write R,JO,J1 as functions of b. It is suffici-

ent to prove that R'(b) > 0 for b > 0.

Using the same methods as before, we show that

3J0 - 4bJ6 - J1
(4.5.2)
15J1 - (12b+4)Ji - 4bJ6

3wwdw
- 4bJ6 - 4Ji. (4.5.3)
to

Now R' has the same sign as S where S - 15[JiJO - J0,J1].

By (4.5.2),

S - (8b-4)JO'Ji - 5(Ji)2 + 4b(Jo,) 2.

By (4.5.3), bJ0 > J. Hence, for b > 2,

S > (J,)2b-1(8b-4-5b+4)
- 3(Ji)2 > 0.

Similarly, if 0 < b < 2, then S > 3b2(Jo)2 > 0. This com-


pletes the proof of Lemma 8.

4.6. More Scaling

In Section 4 we proved the existence of periodic solu-

tions of (4.4.1) which pass through the point (0,(2b)1/2).

Lemma 3 indicates that we may take b to be as large as we


please. However our analysis relied on the fact that the gi

are 0(1) and this is true only for y in a bounded set.

Also, our analysis in Sections 4 and S restricts E1,E2 to a


region of the form {(E1,E2): 1c11 < EO, E1 < (constant) c } .

To remedy this we modify the scaling by setting


6 = 1Ela-111/2h-1, where h is a new parameter with, say,
0 < 4h < 1. The other changes of variables remain the
same. Note that if u,6 lie in a full neighborhood of the
4.6. More Scaling 79

origin then (c1,e2) and (xl,x2) lie in a full neighbor-

hood of the origin.

After scaling, (4.1.1) becomes

Yl - Y2 + 62g1(11,6,h,y)
(4.6.1)
y2 - h2sgn(el)Y1 + 11Y2 - yl + 6yyiy2 + 62g2(i,6,h,Y)

For 0 < 4h < 1 and 11,6 sufficiently small, the gi are

0(1) for y in a bounded set.

Let c1 > 0. Let H2(Y1OY2) _ (Y2/2) - (h2y2/2) +

(y4/4). Then along solutions of (4.6.1), H2 - 11y2 +


y6y2y2 + 0(112 + 62). Following the same procedure as in Sec-
1 2

tion 4, we find that (4.6.1) has a periodic solution passing


through yl = 0, y2 = 1, if and only if

11 = h2111(b,6) = -yh2P(b)6 + 0(62),

where 2b - h-4. An easy computation shows that as h 0,

h2P(b) = K + 0(h2), where J0K = 2 J1,


r1
Ji = fl w
0

Thus, for small h, (4.6.1) has a periodic solution passing

through yl = 0, y2 - 1 if and only if 11 - -yK6 +


0(h2161 + 62). In particular, when el - 0, E2 > 0, B < 0,

(4.1.1) has a periodic solution passing through xl - 0,

x2 =
IaI(-BK)-1c2
+ 0(c3/2)

For < 0, a similar analysis shows that (4.6.1) has


c1
a periodic solution passing through yl = 1, y2 - 0, if and

only if

h2u2(h-1,6) -yh2Q(h-1)6
u = = + 0(62),

and that as h 0, h2Q(h-1) - (K//f) + 0(h2).


80 4. BIFURCATIONS WITH TWO PARAMETERS

4.7. Completion of the Phase Portraits

Our analysis in the previous sections proves the exist-

ence of periodic and homoclinic orbits of (4.1.1) for certain

regions in (el,e2) space. We now show how to complete the

pictures.

Obtaining the complete phase portrait in the different

regions involves many calculations. However, the method is

the same in each case so we only give one representative

example. We prove that if y < 0, then the "outer" periodic

orbit in region 5 is stable. We use the scaling given in

Section 3.

Since we are in region 5, ul(0,d) > u > ul(bl(d),d).

Fix u and d and let b2 be the solution of u = ul(b,d)

with b2 > bl(d). Then we have to prove that the periodic

solution r of (4.4.1) passing through yl = 0, y2 = (2b2)1/2

is stable.

Let b3 be the solution of u = ul(b,d) with

b3 < bl(d). Then there is a periodic solution passing through


yl = 0, y2 = (2b3)1/2.
In fact, we prove that any solution

of (4.4.1) starting "outside" this periodic solution (and in-

side some bounded set) tends to r as t - m.

Let b > b3. Let c(b,+) be the value of yl when

the orbit of (4.4.1) starting at yl = 0, y2 = (2b) 1/2 first

hits y2 = 0. Similarly, c(b,-) is the value of yl when

the orbit of (4.1.1) starting at yl = 0, y2 = (2b) 1/2 is

integrated backwards until it hits y2 = 0. Using f(x,c) _

-f(-x,c), the solution passing through (0,(2b)1/2), spirals

inwards or outwards according to the sign of c(b,+) + c(b,-).

Using the calculation in Lemma 2, c(b,+) + c(b,-) has the


4.8. Remarks and Exercises 81

same sign as H(u,d,b,+) - H(u,d,b,-) which in turn has the


same sign as

S = ul(b2,d) - ul(b,6).

Using the properties of ul(b,b) given in Lemma 5, S is

positive if b < b2 and negative if b > b2. The result now


follows.

4.8. Remarks and Exercises

Remark 1. Consider equation (4.1.1) under the hypothesis


that the linearization is given by x = A(c)x where A(c)

is defined by (4.1.2). Suppose also that (4.2.1) and (Hl)


hold. Make the change of variables x = B(c)x, where
B(c) = I - ra-1A(e) and

a3f1(0,0) a3f2(0,0)
r =
3
ax 3X3

The map x - x will be one-to-one for c sufficiently small.


Using the fact that A(e) and B(c) commute, it is easy to
show that the transformed equation satisfies all the above

hypotheses and that in addition it enjoys (H3). In particu-

lar, if the transformed equation is

x = F(x,e)
then

a3F1(0,0)

ax 0

a3F2(0,0) a3f2(0,0)
3
-3
ax ax
1 1
82 4. BIFURCATIONS WITH TWO PARAMETERS

ax
-
a3F2(0,0) a3f2(0,0)

ax +
3a3f1(0,0)
a3

Remark 2. We have assumed that f(x,e) = -f(-x,e). If we

assume that this is true only for the low order terms then we
would obtain similar results. For example, on L1 we would
get a homoclinic orbit with x1 > 0. Similarly on another
curve Li we would get a homoclinic orbit with x1 < 0. In

general L1 and Li would be different although they would


have the same linear approximation u = -(4/5)6y.

Remark 3. Suppose that we only assume (Hl) and (H3). Then

we can still obtain partial results about the local behavior


of solutions. For example, in Lemma 2 we did not use the
hypothesis B nonzero. Hence, for each b > 0, (4.1.1) has

a periodic solution through x1 = 0, x2 =


IaI-1/2(2b)l/2c1
for some e1 and e2 with e1 > 0. However, we cannot say
anything about the stability of the periodic orbits and we

cannot say how many periodic orbits (4.1.1) has for fixed e1

and 62.

Exercises

(1) Suppose that a and B are negative. Prove

that for (61,62) in regions 5 and 6, (4.1.1) has a connect-


ing orbit. (Hint: Use the calculations in Lemma 1 and

Lemma 6.)

(2) Suppose that a and g are negative and that


(61,62) is in region 3. Let U be the unstable manifold of
the point (0,0) in (4.1.1). Prove that U is in the re-

gion of attraction of the periodic orbit.


4.9. Quadratic Nonlinearities 83

(3) Suppose that a is positive and 8 is negative.

Show that the bifurcation set and the corresponding phase por-

traits are as given in Figures 3 and S.

4.9. Quadratic Nonlinearities

In this section we discuss the local behavior of solu-

tions of (4.1.1) when the nonlinearities are quadratic. Most

of the material in this section can be found in [47].


Suppose that the linearization of (4.1.1) is x = A(e)x

where A(e) is given by (4.1.2) and that

2
f(O,e) = 0, fl(0,0) 0,

2 2
ax2
a - f2(O,O) 0, 8
ax1ax2 f2(O,O) , 0.
1

We also assume that 8 > 0, a < 0; the results for the other

cases are obtained by making use of the change of variables,

t - -t, e2 -1. -e2, x1 -. ±x1, x2 -. +x2,


Introduce parameters u,6, scaled variables y1,y2

and a new time t by the relations

a-1 1/2 2 a 1/263


d = l
e1 1
x1 = 6 Y1, x2 = l l Y2,

tlal-1/26-1

e2 = 'all/26u, t =

We assume that e1 > 0 in what follows, see Exercise 8 for

the case e1 < 0.

After scaling, (4.1.1) becomes

y1 = y2 + 0(62)
(4.9.1)
2
Y2 ' Y1 + "Y2 - y1 + 6YYly2 + 0(6)
2
84 4. BIFURCATIONS WITH TWO PARAMETERS

where the dot means differentiation with respect to t, the


yi lie in a bounded set and y - a a,-1/2. Note that by
making a change of variable we can assume that (1,0) is a

fixed point of (4.9.1) for p,6 sufficiently small.

The object of the following exercises is to show that

the bifurcation set is given by Figure 6 and that the associa-

ted phase portraits are given by Figure 7.

Exercises.

(4) Let H(y1,Y2) (Y2/2) - (y2/2) + (y /3). Show


=

that along solutions of (4.9.1),

(5) Prove that there is a function u = ul(c,6),

0 < c < 1, such that if u = u1(c,6) in (4.9.1) then there


is a periodic solution through (c,0) if 0 < c < 1 and a

homoclinic orbit if c = 0.

(6) Prove that u1(c,6) - -yP(c)6 + 0(62) where

J0(c)P(c) = J1(c),

(c1
Ji(c) = Jc w1R(w)dw,

R(w) _ [w2 - (2/3)w3 + 2b] 1/2,

6b - c2(2c-3), R(c1) - 0, c1 > c.

(7) Prove that P(0) = (6/7), P(l) = 1 and P'(c) > 0

for 0 < c < 1. Deduce that for fixed u,6, (4.9.1) has at
most one periodic solution. (To prove that P'(c) > 0 we can
use the same techniques as in the proof of Lemma 4. An al-

ternative method of proving P'(c) > 0 is given in [47].)


4.9. Quadratic Nonlinearities 85

(8) If el < 0, then after scaling (4.1.1) becomes

Yl ° Y2 + 0(62)
2 2
y2 = -yl + PY2 - yl + y6Y1Y2 + 0(6 )

Put yl s z - 1, u = p - y6. Then

a y2 + 0(62)

y2 = z + PY2 - z2 + 6yzy2 + 0(62)

which has the same form as equation (4.9.1) and hence trans-

forms the results for eI > 0 into results for the case

E1 < 0.
(9) Show that the bifurcation set and the correspond-
ing phase portraits are as given in Figures 6-7.

E1

Bifurcation Set for the Case a < 0, 8 > 0.

Figure 6
86 4. BIFURCATIONS WITH TWO PARAMETERS

REGION 1

REGION 2

Phase Portraits for the Case a < 0, 8 > 0. (For the phase
portraits in regions 4-6 use the transformations in Exer-
cise 8.)

