An Introduction To Dynamical Systems by D. K. Arrowsmith, C. M. Place
An Introduction To Dynamical Systems by D. K. Arrowsmith, C. M. Place
An Introduction To Dynamical Systems by D. K. Arrowsmith, C. M. Place
V L ADIM IR I G O R E V I C H A R N O L D
and
S T E P H E N SMALE
for their inspirational work
UNIVERSIDAD DE CANTABRIA
BIBLIOTECA
00284076
An introduction to
D YNA M I C A L S Y S T E M S
D. K. ARROWSMITH
Lecturer, School of Mathematical Sciences,
Queen Mary & Westfield College, University of London
C. M. PLACE
Lecturer {formerly Department of Mathematics,
Westfield College, University of London)
№ t> tf _j
j
fit .
X: £ 5 X
/ : W X
Cambridge
P UNIVERSITY PRESS
Published by the Press Syndicate of the University of Cambridge
The Pitt Building, Trumpington Street, Cambridge CB2 1RP
40 West 20th Street, New York, NY 10011-4211, USA
10 Stamford Road, Oakleigh, Melbourne 3166, Australia
MP
CONTENTS
Preface
References 413
Index 417
PREFACE
In recent years there has been a marked increase of research interest in dynamical
systems and a number of excellent postgraduate texts have been published. This
book is specifically aimed at the interface between undergraduate and postgraduate
studies. It is intended both to stimulate the interest of final year undergraduates
and to provide a solid foundation for postgraduates who intend to embark on
research in the field. For example, a challenging third-year undergraduate course
can be constructed by selecting topics from the first four chapters. Indeed, lecture
courses taught by one of us (CMP) provided the basis for Chapters 1, 2 and 4.
On the other hand, Chapter 6 is directed at first-year postgraduate students. It
contains a selection of current research topics that illustrate the interaction between
superficially different research problems.
A major feature of the book is its extensive set of exercises; more than 300 in
all. These exercises not only illustrate the topics discussed in the text, but also
guide the reader in the completion of technical details omitted from the main
discussion. Detailed model solutions have been prepared and hints to their
construction are provided.
The reader is assumed to have attended courses in analysis and linear algebra
to second-year undergraduate standard. Prior knowledge of dynamical systems is
not necessary; however, some familiarity with the qualitative theory of differential
equations and Hamiltonian dynamics might be an advantage.
We would like to thank Martin Casdagli for sharpening our understanding of
Birkhoff attractors, David Knowles and Chris Norman for helpful discussions and
Carl Murray for steering some awkward diagrams to a laser printer. We are
grateful to the Quarterly Journal of Applied Mathematics and Springer-Verlag for
allowing us to use diagrams from some of their publications and our thanks go
to Sandra Place for her fast and accurate typing of much of the manuscript. One
of us (CMP) would like to acknowledge the Brayshay Foundation for its financial
support throughout this project. Finally, we must both pay tribute to the patience
and support of our families during the long, and often difficult, gestation period
of the manuscript.
1
Diffeomorphisms and flows
1.1 Introduction
A dynamical system is one whose state changes with time (f). Two main types of
dynamical system are encountered in applications: those for which the time variable
is discrete (ie Z or N) and those for which it is continuous (teR ).
Discrete dynamical systems can be presented as the iteration of a function, i.e.
xl+1=f(x,), ie Z o rN . (1.1.1)
^ = x = X(x). (1.1.2)
dt
In (1.1.1 and 2), x represents the state of the system and takes values in the state
or phase space. Sometimes the phase space is Euclidean space or a subset thereof,
but it can also be a non-Euclidean structure such as a circle, a sphere, a torus or
some other differentiable manifold.
In this chapter we will consider two special cases of the above equations, namely
when:
These two cases have been widely studied and they are fundamental to our
understanding of dynamical systems. Smale, in his definitive work (Smale, 1967),
pointed out that (i) and (ii) are closely related and our discussion emphasises this
connection.
Any description of the theory of (i) and (ii) involves differentiable maps so let
us begin by recalling some definitions. Let U be an open subset of R". Then a
function g: U -+ R is said to be of class O if it is r-fold continuously differentiable,
1 oo. Let V be an open subset of Rm and G: l/-* V. Given coordinates
2 1 Diffeomorphisms and flows
S
4 1 Diffeomorphisms and flows
V y f V
= (h /J.h-«).(ha. f . h - i ).(hoth ^ 1)
=lv W - 0-1-5)
Thus all local representatives of f have the same differentiability class, Ck say, with
k < r. It is important to note that r is determined entirely by the charts and hence
by the structure of Ai. A manifold with overlap maps of class C is called a
C-manifold.
The discussion presented above is, of course, incomplete. We have only
considered maps taking a chart into itself. This is clearly not true in general. Given
f: Ai - » Ai, then f: Wa -» and f: Wt n Wy -* Wf . The generalisation of our
simple arguments that allows for these omissions is considered in Exercise 1.1.2.
Needless to say, the ‘message’ is unchanged by these manipulations.
A more detailed discussion of differentiable manifolds is not necessary here (the
interested reader should consult Arnold (1973) or Chillingworth (1976)). While
the ideas outlined above provide valuable background knowledge, we will rarely
find ourselves involved with charts, atlases, etc. This is because our concern is the
dynamics of maps defined on Ai given that they are diffeomorphisms or flows.