Figure 7
4.9. Quadratic Nonlinearities 87

ON L

REGION 3

Figure 7 (cont.)
CHAPTER 5

APPLICATION TO A PANEL FLUTTER PROBLEM

5.1. Introduction

In this chapter we apply the results of Chapter 4 to a


particular two parameter problem. The equations are

z = Ax + f(x) (5.1.1)

where

x - [xl,x2,x3,x4]T, f(x) = [f 1(x),f2(x),f3(x),f4(x)]T,

r 0 1 0 0 1

al b1 c 0
A --
0 0 0 1

-c 0 a2 b2

f1(x) = f3(x) = 0,

f2(x) = xlg(x),

f4(x) = 4x3g(x),

2g(x) = -n4(kx2 + axlx2 + 4kx2 + 4ax3x4),

c = 8p,
3 a j
-n2j2[n2j2 + r],

bj - -[an4j4 + T 6]; a - 0.005, 6 = 0.1,

as
5.2. Reduction to a Second Order Equation 89

k > 0, a > 0 are fixed and p,r are parameters. The above

system results from a two mode approximation to a certain


partial differential equation which describes the motion of a
thin panel.

Holmes and Marsden [36,38] have studied the above equa-


tion and first we briefly describe their work. By numerical

calculations, they find that for p = p0 = 108, r = r0 =

-2.237r2, the matrix A has two zero eigenvalues and two

eigenvalues with negative real parts. Then for lp-p01 and

Ir-r01 small, by centre manifold theory, the local behavior


of solutions of (5.1.1) is determined by a second order equa-

tion depending on two parameters. They then use some results

of Takens [64] on generic models to conjecture that the local


behavior of solutions of (5.1.1) for lp-p01 and Ir-r0l

small can be modelled by the equation

u + au + bu + u2u + u3 = 0

for a and b small.

Recently, this conjecture has been proved by Holmes

[37], in the case a = 0, by reducing the equation on the


centre manifold to Takens' normal form. We use centre mani-

fold theory and the results of Chapter 4 to obtain a similar

result.

5.2. Reduction to a Second Order Equation

The eigenvalues of A are the roots of the equation

a4 + d1A3 + d2a2 + d3A + d4 = 0 (5.2.1)

where the di are functions of r and p. If A has two

zero eigenvalue% then d3 - d4 - 0. A calculation shows that


90 S. APPLICATION TO A PANEL FLUTTER PROBLEM

if d3 - d4 - 0, then

a1a2 + c2 = 0
(5.2.2)
a1b2 + b1a2 - 0

or in terms of r and p,

64
4n4(n2+r)(4n2+r) + p2 = 0 (5.2.3)

(16an4+6p1/2)(n2+r) + 4(an2+6p1/2)(4n2+r) - 0. (5.2.4)

We prove that (5.2.3), (5.2.4) has a solution r - r0,


p p0'
From (5.2.3) we can express p in terms of r. Sub-

stituting this relation into (5.2.4) we obtain an equation


H(r) = 0. Calculations show that H(r1) < 0, H(r2) > 0 where
rl - -(2.225)r2 and r2 = -(2.23),r2, so that H(r0) = 0 for

some r0 E (r2,r1). Further calculations show that (5.2.3),

(5.2.4) has a solution r0,p0 with 107.7 < p0 < 107.8.

In the subsequent analysis, we have to determine the


sign of various functions of r0 and p0. Since we do not

know r0 and p0 exactly we have to determine the sign of


these functions for r0 and p0 in the above numerical
ranges.

When r = r0, p = p0, the remaining eigenvalues of A


are given by

(b1+b2) t [(b1-b2)2 + 4(a1+a2)]1/2


X3,4

and a calculation shows that they have negative real parts

and non-zero imaginary parts.

We now find a basis for the appropriate eigenspaces

when r = r0, p - p0. Solving Av1 - 0 we find that


5.2. Reduction to a Second Order Equation 91

vl - [1,0,-al/c,0]T. (5.2.5)

The null space of A is in fact one-dimensional so the can-

onical form of A must contain a Jordan block. Solving

A2v2 - 0 we obtain

v2 = [0,1,-b1/c,-a1/c]T, Av2 = v1. (5.2.6)

The vectors vi and v2 form a basis for the generalized

eigenspace of A corresponding to the zero eigenvalues.

Similarly, we find a (real) basis for the space V spanned

by the eigenvectors corresponding to A3 and A4. Solving

Az - A3z, we find that V is spanned by v3 and v4 where

2v3 - z + z, 2v4 = i(z-z),

z - [1,A3,w,A3w]T (5.2.7)

we - b A - b1b2 + a2.
2 3

Let A0 denote the matrix A when I' = r0 and

p = p0. Let S = [vl,v2,v3,v4] where the vi are defined


by (5.2.5), (5.2.6) and (5.2.7) and set y = S-lx. Then

(5.1.1) can be written in the form

jy = By + S-1(A-A0)SY + F(Y,r,P) (5.2.8)

where F(y,I',p) = S-lf(Sy),

0 1 0 0 1

0 0 0 0

0 0
pl p2
0 0
-p2 pl JI

and where aS - Pi + ip2, pl < 0, p2 t 0.


92 S. APPLICATION TO A PANEL FLUTTER PROBLEM

Then for Ir-r0I and Ip-p0I sufficiently small

(5.2.8) has a centre manifold y3 = hl(yl'y2'r'p)'

h2(yl,y2,r,p). The flow on the centre manifold is gov-


y4 =
erned by an equation of the form

yl 0 1 Y1 Y1
] = [ ][ E(r,P) [
Y2 0 0 y2 y2
(5.2.9)

+ N(Y1,Y2,r,P)

where E(r,p) is a 2 x 2 matrix with E(r0,p0) - 0 and

N(yl,y2,r,p) contains no linear terms in yl or y2. We

show that there is a nonsingular change of variables (r,p)

(c1,c2) for (r,p) close to (r0,p0) and a r,p dependent

change of variables such that the lin-


(yl,y2) i (yl'y2)
earized equation corresponding to (5.2.9) is

yl 0 1
yl
(5.2.10)
Y2 E1 c2 y2

The transformation is of the form


(yl,y2) i (yl'y2)
Identity + O(Ir-r0I +Ip-POI) and r - rot p = p0 is mapped

into el = 0, c2 = 0. After these transformations, (5.2.9)

takes the form

yl a 0 1 'l + N(y1,Y2,r0,P0)
Y2 c2 Y2
E1
(S. 2.11)

+ N(Y1,Y2,r,P)

where we have dropped the bars on the yi. The function N

will contain no linear terms in yl or y2 and

N(yl'y2'r0,p0) = 0. Since the nonlinearities in (5.1.1) are

cubic, the same will be true of N and A.


5.3. Calculation of Linear Terms 93

Let

N(Y1,Y2,ro,PO) _ [N1(Y1,Y2,ro,PO), N2(Y1,Y2,r0,P0)]T (5.2.12)

and let
3
al = -s N2(o,o,ro,PO)
ayl
3
ay3 N1(o,o,ro,PO)
(5.2.13)
1

3
S = a N2(o,o,r0,PO)
aylay2

8 = 3r + S.

Using the results in Chapter 4 (see, in particular, Remark 1


in Section 8), if a1 and 0 are non-zero, we can deter-

mine the local behavior of solutions of (5.2.11). By Theorem

2 of Chapter 2, this determines the local behavior of solu-

tions of (5.2.8).

5.3. Calculation of the Linear Terms

From (5.2.1), trace(A) = d3(r,p), det(A) = d4(r,p)

where

d3(r,P) = 4n2(n2+r) (ant+r) + 69 p2

d4(r,P) - n2(16an4+6p1/2)(n2+r) + 4n2(air2+6P1/2)(4n2+r)

Calculations show that the mapping

(r,p) (d3(r,P),d4(r,P)) (5.3.1)

has non-zero Jacobian at (r,p) = (ro,po).

Define the matrix C(r,p) by


94 S. APPLICATION TO A PANEL FLUTTER PROBLEM

C(r,p) = + E(r,p)

and let J be the value of the Jacobian of the mapping

(r,p) -. (trace(C(r,p)), det(C(r,p))), (5.3.2)

evaluated at (r0,p0). By considering the 4 x 4 matrix

B + S-1(A-AO)S, it is easily seen that J is a non-zero

multiple of the Jacobian of the mapping given by (5.3.1).

Hence J is non-zero, so by the implicit function theorem we

can use E1 = -det(C(r,p)), E2 ' trace(C(r,p)) as our bifur-

cation parameters. Approximate formulae for E1 and E2

can easily be found if so desired.

Let C(r,p) - [cij],

M =
[:11 :12]

yl -M Y1

Y2 Y2
J

Then the linearized equation corresponding to (5.2.9)

is

yl 0 1
yl
y2 E1 E2
Y2

Note that M is equal to the identity matrix when

e1 = E2 = 0.
5.4. Calculation of the Nonlinear Terms 95

5.4. Calculation of the Nonlinear Terms

We now calculate the nonlinear term in (5.2.9) when


r - rot p - p0. Since the nonlinearities in (5.1.1) are
cubic, the centre manifold has a "cubic zero" at the origin.

Using x - Sy, on the centre manifold

xl - Yl + 03, x2 y2 + 03

-a bl (5.4.1)
x3 - yl - E Y2 + 03, x4 al
c Y2 + 03
where 03 - 0(IY1I3 + IY2I3).
Let S-1 - [t..] and let

Fj(Y1,Y2) - fj(Y1,Y2,hl(Y1,Y2,r0,PO)).

Then using the notation introduced in (5.2.12),

N1(Y1,Y2,r0,PO) ' t12F2(Y1,Y2) + t14F4(Y1,Y2)


(5.4.2)
N2(Y1,Y2,r0,PO) - t22F2(Y1,Y2) + t24F4(Y1,Y2)

Using (5.4.1),

al 8ka1b1 2
n4 3 2 3 2
F2(Y1,Y2) ' - [kyl + 4k(c) yl + ayly2 + yly2
c
+ 4aal 2
Y1Y2]
2
+ terms in
yly2 + 05

4
-4c 1
F2(Yl,y2) + 2bcn rkyiy2
F4(Yl,y2) = + 4k( c1)2 YlY21

+ terms in yly2 and y3 + 05

where 05 - 0(Iyl1S + IY21 ). Note also that since r - r0


and p - pot (5.2.2) holds.

From (5.2.13) and (5.4.2),


96 5. APPLICATION TO A PANEL FLUTTER PROBLEM

a,
-T- k(l + 4 (c) 2(t22
- 74
al - 4 ac t24)
4 a a
r k(1 + 4(c)2)(t12 4
- c t14)
2
_ _n4 8kalbl 4aa1 al
2ir4
(a + + )(t22 - 4 c t24) + -3b1kt24(c 2+4a1).
2

Routine calculations show that

t12 = m2b2(2a1-b1b2)

t22 = mat

t14 = m2c(bl+b2)

t24 = -mc

m = (al+a2-blb2)-1 < 0.

Using (5.2.2) and numerical calculations, we find that


a
t22 - 4 c t24 = m(4a1+a2) > 0

t14 a b1m2(2a2-bZ-4a1) > 0


t12 - 4 ac

so that al and r are negative. Similarly, the coefficient

of k in is

7 a1b1m(12a1+a2) < 0.
c

Hence a1 and a are negative and the local behavior of solu-

tions of (5.1.1) can be determined using the results of

Chapter 4.
CHAPTER 6
INFINITE DIMENSIONAL PROBLEMS

6.1. Introduction

In this chapter we extend centre manifold theory to a


class of infinite dimensional problems. For simplicity we
only consider equations of the form

w = Cw + N(w), w(O) E Z

where Z is a Banach space, C is the generator of a strongly


continuous semigroup on Z and N: Z - Z is smooth. In the

next section we give a brief account of semigroup theory.


For additional material on semigroup theory see [6,43,50,55,
561. For generalizations to other evolution systems see
[34,51).

6.2. Semigroup Theory

In earlier chapters we studied centre manifold theory

for the finite dimensional system

w - Cw + N(w). (6.2.1)

The most important tools for carrying out this program were:

97
98 6. INFINITE DIMENSIONAL PROBLEMS

(1) The solution of the linear problem = Cw and estimates


of exp(PCt), exp((I-P)Ct), where P is the projection

onto the space associated with eigenvalues of C with

zero real parts.

(2) The variation of constants formula

rt
w(t) = exp(Ct)w(O) + exp(C(t-s))N(w(s))ds
J
0

for the solution of (6.2.1) and Gronwall's inequality.

To carry out this program for partial differential

equations we have to study ordinary differential equations in

infinite dimensional spaces. We first study linear problems.

Let Z be a Banach space and C a linear operator

from some domain D(C) of Z into Z. We wish to solve the


problem

w = Cw, t > 0, (6.2.2)

w(0) = w0 E Z. (6.2.3)

Suppose that for each w0 E Z the above equation has a

unique solution w(t). (We define later the meaning of the

term solution). The solution w(t) is a function of t and

w0 and we write w(t) = T(t)w0. Then T(t) is a linear

mapping from Z into Z with T(0) = I. If w(t) is a

solution of (6.2.2) - (6.2.3), then for any s > 0, w(t)

w(t+s) = T(t+s)w0 is a solution of (6.2.2) with w(0) - w(s).

Hence T(t+s)w0 = T(t)T(s)w0. If solutions depend continu-

ously on initial data then T(t)wn T(t)w whenever wn - w


in Z, so that T(t) must be bounded. Finally, we require
that T(t)w0 + w0 as t - 0+.
6.2. Semigroup Theory 99

Definition. A one parameter family T(t), 0 < t < of

bounded linear operators from Z into Z is a (strongly


continuous) semigroup if

(i) T(O) = I,

(ii) T(t+s) = T(t)T(s), t,s > 0 (semigroup property),

(iii) IIT(t)w - wII- 0 as t - 0 for every w E Z.

Example 1. Let L(Z) denote the set of bounded linear opera-


tors from Z into Z and let C E L(Z). Define T(t) by

n
T(t) = eCt L Cn.
n=0

The right hand side converges in norm for each t > 0 and it

is easy to verify conditions (i), (ii), (iii). Thus eCt

defines a semigroup.

Example 2. Let Z be a Banach space of uniformly continuous

bounded functions on [0,m) with the supremum norm. Define

(T(t)f)(8) = f(O+t), f E Z, 8 > 0, t > 0.

Conditions (i) and (ii) are obviously satisfied and since

IIT(t)f - f11 = sup{If(0+t) - f(8)I: 8 > 0} - 0 as t - 0+,

(iii) is satisfied. Hence T(t) forms a semigroup.

Definition. The infinitesimal generator C of the semigroup


T(t) is defined by
T(t)z - z
Cz = lim+
t+0 t

whenever the limit exists. The domain of C, D(C), is the set


of all elements z E Z for which the above limit exists. We

also say that C generates T(t).


100 6. INFINITE DIMENSIONAL PROBLEMS

In Example 1 the infinitesimal generator is C E L(Z)

while in Example 2 the infinitesimal generator is

(Cf) (e) = f' (e) , D(C) - {f: f' E Z}.


Exercise 1. If C E L(Z) prove that IleCt - 111- 0 as

t - 0.
Exercise 2. Let Z - R2 and define T(t): Z - Z by

(e-ntan),
T(t)(an) = (an) E Z, t > 0.

Prove that T(t) is a semigroup on Z with infinitesimal

generator given by

C(an) = (-nan), D(A) - {(an): (nan) E Z).

Does II T (t) - III - 0 as t - 0+?


Exercise 3. Prove that if T(t) is a semigroup then

IIT(t)II < Mewt, t > 0, (6.2.4)

for some constants w > 0, M > 1. (Hint: Use the uniform


boundedness theorem and property (iii) to show that IIT(t)11

is bounded on some interval [O,e]. Then use the semigroup

property.)

Exercise 4. Find a 2 x 2 matrix C such that if

Il eCt Il < M for all t > 0 then M > 1.

Exercise S. Let C be the generator of a semigroup T(t)

which satisfies IIT(t)II < Mewt, t > 0. Prove that C - mI

is the generator of the semigroup S(t) - e-WtT(t) and that

11 S(t) 11 < M for all t > 0.


6.2. Semigroup Theory 101

Exercise 6. If T(t) is a semigroup, prove that for each

w E Z, T(t)w is a continuous function from [0,m) into Z.

Let T(t) be a semigroup with generator C. Now

h-1[T(t+h)w - T(t)w] = h-1[T(h) - I]T(t)w


(6.2.5)
- T(t)h-1[T(h)w - w].

If w E D(C) then the right side of (6.2.5) converges to

T(t)Cw as h - 0+. Thus the middle term in (6.2.5) con-

verges as h - 0+ and so T(t) maps D(C) into D(C). Also

the right derivative satisfies

dT t w = CT(t)w - T(t)Cw, t > 0, w E D(C). (6.2.6)


t

Similarly, the identity

h-1[T(t)w - T(t=h)]w - T(t-h)h-1[T(h)w - w]

shows that (6.2.6) holds for the two-sided derivative.

Exercise 7. For w E Z, prove that w(t) E D(C) where

t
w(t) = T(s)w ds,
J0

and that t-1w(t) - w as t + 0+. Deduce that D(C) is

dense in Z.

Exercise 8. If wn E D(C), then from (6.2.6)


t
T(t)wn - wn = T(s)Cwnds.
J0

Use this identity to prove that C is closed.

Equation (6.2.6) shows that if w0 E D(C) and C is

the generator of a semigroup T(t), then w(t) = T(t)w0 is a

solution of the Cauchy problem (6.2.2) - (6.2.3). In applica-


102 6. INFIMITE DIMENSIONAL PROBLEMS

tions to partial differential equations, it is important to

know if a given operator is the generator of a semigroup. To

see the problems involved, suppose that C generates the con-

traction semigroup T(t), that is IIT(t)II < 1 for all t > 0.

The Laplace transform

R(a)w = J e-AtT(t)w dt, (6.2.7)


0

exists for Re(A) > 0 and IIR(A)II< A 1 for A > 0. Since

T(t) is in some sense exp(Ct), we expect R(A) to be the

resolvent of C, that is, (XI - C)-1. This is indeed the case

and is easy to prove. The main problem is in finding the in-

verse Laplace transform, that is, given C such that

(XI - C)-1 exists for A > 0, say, is C the generator of a

semigroup? The basic result is the Hille-Yosida Theorem.

Theorem 1 (Hille-Yosida Theorem). A necessary and sufficient

condition for a closed linear operator C with dense domain

to generate a semigroup of contractions is that each A > 0

is in the resolvent set of C and that

I IR(A) I I = I I ( A I - C ) - 1 11 <A1 for A > 0. (6.2.8)

The reader is led through the proof of the above

theorem in the following exercises.

Exercise 9. Let T(t) be a contraction semigroup and for

A > 0 define R(A) as in (6.2.7) for w E Z. Use the identity

Ah
h-1(T(h)-I)R(A)w = h-'[(e - 1)j e-AtT(t)wdt
0
h
- eah
1
e-AtT(t)w dt]
0

to prove that R(A) maps Z into D(C) and that


6.2. Semigroup Theory 103

(XI - C)R(A) = I.

Prove also that for w E D(C) , R(X) (XI - C)w = w and deduce

that R(X) is the resolvent of C.

Exercise 10. Suppose that C is a closed linear operator on

Z with dense domain such that (0,m) is in the resolvent

set of C with (6.2.8) satisfied. For A > 0 define

C(X) E L(Z) by

C)-1 C)-1
C(X) = XC(XI - = X2(XI - - XI.

Prove that

(i) C(a)w - Cw as A for w E D(C),

(ii) Il exp (tC(X)) II < 1 and


Ilexp(tC(X))w - exp(tC(u))wII < tlIC(X)w - C(u)wII

for t > 0, X,u > 0 and w E Z.

Use the above results to define T(t) by

T(t)w = lim exp(C(A)t)w


aim

and verify that T(t) is a semigroup of contractions on Z

with generator C.

Exercise 11. Let Z = R2 and let {an} be a sequence of


complex numbers such that a = sup{Re(an): n = 1,2,...} <
Define C by

C(an) = (an an ), D(C) = {(an): (an an) E Z}.

Use the Hille-Yosida Theorem to prove that C - aI generates

a contraction semigroup and deduce that C generates a semi-

group T(t) with IIT(t)II < eat for t > 0.


104 6. INFINITE DIMENSIONAL PROBLEMS

Remark. The above version of the Hille-Yosida Theorem char-

acterizes the generators of semigroups which satisfy


eWt
IIT(t)II < (see Exercise 5). A similar argument gives the

characterization of the generators of all semigroups.

Remark. Let C be a closed operator with dense domain such


that all real nonzero A are in the resolvent set of C and

II(AI - C)_111.' IAI 1 for such A. Then C is the generator

of a group T(t) , t EIR, that is, T(0) = I, T(t+s) = T(t)T(s)

all t,s E IR and T(t)z - z as t + 0 for each z E Z.


Example 3. Let C be a selfadjoint operator on a Hilbert

space H with C < 0 . Then 11 (XI - C) 1 1 1 < A 1 for A > 0,

so by the Hille-Yosida Theorem, C generates a contraction


semigroup.

Example 4. Let A = iC where C is a selfadjoint operator

on a Hilbert space H. Then for all nonzero real A,

I I (AI -A)_111 < IAI so that A generates a unitary group.


Let f E C([0,T];Z) and let C be the generator of a
semigroup T(t) on Z. Then formally the solution of

w = Cw + f(t), 0 < t < T, w(0) = w0, (6.2.9)

is given by the variation of constants formula

t
w(t) = T(t)w0 + I
T(t-s)f(s)ds. (6.2.10)
0

If f is continuously differentiable and w0 E D(C) then it

is easy to check that w(t) defined by (6.2.10) satisfies

(i) w(t) is continuous for 0 < t < T,

(ii) w(t) is continuously differentiable for

0 < t < T,
6.2. Semigroup Theory 105

(iii) w(t) E D(C) for 0 < t < T and (6.2.9) is sat-

isfied.

A function w(t) which satisfies (i), (ii), (iii) is


said to be a strong solution of (6.2.9). It is easy to check
that such a solution is unique in this class. If w0 t D(C)

or f is only continuous, then in general w(t) defined by

(6.2.10) is not in D(C) so (6.2.9) does not make sense. We

now define what we mean by a solution in this case.

Definition. A function w E C([0,T];Z) is a weak solution

of (6.2.9) if w(0) = w0 and if for every v E D(C*), the

function <w(t),v> is absolutely continuous on [0,T] and

dt<w(t),v> = <w(t),C*v> + <f(t),v> a.e.

where C* is the adjoint of C and < , > denotes the pair-

ing between Z and its dual space.

The proof of the following theorem can be found in 191.

Theorem 2. The unique weak solution of (6.2.9) is given by


(6.2.10).

A semigroup T(t) is said to be an analytic semigroup


if in addition the map t - T(t)z is real analytic on (0,W)

for each z E Z. Analytic semigroups have many additional


properties. In particular, if C is the generator of an analy

tic semigroup T(t), then T(t) maps Z into D(C) for each

t > 0. Thus if f is continuously differentiable, then w(t)

defined by (6.2.10) is a strong solution of (6.2.9) for each


w0 E Z if T(t) is an analytic semigroup. An example of a

generator of an analytic semigroup is a nonpositive selfad-


joint operator and in general, solutions of parabolic equations
106 6. INFINITE DIMENSIONAL PROBLEMS

are usually associated with analytic semigroups. Hyperbolic

problems, however, do not in general give rise to analytic

semigroups. To see this, suppose that C is the generator

of a group T(t) and that T(t) maps Z into D(C) for

t > 0. By the closed graph theorem, CT(l) E L(Z) and hence

C = CT(1)T(-l) is a bounded linear operator.

In applications, it is sometimes difficult to prove


directly that C + A is a generator, although C is a gen-

erator and A is in some sense small relative to C. The

simplest such result is the following.

Theorem 3. Let C be the generator of a semigroup (group)

T(t) and let A E L(Z). Then C + A generates a semigroup


(group) U(t). If IIT(t)II < McWt for t > 0 then

IIU(t) II < Me(-+MIIAII)t for t > 0.


Outline of the Proof of Theorem 3: If A + C is the genera-

tor of U(t), then by the variation of constants formula, U(t)

must be the solution of the functional equation


t
U(t) - T(t) + T(s)AU(t-s)ds. (6.2.11)
J
0

Equation (6.2.11) is solved by the following scheme:

U(t) - F Un(t) (6.2.12)


n-0

where U0(t) - T(t) and

rt
Un+1(t) - T(s)AUn(t-s)ds.
0

A straightforward argument shows that U(t) defined by

(6.2.12) is a semigroup with generator A + C.

Exercise 12. Prove Theorem 3 in the case M - 1, w - 0 in


6.2. Semigroup Theory 107

the following way: let B = A + C so that B is a closed


operator with dense domain D(B) - D(C). Show that for

A > IIAII, IIA(AI - C)_111 <1 so that (I - A(AI - C)-1)-1

exists for A > IIAII Hence show that II (AI - B)_111 <

(A - IIAII)-1 for A > IIAII and use the Hille-Yosida Theorem.

Example 5 (Abstract Wave Equation). Let A be a positive


selfadjoint operator on a Hilbert space H and let B E L(H).

Consider the equation

v + B4 + Av - 0. (6.2.13)

Define a new Hilbert space Z = D(A1/2) x H with inner pro-


duct < , > defined by

<w,w> - (A1/2w1,A1"2w1) + (w2,w2)

where ( , ) is the inner product in H and w = [wl,w21,

w = [wl,w21. Equation (6.2.13) can now be written as a first


order system on Z,

w = Cw (6.2.14)

where

C -

D(C) = D(A) x D(A1/2). If B - 0 then C is skew-selfad-

joint so by Example 4 it generates a group. For B # 0, C is

is the sum of a generator of a group and a bounded operator


so by Theorem 3 it generates a group.

As a concrete example of the above, consider the wave


equation
108 6. INFINITE DIMENSIONAL PROBLEMS

vtt + vt - Ov = 0, x E SI, v(x,t) = 0, x E BSI,

where D is a bounded open subset of IRn with boundary BSI.

This equation has the form (6.2.13) if we set H L2(SI),

D(A) _ {w E H: -Ow E H, w = 0 on 801, A = -0, B I. In


this case Z = H01 (0) x L2(SI) (see [2] for a definition of
1
the Sobolev space H0 (S2)).

Let T(t) be a contraction semigroup on Hilbert space

H with generator C. Then for s > 0,

IjT(s+t)zII _ IIT(t)T(s)zjl < IjT(t)zjj

so that IIT(t)zII is nonincreasing. For z E D(C), v(t)

11T(t) zII 2 is differentiable in t and

v1(t) _ <T(t)Cz, T(t)z> + <T(t)z, T(t)Cz>.

Since v'(0) < 0, we have that

<Cz,z> + <z,Cz> < 0, z E D(C). (6.2.15)

A linear operator C with dense domain is said to be dissi-

pative if (6.2.15) holds.

Exercise 13. If C is dissipative and is the generator of a

semigroup, prove that the semigroup is a contraction.

Theorem 4 (Lumer-Phillips Theorem). Let C be a dissipative

operator on the Hilbert space H. Suppose that for some


a > 0 the range of BI - C is H. Then C is the generator

of a contraction semigroup.

Exercise 14. Prove the above theorem. (Hint: deduce from

(6.2.15) that II (XI-C) z II > Ali z it for A> 0 and z E D(C).


Use this and the fact that the range of BI - C is II to
6.2. Semigroup Theory 109

deduce that C is closed. Show that the non-empty set


(X > 0: range of XI - C is H) is open and closed and use
the Hille-Yosida Theorem.)

Remark. There is also a Banach space version of the above

result which uses the duality map in place of the inner pro-
duct.

Exercise 15. Let C be the generator of a contraction semi-

group on a Hilbert space and let B be dissipative with


D(B) D(C) and

IIBwll < alUCwll + bllwll , w E D(C)

for some a with 0 < 2a < 1. Use the inequality


IIB(AI - C) lwll < (2a + bX 1)Ilwll to show that AI - B - C is
invertible for large enough A. Deduce that B + C generates

a contraction semigroup.

Exercise 16. Let C be a closed linear operator on a Hil-


bert space H and suppose that C and C* are dissipative.

Prove (i) the range of I - C is closed in H,

(ii) <h - Ch,z> - 0 for all h E D(C) implies z = 0. De-

duce that the range of I - C is H so that by Theorem 4,

C generates a contraction semigroup on H.

If all the eigenvalues of a matrix C have negative


real parts then Ilexp(Ct)II < Met, t > 0, for some M > 1

and w > 0. The corresponding question in infinite dimen-


sions is: if C is the generator of a semigroup T(t), does

Re(a(C)) < 0 imply that IIT(t)II < Me-wt, t > 0, for some

w > 0 and M > 1? (a(C) denotes the spectrum of Q. In

general, the answer is no. Indeed, it can be shown that if


110 6. INFINITE DIMENSIONAL PROBLEMS

a < b then there is a semigroup T(t) on a Hilbert space

such that sup Re(a(C)) - a and IIT(t)II = ebt [69]. Thus to

obtain results on the asymptotic behavior of the semigroup in

terms of the spectrum of the generator, we need additional

hypotheses.
W t
If e 0 is the spectral radius of T(t), that is

eWOt - lim IIT(t)kIIl/k,

then a standard argument shows that for m > m0 there exists

M(w) such that IIT(t)II < M(w)eWt. Hence we could determine


the asymptotic behavior of T(t) from the spectral radius

of T(t). However the spectrum of T(t) is not faithful to

the spectrum of C, that is, in general the mapping relation

a(T(t)) - exp(ta(C)) (6.2.16)

is false. While (6.2.16) is true for the point and residual


spectrum (55] (with the possibility that the point 0 must

be added to the right hand side of (6.2.16)) in general we


only have

continuous spectrum of T(t) exp(t(continuous spectrum of C)).

R2
Example 6. Let Z - and C(xn) - (inxn), D(C) =
{(xn) E Z: (nxn) E Z}. Then C generates the semigroup

T(t) given by T(t)(xn) _ (eintxn). The spectrum of C is

{in: n - 1,2,...,} while the spectrum of T(l) is the unit

circle so (6.2.16) is false.

For special semigroups, e.g., analytic semigroups,

(6.2.16) is true (the point zero must be added to the right

hand side of (6.2.16) when the generator is unbounded), but


6.2. Semigroup Theory 111

this is not applicable to hyperbolic problems. For the prob-


lem studied in this chapter, the spectrum of the generator
consists of eigenvalues and the associated eigenfunctions
form a complete orthonormal set. Thus the asymptotic be-

havior of the semigroup could be determined by direct compu-


tation. It seems worthwhile however to give a more abstract
approach which will apply to more general problems.

Let C be the generator of a semigroup T(t) with

IIT(t)II< McWt. Hence the spectrum of T(t) lies in the disc

IaI < exp(ort). Let A E L(Z). Then A + C generates a semi-

group U(t). Suppose that there are only isolated eigen-


values of A + C in Q - {A: Re(A) > w}. Then if A E Q,

eXt is an eigenvalue of U(t). We prove that if A is com-


emt}
pact then the spectrum,of U(t) in the set {A: IXI >

consists only of points e?t, A E Q.

Theorem 5 (Vidav [67], Shizuta [62]). Let C be the genera-

tor of a semigroup T(t) with IIT(t)II < McWt and let

A E L(Z) be compact. Then A + C generates a semigroup

U(t) such that U(t) - T(t) is compact.

Proof: From Theorem 3, A + C generates a semigroup U(t)

which is a solution of the functional equation

rt
U(t) - T(t) + T(s)AU(t-s)ds. (6.2.17)
I
0

We prove that the map s - f(s) = T(s)AU(t-s) is continuous

in norm on [O,t]. Thus the integral in (6.2.17) converges

in norm. Since f(s) is a compact linear operator and the


set of compact operators in L(Z) is closed in the uniform
operator topology, U(t) - T(t) is compact.
112 6. INFINITE DIMENSIONAL PROBLEMS

To prove the claim on the continuity of f we first

prove that s -r AU(t-s) is continuous in norm. Let

t > s > 0 and h1 > 0 with h1 sufficiently small. Then

the set

{A(U(t-s-h) - U(t-s))w: IIwII = 1, IhI < hl}

has compact closure. For e > 0 it can be covered with a

finite number of balls with radius a and centres at

wl,w2,...,wn. Let g(s) = AU(t-s). Then for IIwII = 1,

Ilg(s+h)w - g(s)wII < IIA(U(t-s-h) - U(t-s)) (w-wk) II


+ IIA(U(t-s-h) - U(t-s))wkll

Hence if 6 is chosen such that

IIA(U(t-s-h) - U(t-s))wkll < E

for IhI < 6 and k = 1,2,...,n then IIg(s+h) - g(s)II <

(constant) E. The continuity of f follows from similar

reasoning applied to

f(s+h) - f(s) - T(s+h)[g(s+h)-g(s)] + (T(s+h)-T(s))AU(t-s).

This completes the proof of the theorem.

Since U(t) - T(t) is compact, the essential spectrum

of U(t) and T(t) coincide [43]. (It is important to note

that we are using Kato's definition of essential spectrum in

this claim.) Hence we have proved:

Corollary to Theorem S. The spectrum of U(t) lying outside


emt
the circle IxI = consists of isolated eigenvalues with

finite algebraic multiplicity.


6.2. Semigroup Theory 113

The above result is also useful for decomposing the


space Z. Suppose that the assumptions of Theorem S hold.
Since there are only isolated points in the spectrum of U(t)

outside the circle IXI = eWt, there is only a finite number


of points in the spectrum of A + C in the halfplane

Re(X) > m + e for each e > 0. Let A1,...''n denote these

eigenvalues for some fixed e > 0 and let P be the corres-

ponding projection. Then we can decompose U(t) into the

sum PU(t) + (I-P)U(t) and for any 6 > e there exists

M(6) such that ii(I-P)U(t)ii < M(6)eN+6)t Finally, PZ is

finite dimensional and so it is trivial to make further de-


compositions of PU(t).

We now show how the above theory can be applied to


hyperbolic problems. Let H be a Hilbert space with inner
product ( , ). Let A be a positive selfadjoint operator

on H such that A-1 is defined on all of H and is com-

pact. Consider the equation

v + 2a' + (A+B)v = 0

where B E L(H) and a > 0. Let Z be the Hilbert space

D(A1/2) x H with inner product < , > as defined in Example

S. We can recast the above equation in the form

w = Cw

r0 I
C =
A-B -2aI ].
As in Example 5, C is a bounded perturbation of a skew-
selfadjoint operator and so C generates a group S(t). Let
114 6. INFINITE DIMENSIONAL PROBLEMS

1
I 01 I 0
U(t) S(t)
IJ
I

aI aI I

Then U(t) is a group with generator

L
-A
aI I
-aI
r 0

a2I-B
0

Cl + C2

and the compactness of the injection D(A1/2) -r H implies

that C2: Z + Z is compact. Thus the above theory applies.

Since C1 generates a group U1(t) with HU1(t)II < e-at,

the asymptotic behavior of U(t) (and hence S(t)) depends

on a finite number of eigenvalues.

The above theory applies, for example, to the case


H = L2(0), where fl is a bounded domain in IRn, A . -D,

D(A) _ {v E H: Av E H and v - 0 on 9c} since by standard

elliptic theory, A-1 is compact.

Example 7. Consider the coupled set of wave equations


Tr

utt + taut - uxx + of g(x,s)u(s,t)ds - By = 0


0 (6.2.18)
vtt + 2avt - vxx . 0

for 0 < x < Tr with u = v = 0 at x = 0,Tr, where

g(x,s) = E n 2 sin nx sin ns.


n=1

As in Example 5 we can write (6.2.18) as a first order system

w = Cw

on Z = (H0(0,7r) x is a bounded pertur-


L2(O,Tr))2. Since C

bation of a skew-selfadjoint operator it generates a


6.2. Semigroup Theory 115

group T(t).

An easy computation shows that the eigenvalues of C


- a n) 1/2,
are an = -a ± (a2 n = 1,2,..., where an = n2 or

an = n2 + (Bn)/(2n2). For a > 0 and B small enough the


real parts of all the eigenvalues are negative so we expect
that IIT(t)II < Me--t, t > 0, for some w > 0. This formal

argument can be rigorized by applying the above theory.

We now suppose that a = 0. In this case, for B

sufficiently small all the eigenvalues of C are purely ima-

ginary and they are simple. By analogy with the finite di-

mensional situation we expect that IIT(t)II < M, t > 0, for

some constant M. We show that this is false.

Consider the following solution of (6.2.18):

u(x,t) = 2n-1 m(cos mt - cos amt)sin mx


(6.2.19)
v(x,t) = m-l sin mx cos mt

where m is an integer and am = m2 + (Bn)/(2m2). Note that

the initial data corresponding to (6.2.19) is bounded inde-

pendently of in. Let tm 2m3n. Then for large m,

2
cos mtm - cos amtm = 1 - cos(n26) + 0(m 4) > constant,

so that Iux(x,tm)I > (constant)m2. Thus IIT(tm)II

(constant)m2.

The above instability mechanism is associated with the

fact that for a = 0, B # 0, the eigenfunctions of C do not

form a Riesz Basis. For further examples of the relationship

between the asymptotic behavior of solutions of w = Cw and

the spectrum of C see [15,16].

We now consider the nonlinear problem


116 6. INFINITE DIMENSIONAL PROBLEMS

w = Cw + N(w), w(0) = w0 E Z, (6.2.20)

where C is the generator of a semigroup T(t) on Z and

N: Z - Z. Formally, w satisfies the variation of constants

formula
rt
w(t) = T(t)w0 + T(t-s)N(w(s))ds. (6.2.21)
J
to

Definition. A function w E C([O,T];Z) is a weak solution

of (6.2.20) on [0,T] if w(0) = w0; E L1([O,T];Z)

and if for each v E D(C*) the function <w(t),v> is ab-

solutely continuous on [0,T] and satisfies

Ut <w(t),v> = <w(t),C*v> + <N(w(t)),v> a.e.

where C* is the adjoint of C and < , > denotes the pair-

ing between Z and its dual space.

As in the linear case, weak solutions of (6.2.20) are

given by (6.2.21). More precisely:

Theorem 6 [9]. A function w: [0,T] -. Z is a weak solution

of (6.2.20) on [O,T] if and only if E L1([O,T];Z)

and w is given by (6.2.21).

As in the finite dimensional case, it is easy to solve

(6.2.20) using Picard iteration techniques.

Theorem 7 [61]. Let N: Z -. Z be locally Lipschitz. Then

there exists a unique maximally defined weak solution

w E C([O,T);Z) of (6.2.20). Furthermore, if T < W then

I Iw(t) I I -- as t-T . (6.2.22)

As in the finite dimensional case, (6.2.22) is used as

a continuation technique. Thus if for some w0 E Z, the


6.3. Centre Manifolds 117

solution w(t) of (6.2.20) remains in a bounded set then

w(t) exists for all t > 0.

6.3. Centre Manifolds

Let Z be a Banach space with norm We consider


ordinary differential equations of the form

w = Cw + N(w), w(0) E Z, (6.3.1)

where C is the generator of a strongly continuous linear

semigroup S(t) and N: Z - Z has a uniformly continuous

second derivative with N(0) = 0, N'(0) = 0 [N' is the

Frechet derivative of N].

We recall from the previous section that there is a

unique weak solution of (6.3.1) defined on some maximal inter-

val [0,T) and that if T < W then (6.2.22) holds.

As in the finite dimensional case we make some spec-

tral assumptions about C. We assume from now on that:

(i) Z = X ® Y where X is finite dimensional and Y

is closed.

(ii) X is C-invariant and that if A is the restric-

tion of C to X, then the real parts of the

eigenvalues of A are all zero.

(iii) If U(t) is the restriction of S(t) to Y, then

Y is U(t)-invariant and for some positive con-

stants a,b,

ae-bt,
IIU(t)1i< t > 0. (6.3.2)

Let P be the projection on X along Y. Let

B - (I-P)C and for x E X, y E Y, let


118 6. INFINITE DIMENSIONAL PROBLEMS

f(x,y) = PN(x+y), g(x,Y) ' (I-P)N(x+Y). (6.3.3)

Equation (6.3.1) can be written

z Ax + f(x,y)
(6.3.4)
Y By + g(x,y).

An invariant manifold for (6.3.4) which is tangent to


X space at the origin is called a centre manifold.

Theorem 8. There exists a centre manifold for (6.3.4),

y = h(x), lxi < 6, where h is C2.

The proof of Theorem 8 is exactly the same as the

proof given in Chapter 2 for the corresponding finite dimen-


sional problem.

The equation on the centre manifold is given by

u - Au + f(u,h(u)). (6.3.5)

In general if y(O) is not in the domain of B then y(t)

will not be differentiable. However, on the centre manifold

y(t) - h(x(t)), and since X is finite dimensional x(t),

and consequently y(t), are differentiable.

Theorem 9. (a) Suppose that the zero solution of (6.3.5) is

stable (asymptotically stable) (unstable). Then the zero

solution of (6.3.4) is stable (asymptotically stable)

(unstable).

(b) Suppose that the zero solution of (6.3.5) is


stable. Let (x(t),y(t)) be a solution of (6.3.4) with

II(x(0),y(0))II sufficiently small. Then there exists a solu-

tion u(t) of (6.3.5) such that as t -


6.3. Centre Manifolds 119

x(t) - u(t) + O(e-Yt)


(6.3.6)
Y(t) - h(u(t)) + O(e_Yt)

where Y > 0.

The proof of the above theorem is exactly the same as


the proof given for the corresponding finite dimensional
result.

Using the invariance of h and proceeding formally

we have that

h'(x)[Ax + f(x,h(x))] - Bh(x) + g(x,h(x)). (6.3.7)

To prove that equation (6.3.7) holds we must show that h(x)

is in the domain of B.

Let x0 E X be small. To prove that h(x0) is in

the domain of B it is sufficient to prove that


U(t)h(x0) - h(x0)
lim
ti0+ t

exists. Let x(t), y(t) = h(x(t)) be the solution of


(6.3.4) with x(0) = x0. As we remarked earlier, y(t) is

differentiable. From (6.3.4)

r0 t

y(t) = U(t)h(x0) + U(t-T)g(x(T),Y(T))dT,


1

so it is sufficient to prove that

t
lim+ i U(t-T)g(x(T),y(T))dT
0

exists. This easily follows from the fact that U(t) is a

strongly continuous semigroup and g is smooth. Hence

h(x0) is in the domain of B.


120 6. INFINITE DIMENSIONAL PROBLEMS

Theorem 10. Let 0 be a C1 map from a neighborhood of the

origin in X into Y such that 0(0) - 0, 0'(0) = 0 and

O(x) E D(B). Suppose that as x _ 0, (Mo)(x) = O(jxjq),

q > 1, where

(MO) (x) _ 0' (x) [Ax + f (x, 0) 1 - BO (x) - 9 (x, 0 (x) )

Then as x - 0, IIh(x) - O(x) 11


= O(Ixlq).
The proof of Theorem 10 is the same as that given for

the finite dimensional case except that the extension

6: X -r Y of 0 must be defined so that 8(x) is in the

domain of B.

6.4. Examples

Example 8. Consider the semilinear wave equation

vtt + vt - vxx - v + f(v) ` 0, (x,t) E (0,Tr) x (0,°°)


(6.4.1)
v= 0 at x - O,Tr

where f is a C3 function satisfying f(v) - v3 + 0(v4)

as v -r 0. We first formulate (6.4.1) as an equation on a

Hilbert space.

Let Q - (d/dx) 2 + I, D(Q) - H2(0,Tr) f1 H1(O,Tr). Then


Q is a self-adjoint operator. Let Z - H1(0,Tr) x L2(0,Tr),

then (6.4.1) can be rewritten as

w = Cw + N(w) (6.4.2)

where

w2
Cw N(w)
Qwl-w2 - f(wl)

Since C is the sum of a skew-selfadjoint operator and a


6.4. Examples 121

bounded operator, C generates a strongly continuous group.

Clearly N is a C3 map from Z into Z.

The eigenvalues of are an = [-1 ±


(5-4n2)1/2]/2.

ai = 0 and all the other eigenvalues have real part less

than 0. The eigenspace corresponding to the zero eigen-

value is spanned by ql where

1
ql(x) I Js in x.
0

To apply the theory of Section 3, we must put (6.4.2)


into canonical form. We first note that Cq2 - -q2 where

1
q2(x) ]sin x

and that all the other eigenspaces are spanned by elements of

the form an sin nx, n > 2, an EIR2. In particular, all


other eigenvectors are orthogonal to ql and q2. Let

X - span(gl), V - span(gl,g2), Y span(q 2) ® V1, then

Z - X ® Y. The projection P: Z X is given by

wl
P = + w2)gl (6.4.3)
1
w2
L
where
n
2
w. = w. (6) sin e do.
1
0

Let w = sql + y, s EIR, y E Y and B = (I-P)C. Then we can

write (6.4.2) in the form

sql = PN(sgl+y)
(6.4.4)
y - By + (I-P)N(sgl+y)
122 6. INFINITE DIMENSIONAL PROBLEMS

By Theorem 8, (6.4.4) has a centre manifold y - h(s),

h(O) = 0, h'(0) - 0, h: (-6,6) - Y. By Theorem 9, the equa-

tion which determines the asymptotic behavior of solutions of


(6.4.4) is the one-dimensional equation

sql - PN(sgl + h(s)). (6.4.5)

Since the nonlinearities in (6.3.4) are cubic, h(s) - 0(s3),

so that
rn
s = n f(s+0(s3))sin 46 d9
1
0

or

s = -33 s3 + 0(s4). (6.4.6)

Hence, by Theorem 9, the zero solution of (6.4.4) is asymptoti-

cally stable. Using the same calculations as in Section 1 of

Chapter 3, if s(O) > 0 then as t - -,

s(t) = ( t) 1/2 + o(t-1/2). (6.4.7)

Hence, if v(x,t) is a solution of (6.4.1) with v(x,O),

vt(x,O) small, then either v(x,t) tends to zero exponen-

tially fast or

v(x,t) - ±s(t)sin x + 0(s3) (6.4.8)

where s(t) is given by (6.4.7).

Further terms in the above asymptotic expansion can be

calculated if we have more information about f. Suppose that

f(v) = v3 + av5 + 0(v7) as v - 0. In order to calculate an

approximation to h(s) set

(MO)(s) ` O'(s)PN(sgl+0(s)) - BO(s) - (I-P)N(sgl+0(s))


(6.4.9)
6.4. Examples 123

where 0:IR + Y. To apply Theorem 10 we choose O(s) so that

(Mo)(s) = 0(s5). If O(s) = 0(s 3) then

MO(s) = -BO(s) - (I-P)N(sgl) + 0(s5)


(6.4.10)

-BO(s) - s3g2 - 0 ]q + o(s5)


1 1

where q(x) = sin 3x. If

s
(s) - ag2s3 + 1
]qs3 (6.4.11)
s2

then substituting (6.4.11) into (6.4.10) we obtain

(MO) (s) = ag2s3 +


- S3 0
801+02
2
qs3 - s3g2 -
1
q + 0(s5)

Hence, if a - 3/4, 01 - 1/32, 02 0, then MO(s) - 0(s5),

so by Theorem 10
1
3 ]qs3 + 0(s5).
h(s) = g2s3 + 3 (6.4.12)
0

Substituting (6.4.12) into (6.4.5) we obtain

-3s3 _ 213 + 5a s5 + 0 s7
s = 2T y a ( )
4 1

The asymptotic behavior of solutions can now be found using


the calculations given in Section 1 of Chapter 3.

Example 9. In this example we apply our theory to the equa-

tion
1

vtt + vt + vxxxx - [B + (2/r4)J (vs(s,t))2dslvxx = 0,


0
(6.4.13)

with v - vxx - 0 at x - 0,1 and given initial conditions


v(x,0), vt(x,0), 0 < x < 1. Equation (6.4.13) is a model for
124 6. INFINITE DIMENSIONAL PROBLEMS

the transverse motion of an elastic rod with hinged ends, v

being the transverse deflection and B a constant. The

above equation has been studied by Ball [7,8] and in particu-

lar he showed that when a = -Tr2 the zero solution of

(6.4.13) is asymptotically stable. However, in this case the

linearized equations have a zero eigenvalue and so the rate

of decay of solutions depends on the non-linear terms. In

[13] the rate of decay of solutions of (6.4.13) was found

using centre manifold theory. Here we discuss the behavior

of small solutions of (6.4.13) when B + Tr2 is small.

As in the previous example we formulate (6.4.13) as an

ordinary differential equation. We write ' for space deri-

vatives and for time derivatives. Let

Z = H2(0,1) f1 H1(0,1) x L2(0,1), Qv = -V,,,, + By",

wl w2
w= C w= N (w) _
w2 Qww2 [f:1] )

2w 1

Tr 0

Then we can write (6.4.13) as

w = Cw + N(w). (6.4.14)

It is easy to check that C generates a strongly continuous

group on Z and that N is a C" map from Z into Z.

If A is an eigenvalue of C then we must have a non-

trivial solution u(x) of

u,,,, - Bu" + (),+A2)u = 0

u(0) = u"(0) = u(1) = u"(1) = 0.


6.4. Examples 125

An easy computation then shows that A is an eigenvalue of

C if and only if

2A = -1 ± [1 - 4 (n4Tr4+Bn2Tr2) ] 1/2.

Let E = r20 + Tr4. Then the eigenvalues of C are an(E),


where 2A1(E) = -1 + [1-4E]1/2, A2(E) = -1 - A1(E), and all

the rest of the an(E) have real parts less than zero for
E sufficiently small. The eigenspaces corresponding to

al(E) and A2(E) are spanned by ql and q2 where

1 1
ql(x) sin Trx, q2(x) _ sin Trx
1
(E) 2
(E)

while the eigenspaces corresponding to An(E) for n > 2

are spanned by elements orthogonal to ql and q2.

Let X = span(gl), V = span(gl,g2), Y = span(g2) 0 V


then Z = X 0 Y and the projection P: Z - X is given by

wl
PI = (XI(E) - )2(E)) 1(w2-a2(E)WI)gl
W2
I

where
1
wj = 2 j0 wj (9)sin re d9.

Let w = sql + y, where s EIR and y E Y. Then we

can write (6.4.14) in the form

Sgl = A1(E)sg1 + PN(sgl+y)


= By + (I-P)N(sgl+y) (6.4.15)

= 0.

By Theorem 8, (6.4.15) has a centre manifold y = h(s,E),

IsI < 6, IEI < CO. Using h(s,E) = O(s2+JESI), if N2(w)

is the second component of N(w) then


126 6. INFINITE DIMENSIONAL PROBLEMS

1
N2(sgl+h(s,e))(x) _ 2 [JO s2Tr4cos2Trede]Tr2s sin Trx

+ O(s4 + ics31)

= -s3sin Trx + O(s4 + Ies3I).

Hence PN(sgl+h(s,e)) _ (-s3+O(s4+1Cs31))gl, so that by

Theorem 9, the asymptotic behavior of small solutions of

(6.4.15) is determined by the equation

s - al(e)s - s3 + O(les3I + Is41). (6.4.16)

We can now determine the asymptotic behavior of small solu-

tions of (6.4.14). For 0 < e < 6, solutions of (6.4.14) are

asymptotically stable. For -6 < c<0, the unstable manifold

of the origin consists of two stable orbits connecting two

fixed points to the origin. (See Figure 1.)

Exercise 17. For c = 0, show that equation (6.4.16) can be

written as

s = -s3 - 3s5 + O(s7)

Phase Portrait for Small Negative c

Figure 1
6.4. Examples 127

Example 10. Consider the equation

utt + tut - uxx + a2v + f(u,v) = 0


(6.4.17)
vtt + 2vt - vxx - u + g(u,v) - 0

for (x,t) E (0,Tr) x (O,Tr) with u - v - 0 at x - 0,7r,


where f(u,v), g(u,v) have a second order zero at u = v - 0.

For a = 2, we show that the linearized problem has two purely


imaginary eigenvalues while all the rest have negative real

parts. We then use centre manifold theory to reduce the prob-

lem of bifurcation of periodic solutions to a two-dimensional


problem.

Let w - (u,v,u,S)T, then we can write (6.4.17) as

w = Cw + N(w) (6.4.18)

on Z - (H0(O,Tr))2 x (L2(0,n))2. Let

uxx + av
2

-v - u
xx

If u is an eigenvalue of A then A is an eigenvalue of

C where a2 + 2A + u - 0 and all the eigenvalues of C

arise in this way. An easy calculation shows that the eigen-


values of C are given by

A -1 ± (1-n2tia)1"2, n - 1,2,.... .

For a - 2, the eigenvalues of C are al - i, al -i, while

all the rest have negative real parts. Also

(Re A1(2)) > 0

so that al(a) and a1(a) cross the imaginary axis with non
zero speed. It is now trivial to apply centre manifold theory
128 6. INFINITE DIMENSIONAL PROBLEMS

to conclude that for a - 2 small, the behavior of small

solutions is determined by an equation of the form

r a 1l
s Is + J(s,a) (6.4.19)
L 1 a

where s E ]R2, a is a real parameter and J(s,a) - O(s2).

To apply the theory in Section 2, Chapter 3, we need to calcu-

late the quadratic and cubic terms in J(s,O). To do this we

need to put (6.4.18) into canonical form and to calculate the

centre manifold when a - 2. From now on we let a - 2.

On the subspace {r sin nx: r E]R4} the operator C

can be represented by the matrix Cn where

0 0 1 0

0 0 0 1
C
n
-n2 -4 -2 0

L 1 -n2 0 -2 J

Note that the eigenvalues of C are given by the eigenvalues

of Cn for n = 1,2,... . To put (6.4.18) into canonical

form we first find a basis which puts C1 into canonical

form. Calculations show that if


-2 0 0

ql 0 q2 1 q3 -1
0 -2 -2
-1 0 2

-2
0
q4
4

L 1J
then Clgl - -q2, Clg2 ql, Clg3 = -2q3 + q4, Clg4 - -q3-2q4.

Let w - Qz where Q - [g1,g2,q3,q4}, then we can rewrite


6.4. Examples 129

(6.4.18) as

i = Q-1CQz + Q-1N(Qz)
(6.4.20)

Let X= s1,s2 E IR}, V =


rl,r2 E]R}, Y = V 0 [X ® V]1 where ,y(x) = sin x. Then
Z = X 0 Y and the projection P on X along Y is given by

wl

w2
Pw =
0

0 J

2
fn
w. = wj(e)sin a de.
1
0

Let z - [s1,s200,0]T + y, si EIR, y E Y, then we can

write (6.4.20) in the form

0 1
Is + <PQ-1N(Qz),4>
-1 0
(6.4.21)

y = By + (I-P)Q 1N(Q z)

where B - (I-P)Q 1CQ, 11,1,0,0]T,y and s = [s1,s21T.

By Theorem 8, (6.4.21) has a centre manifold y h(s).

From w = Qz we have that w1 = -2z1 - 2z4 and

w2 - z2 - z3. Let Q-1 = (tij),

F(z) f(-2z1 - 2z4, z2 - z3)


t13 t14
g(-2z1 - 2z4, z2 z3)
G(z) J L t23 t24 -

Suppose that F(z) a F3(z) + O(IzI 4) and G(z) = G3(z) +


O(1zI4) where F3 and G3 are homogeneous cubics. Then if

J1(s), J2(s) denote the first two components of


130 6. INFINITE DIMENSIONAL PROBLEMS

on the centre manifold,

r n
J1(s) - n J F3(slsin 6, s2sin 6,0,0)sin 6 d6 + O(Isl 4)
0

with a similar expression for J2(s). Hence, on the centre

manifold

0 1 s1 + [Jl(s)
(6.4.22)
1 0 s2 J2(s)

and we can apply the theory given in Section 2 of Chapter 3.

If the constant K associated with (6.4.22) is zero (see

Section 2 of Chapter 3 for the definition of K) then the

above procedure gives no information and we have to calculate

higher order terms.

If F(z) = F2(z) + F3(z) + O(1z14) and G(z) - G2(z) +

G3(z) + O(Izl 4) where F2 and G2 are homogeneous quad-


ratics, then the calculation of the nonlinear terms is much

more complicated. On the centre manifold, z1 - s1iy + O(s2),

z2 = s2 + O(s2), z3 - O(s2), and z4 = O(s2). The terms of

order s2 make a contribution to the cubic terms in

PQ-1N(Qz). Hence, we need to find a quadratic approximation

to h(s). This is straightforward but rather complicated so

we omit the details.

Example 11. Consider the equations

BA-1u2
ut - Duxx + (B-l)u + A2v + 2Auv + u2v +

BA-1u2
vt = 6Dvxx - Bu - A2v - 2Auv - u2v - (6.4.23)

u = v - 0 at x - 0,1,
where A,B,6,D are positive. The above equations come from
a simplified model of a chemical reaction with u + A and
6.4. Examples 131

v + BA-1 as the chemical concentrations [5,12].

We study (6.4.23) on Z = (H0(011))2. Set

2
d
D + (B-1) A2
dx
W = , C-
C] B 9D
d
2
- A2
dx

N(w) - 1 (2Auv + u2v + BA lug)


C

then we can write (6.4.23) as

w - Cw + N(w). (6.4.24)

We analyze the situation in which for some value of the

parameters A,B,9,D, C has two zero eigenvalues such that

the restriction of C to the zero eigenspace has a Jordan

block. The bifurcation of static solutions where C has two

zero eigenvalues and the restriction of C to the zero eigen-

space is zero, has been studied in [25,26].

On the subspace {r sin nirx, r E R2} the operator C

can be represented by the matrix

rn2n2D + B - 1 A2
C =
n
-B -9Dn2n2 - A2

The eigenvalues of C are given by the eigenvalues of Cn

for n = 1,2,... . We suppose that two of the eigenvalues of

C are zero while all the rest have negative real parts.
For simplicity we assume that the eigenvalues of C1 are zero.

If C1 is to have two zero eigenvalues then


132 6. INFINITE DIMENSIONAL PROBLEMS

trace(C1) - B - 1 - rr2D - A2 - Oir2D 0


(6.4.25)
det(C1) - A2B - (B-1-ir2D) (A2+Oir2D) = 0.

We make the following hypotheses:

(Hl) There exists AO,BO,9O,D0 such that (6.4.25) is satis-

fied and the real parts of the rest of the eigenvalues

of C are negative.

(H2) For A,B,0,D in a neighborhood of AO,BO,9O,DO, we can

parametrize trace(C1) and det(C1) by

trace(C1) = E2, det(C1) _ -E1. (6.4.26)

The first hypothesis is satisfied, for example, if


D0 > 0 and

00 = 1, A0 ,r2 DO, B0 - (1 + rr2D0)2. (6.4.27)

If we vary B and 0 and keep A - A0, D = DO, then the map-

ping (B,9) + (det(C1), trace(C1)) has a non-zero Jacobian

at B = BOB 0 = 00 if AO,BO,00 are given by (6.4.27), so

by the implicit function theorem, (H2) is satisfied. In

order to simplify calculations we assume (6.4.27) from now on.

Let X = {siy: s i 2 , Y = X1 where iy(x) = sin rrx,


then Z = X ® Y. By Theorem 8, the system

if Cw + N (w)
e= 0

has a centre manifold h: (neighborhood of X x]R2) Y,

where we have written c _ (E1,E2). On the centre manifold,

the equation reduces to

s = C1(E)s + l(siy + h(s,E)) (6.4.28)


6.4. Examples 133

where s = (sl,s21T and

r0 1

Ni(z) = 2 Ni(z(6))sin n9 d9.


1

We treat the linear and nonlinear parts of (6.4.28) separately.


If ql = [1,P]T, q2 = [-p,l]T where p =

then C1(0)g1 = 0 and C2(0)g2 = (A2+B)ql. Let Q = (g1,g2]'


then

0 A2+B
Q 1C1(0)Q =
0 0

Let Q-1C1(e)Q = T = (tip) and let

1 0 1 0
M(e) = = 2 + 0(e)
0
A +B
t11 t12

Then for c sufficiently small, M(c) is nonsingular and

0 1
M(e)Q-1C1(e)QM-1(E)

_
[-detT) trace(T)

0 1 0 1

det(C1) trace(C1) el e2

by (6.4.26). Let s = QM-1(e)r, then (6.4.28) becomes

r 0
I
1 _
r + M(e)Q 1N(QM 1(e)r,
r + h(s,c)). (6.4.29)
L el e2

To check the hypotheses of Section 9 of Chapter 4 we


only need calculate the nonlinear terms when c = 0. Using

the fact that h(s,0) = 0(s2), routine calculations show that

M(0)Q 1N(QM-1(0)r
+ h(s,O)) - [P1,P2ITR(rl,r2)

+ 0(jr1I + I r2I3)
L34 6. INFINITE DIMENSIONAL PROBLEMS

where
j(l+p2)-1(l+p)(A2+B)
p1 - 3n(l+p2)-1(1-P), p2

R(r1,r2) a a1r1 + a2r1r2 + a3r2

BA-1 a2 - 2(A2+B)-1(B3/2A-2+A-BA-1).
al -
2B1/2,

Using (6.4.27), al = (r2D0) 1(1-n4D2) so that a1 is non-

zero if n2D0 + 1. We assume that n2D0 } 1 from now on.


1,
Note also that since 1 + p =-(n2D0) we have that p2

is non-zero.

To reduce (6.4.29) to the form given in Section 9,


Chapter 4,we make the substitution

p - (I-p1p21A(e))r. (6.4.30)

Substituting (6.4.30) into (6.4.29) and using the above cal-

culations we obtain

0 1
p + F(p,c) (6.4.31)
Lel e2

a2F2(0,0) a2F2(0,0)
a = ate- a B -
ap- p - = 2alp1 + a2p2
p
1

-=0
a2F1(0,0)

apl

We have already checked that a + 0. For most values

of DO,B is nonzero [B is only zero when D0 is a solu-

tion of a certain algebraic equation]. If a + 0 then we

can apply the theory given in Section 9, Chapter 4 to obtain

the bifurcation set for (6.4.23). If B - 0 then the theory

given in Section 9, Chapter 4 still gives us part of the


6.4. Examples 135

bifurcation set; the full bifurcation set would depend on

higher order terms.

Remark. If we vary 9 in (6.4.23) the theory given in Sec-

tion 3 does not apply since the map (9,v) - is not


evxx
even defined on the whole space. However, it is easy to
modify the results of Section 3 to accommodate the above sit-
uation. (See, for example, Exercises 1-2 in Section 3.4 of
[34].)
REFERENCES

1] R. Abraham and J. Marsden, Foundations of Mechanics,


2nd ed., Benjamin, New York (1978).

:2] R. A. Adams, Sobolev Spaces, Academic Press, New York


(1975).

[3] V. I. Arnold, Bifurcations in versal families, Russian


Math. Surveys 27, 54-123 (1972).

[4] V. I. Arnold, Loss of stability of self-oscillations


close to resonance and versal deformations of equivari-
ant vector fields, Functional Analysis and its Appl. 11,
No. 2, 1-10 (1977).

[5] J. F. G. Auchmuty and G. Nicolis, Bifurcation analysis


of nonlinear reaction-diffusion equations, Bull. Math.
Biology 37, 323-365 (1975).

[6] A. V. Balakrishnan, A lied Functional Analysis, Applica-


tions of Mathematics, Vol. pringer er ag, Berlin
and New York (1976).

[7] J. M. Ball, Stability theory for an extensible beam,


J. Diff. Egns. 14, 399-418 (1973).

[8] J. M. Ball, Saddle point analysis for an ordinary dif-


ferential equation in a Banach space and an application
to dynamic buckling of a beam, Nonlinear Elasticity,
J. Dickey (ed.), Academic Press, New York
[9] J. M. Ball, Strongly continuous semigroups, weak solu-
tions, and the variation of constants formula, Proc.
Amer. Math. Soc. 63, 370-373 (1977).

[10] J. M. Ball and J. Carr, Decay to zero in critical cases


of second order ordinary differential equations of
Duffing type, Arch. Rat. Mech. Anal. 63, 47-57 (1976).

136
References 137

[11] M. Berger, Nonlinearity and Functional Analysis,


Academic Press, New or
[12] J. Boa and D. S. Cohen, Bifurcation of localized dis-
turbances in a model biochemical reaction, SIAM J. Appl.
Math. 30, 123-135 (1976).

[13] J. Carr and Naji-Al-Amood, Rate of decay in critical


cases, J. Math. Anal. and Appl. 75, 242-250; II, Ibid.,
293-300

[14] J. Carr and R. G. Muncaster, The application of centre


manifolds to amplitude expansions, submitted to SIAM J.
Appl. Math.

[15] J. Carr and M. Z. M. Malhardeen, Beck's Problem, SIAM


J. Appl. Math. 37, 261-262 (1979).

[16] J. Carr and M. Z. M. Malhardeen, Stability of noncon-


servative linear systems, S rin er Lecture Notes in
Math., Vol. 799, Springer er ag, Berlin and New York
MO).
[17] N. Chafee, The bifurcation of one or more closed orbits
from an equilibrium point of an autonomous differential
system, J. Diff. Egns. 4, 661-679 (1968).
[18] N. Chafee, A bifurcation problem for a functional dif-
ferential equation of finitely retarded type, J. Math.
Anal. Appl. 35, 312-348 (1971).
[19] S. N. Chow and J. Mallet-Paret, Integral averaging and
bifurcation, J. Diff. Egns. 26, 112-159 (1977).
[20] S. N. Chow and J. Mallet-Paret, Fuller index and Hopf
bifurcation, J. Diff. Egns. 29, 66-85 (1978).
[21] S. N. Chow, J. Mallet-Paret, and J. K. Hale, Applications
of generic bifurcation, I, Arch. Rat. Mech. Anal. 59,
159-188 (1975); II, Ibid. 62, (1976).
[22] E. A. Coddington and N. Levinson, Theory of Ordinar
Differential Equations, McGraw-Hill, New York (1955).
[23] N. Fenichel, Persistence and smoothness of invariant
manifolds for flows, IndianaUniv. Math. J. 21, 193-226
(1971).
[24] N. Fenichel, Geometric singular perturbation theory for
ordinary differential equations, J. Diff. Egns. 31,
53-98 (1979).

[25] M. Golubitsky and D. Schaeffer, A theory for imperfect


bifurcation via singularity theory, Commun. Pure Appl.
Math. 32, 21-98 (1979).

[26] M. Golubitsky and D. Schaeffer, Bifurcation analysis


near a double eigenvalue of a model chemical reaction,
MRC Rep. 1859, Univ. of Wisconsin (1978).
58 REFERENCES

27] J. K. Hale, Ordinary Differential Equations, Wiley-


Interscience, New York 1

28] J. K. Hale, Critical cases for neutral functional dif-


ferential equations, J. Diff. Egns. 10, 59-82 (1970).

29] J. K. Hale, Generic bifurcation with applications,


Nonlinear Analysis and Mechanics: Heriot-Watt Symposium
Ced. R. J. Knops) Vol. 1, Pitman, London (1977).

30] P. Hartman, Ordinary Differential Equations, Second


Edition, P. artman

31] P. Hassard and Y. H. Wan, Bifurcation formulae derived


from center manifold theory, J. Math. Anal. Appl. 63(1),
297-312 (1978).

32] A. R. Hausrath, Stability in the critical case of pure


imaginary roots for neutral functional differential
equations, J. Diff. Egns. 13, 329-357 (1973).

33] F. G. Heinenken, H. M. Tsuchiya and R. Aris, On the


mathematical status of the pseudo-steady state hypothe-
sis of biochemical kinetics, Math. Biosci. 1, 95-113
(1967).

:34] D. Henry, Geometric Theor of Parabolic Equations, to


be published by pringer er ag.

:35] M. Hirsch, C. Pugh and M. Schub, Invariant Manifolds,


Springer Lecture Notes in Math., o. 5B3, Springer-
Verlag, Berlin and New York (1977).

[36] P. J. Holmes, Bifurcations to divergence and flutter in


flow induced oscillations - a finite dimensional analy-
sis, J. Sound and Vib. 53, 471-503 (1977).

[37] P. J. Holmes, Coupled flutter and divergence in auto-


nomous nonlinear systems, to appear in A.S.M.E.
Symposium on Nonlinear Dynamics.

[38] P. J. Holmes and J. E. Marsden, Bifurcations to diver-


gence and flutter in flow induced oscillations - an
infinite dimensional analysis, Automatica 14(4), 367-
384 (1978).

[39] E. Hopf, Abzweigung einer periodischen Losung von einer


stationaren Losung eines Differential systems, Ber.
h. Phys., K1. Sachs Akad. Wiss. Leipzig 94, T2
Mat_

[40] G. Iooss, Bifurcation of Maps and A lications, North-


Holland Mathematics Studies, Vol. North-Holland
Publishing Company, Amsterdam, New York and Oxford
(1979).
References 139

[41) G. Iooss and D. D. Joseph, Bifurcation and stability


of nT-periodic solutions branching from T-periodic
solutions at points of resonance, Arch. Rat. Mech. Anal.
66, 2, 135-172 (1977).

[42) M. C. Irwin, On the stable manifold theorem, Bull.


London Math. Soc. 2, 196-198 (1970).
[43) T. Kato, Perturbation Theory for Linear 0 erators,
Springer er ag, Berlin and New or

[44) A. Kelley, The stable, center-stable, center, center-


unstable, and unstable manifolds, J. Diff. Egns. 3,
546-570 (1967).

[45) A. Kelley, Stability of the center-stable manifold,


J. Math. Anal. Appl. 18, 336-344 (1967).

[46) W. Klingenberg, Lectures on Closed Geodesics, Springer-


Verlag, Berlin and New York
[47) N. Kopell and L. N. Howard, Bifurcations and trajector-
ies joining critical points, Adv. Math. 18, 306-358
(1976).

[48) 0. E. Langford, Bifurcation of periodic solutions into


invariant tori: the work of Ruelle and Takens,
S rin er Lectures Notes in Mathematics, Vol. 322,
pp. 159-19Z, pringer- er ag, Berlin and New York (1972).
[49) J. E. Marsden, Qualitative methods in bifurcation
theory, Bull. Amer. Math. Soc. 84, 1125-1147 (1978).
[50) J. E. Marsden and T. J. R. Hughes, Topics in the mathe-
matical foundations of elasticity, Nonlinear Anal sis
and Mechanics: Heriot-Watt Symposium (ed. Knops)
Vol. itman, London (1978).
[51) J. E. Marsden and M. F. McCracken, The Hopf Bifurcation
and Its A lications, Applied Math. sciences, Vol. 19,
pringer er ag, Berlin and New York (1976).

[52) S. J. Merrill, A model of the stimulation of B-cells by


replicating antigen, I., Math. Biosci. 41, 125-141
(1978) ; II. Ibid., 41, 143-155 (1978). -
[53) J. Moser, Proof of a generalized form of a fixed point
theorem, S rin er Lecture Notes in Math., Vol. 597,
Springer- er ag, Berlin and New York 977).

[54) A. I. Neishtadt, Bifurcation of the phase pattern of an


equation system arising in the problem of stability
loss of selfoscillations close to 1 : 4 resonance.
P.M.M. 42, 896-907 (1978).

[55] A. Pazy, Semigroups of linear operators and applica-


tions to partial differential equations, University of
Maryland Lecture Notes, Vol. 10, (1974).
40 REFERENCES

56] M. Reed and B. Simon, Methods of Modern Mathematical


Physics, Vol. II: Fourier Analysis, a -a ointness,
Academic Press, New or (1975).

57] D. Ruelle and F. Takens, On the nature of turbulence,


Comm. Math. Phys. 20, 167-192 (1971).

:58] R. J. Sacker, On invariant surfaces and bifurcation of


periodic solutions of ordinary differential equations,
New York Univ. IMM-NYU, 333 (1964).

[59] D. H. Sattinger, Topics in Stability and Bifurcation


The pry, Springer Lecture Notes in Math., Vol. ,

pier-Verlag, Berlin and New York (1973).


[60] J. T. Schwartz, Nonlinear Functional Analysis, Gordon
and Breach Science Publishers, New York (1969).

[61] I. Segal, Nonlinear semigroups, Ann. Math. 78, 339-364


(1963).
[62] Y. Shitzuta, On the classical solutions of the Boltz-
mann equation, to appear in Commun. Pure Appl. Math.

[63] S. J. van Strien, Center manifolds are not C", Math. Z.


166, 143-145 (1979).

[64] F. Takens, Forced Oscillations and Bifurcations,


Communication Math. Institute, ij suniversiteit,
Utrecht (1974).

[65] F. Takens, Singularities of vector fields, Publ. Inst.


Hautes Etudes Sci. 43, 47-100 (1974).

[66] D. A. Vaganov, N. G. Samoilenko and V. G. Abramov,


Periodic regimes of continuous stirred tank reactors,
Chem. Engrg. Sci. 33, 1133-1140 (1978).

[67] I. Vidav, Spectra of perturbed semigroups with applica-


tions to transport theory, J. Math. Anal. Appl., 30,
264-279 (1970).

[68] Y. H. Wan, Bifurcation into invariant tori at points of


resonance, Arch. Rat. Mech. Anal. 68, 343-357 (1978).

[69] J. Zabczyk, A note on Co semigroups, Bulletin de


L'Academie Polonaise des Sciences, Serie e-3 ssciences
mat aastr. et p ys., Z3, 895-898 (197S).
INDEX

Analytic semigroup, 105, 106, Contraction mapping, 14, 17, 19,


110 22, 24-26, 102-104, 108-109

Asymptotic behavior, 4, 8-12 Coupled equations, 1


19, 29, 32, 37-39, 46, 89,
114-115, 118, 122-123, 126 Dissipative operator, 108-109
128
Eigenvalues, 3, 7, 11-13, 16,
Asymptotic stability, 1, 2, 18, 20, 25, 34, 40, 44, 46-47,
4, 6-8, 21, 35-36, 118, 50, 53-55, 89-91, 98, 109,
122, 124, 126 111-112, 114-115, 117, 120,
124-125, 127-128, 131-132
Averaging, 42
Equilibrium point, 8, 13, 28,
Autonomous equations, 13 55, 57, 59, 65-67, 69, 73-75,
82, 84, 126, 131
Bifurcation (see also
Hopf bifurcation), 11-12, Evolution equation, 13, 97
40-41, 43, 47, 50, 52-53,
55-59, 82, 84-85, 94, 127, First integral, 65
131, 134-135
Generator, 97, 99-134
Cauchy problem, 101
Geodesics, 53
Center manifold
Gronwall's inequality, 18, 20,
approximation of, 2, 5-7, 23-24, 98
9, 13, 25, 32, 35, 38,
49, 120, 122-123 Group, 106-107, 113-114, 124

definition, 3 Hamiltonian system, 53


existence of, 1-4, 6-7, Hille-Yosida theorem, 102-104,
9, 12-17, 29-34, 38, 46, 107, 109
89, 93, 97, 117-118, 122,
124-125, 127-129, 132 Homoclinic orbit, 57, 67-68, 74-
75, 80, 82, 84
flow on, 2-6, 8, 10, 12,
19, 21-22, 29, 32, 35, Hopf bifurcation, 39-50
38, 46-47, 89, 92, 118-
119, 126, 128-129, 132 Implicit function theorem, 40-
41, 43, 48, 52, 56, 67-68,
properties of, 19-20, 28-30, 70, 73, 94, 132
95, 118
Infinitesimal generator (see
Chemical reactions, 8, 12, generator)
130-131
Instability, 2, 4, 6-7, 21, 35,
Closed graph theorem, 106 43-44, 115, 118

Compact operator, 111-112, Invariance of domain theorem, 24


114
Invariant manifold, 2-3, 13, 16,
Continuation, 116 30-34

141
.42 INDEX

Invariant set, 13, 33-34, Separatrix, 8


51-53, 117
Skew-selfadjoint operator, 107,
Jacobian, 3, 93-94, 132 113, 120

Laplace transform, 102 Spectral radius, 110

Liapunov function, 37 Spectrum, 109-113, 115, 117

Liapunov-Schmidt procedure, Stability, 2, 4, 6, 8, 19, 21-


12 22, 35, 43-44, 50, 57, 118

Linearization, 7, 11-13, 15, Stable manifold, 3, 8, 13, 15,


34, 37, 40, 44, 46-47, 50, 67
53-54, 57, 71, 81-82, 88,
92-94, 127 Strong solution, 104-105

Lumer-Phillips theorem, 108 Symplectic mapping, 53

Neutral functional differen- Uncoupled equations, 1


tial equations, 13
Uniform boundedness theorem,
Nilpotent, 20 100

Normal form, S2, 56, 89 Unstable manifold, 8, 13, 67,


82, 126
Parabolic equations, 105
Weak solution, 105, 116
Partial differential equa-
tions, 13, 98, 102

Periodic solutions, 28, 33,


40-45, 47, 50-51, 55, 57,
65-67, 69-71, 73-80, 82,
84, 127

Perturbation, 9, 12, 30, 33,


44-45, 113-114

Phase portraits, 11-13, 54,


60-63, 80, 82, 84-87, 126

Poincare map, 33

Rate of decay, 2, 4, 10, 37,


124

Reduction of dimension, 12,


19, 89

Resolvent, 102-103

Saddle point, 67

Selfadjoint operator, 104-


105, 107, 113, 120

Semigroup, 97-134

You might also like