W------- ------ -W
h 1 h
U ------- ;-------U
f
Figure 1.3 Illustration of the definition of the overlap maps h ^ and hfa.
Note that his, = h ^ '.
Wa nW p Wa n W f
These maps are usually presented to us in local coordinates so that the manifold
structure does not appear explicitly.
Definition 1.2.1 A point x*e M is called a fixed point o /f i/r"(x*) = x* for all m e Z.
Definition 1.2.2 A point x*eM is a periodic point of f if P(x*) = x*, for some
integer q > 1.
The least value of q satisfying Definition 1.2.2 is called the period of the point x*
and the orbit of x*, i.e.
(x*,f(x*)....... f « - V ) } , (1.2.1)
Definition 1.2.3 A fixed point, x*, is said to be stable if, for every neighbourhood
N of x*, there is a neighbourhood N' of x* such that if x e N ' then f"(x)e)V
for all m > 0.
Essentially, Definition 1.2.3 says that iterates of points ‘near to’ a stable fixed
point, remain ‘near to’ it for m e Z +. If a fixed point x* is stable and Lim P(x) = x*,
m-*co
for all x in some neighbourhood of x*, then the fixed point is said to be
asymptotically stable. Trajectories of points near to an asymptotically stable fixed
point move toward it as m increases. Fixed points that are stable, but not
6 1 Diffeomorphisms and flows
asymptotically stable, are said to be neutrally or marginally stable and those that
are not stable in the sense of Definition 1.2.3 are unstable.
K«(0) = (0 + a )m o d l. ( 1.2.2)
Here we have assumed that 0 is measured in units of 2n. If a = p/q, p,qe Z and
relatively prime, then
R2(0) = (0 + p )m o d l = 0 (1.2.3)
and we conclude (cf. Definition 1.2.2) that every point of the circle is a periodic
point of period-0, i.e. the orbit of any point is a 0-cycle (see Figure 1.4). If a is
irrational then
for any 0 and, in fact, the orbit of any point fills the circle densely (see Exercise 1.2.1).
Obviously more general diffeomorphisms of S' do not simply rotate all points
uniformly. Crudely speaking they compress some arcs of the circle and stretch
Figure 1.4 Typical orbit of the pure rotation R, for a = p/q — 2/5. Observe
that the orbit of 0 winds around the circle p = 2 times before returning
to 0 on the fifth iteration.
1.2 Elementary dynamics o f diffeomorphisms 7
others. It is then difficult to recognise fixed or periodic points from the representation
of orbits on the circle itself. This is a problem for any map ( / ) of the circle,
whether it is a diffeomorphism or not, and it is solved by considering a lift of /.
The natural setting for introducing the lift of f : S 1->Sl is when / is a
homeomorphism rather than a diffeomorphism and it would be perverse to
artificially confine our discussion to the differentiable case. Moreover, by taking
/ to be a homeomorphism at this point we can better appreciate the consequences of
imposing differentiability on f and f ~ l. Thus, let f: S' -* S‘ be a homeomorphism
and suppose there is a continuous function/: R -+ R such that
/ ( * + ! ) = /( * )+ 1 (1.2.7)
for every x e R.
Suppose fc > 2, then f(x) and ]{x 4 1) differ by more than two and / takes the
form shown schematically in Figure 1.6(a). Clearly, the points x 0 and x, satisfying
/(x 0) = 1 and /(Xj) = 2 are both less than unity. This means that n maps them to
distinct points on S1. However, /(x 0) and /( x t) differ by unity and therefore
represent the same point on S1. This contradicts the hypothesis that / is a
homeomorphism. Hence k < 1.
If fc = 0 ,/(0 ) =7(1) and 7 fails to be injective on (0,1) (see Figure 1.6(6)). Again
this contradicts the fact that / is a homeomorphism.
If fc < 0 then continuity of / can only be maintained if / is orientation-reversing
in contradiction to hypothesis. Moreover, it is clear that similar arguments would
lead to a minus sign in the right hand side of (1.2.7) for orientation-reversing / .
Finally, we conclude that k = 1 and (1.2.7) follows. □
ff( x ) = - x 2 + 2x + i ( 1 .2. 11 )
Figure 1.6 Schematic forms for f when (1.2.10) has (a) fc = 2; (6) fc = 0.
In both cases, the hypothesis that / is a homeomorphism is contradicted.
(a) (6)
1.2 Elementary dynamics o f diffeomorphisms 9
but not a diffeomorphism of S '. To obtain the latter, / must be a bijection and
differentiable for all * e R. An example of this type is shown in Figure 1.8(b) where
« [ 0 ,1 ] .
Notice, we have, without loss of generality, taken /(0) e [0,1) in both of the
above examples. Observe that, n(f(x) + k) = n(f(x)), for any ke Z. Thus if f(x) is
a lift of / then so is f k{x) = f(x) + k , k e l . Therefore, unless otherwise stated, we
will assume that / is the member of this family of lifts satisfying /(0)6 [0,1).
Figure 1.7 The function / shown here cannot be the lift of a homeo-
morphism / : S1 -* S1 because it is not injective.
or
7(x*) = x* + 1. (1.2.13b)
In either case,
f(x*) = x* + k, k e l. (1.2.17)
f(y*) + l = y* + l+ k. (1.2.18)
Thus 7 2 is a lift of f 2.. It only remains to ensure that 7 2( 0 )e [0 ,1) (i.e. choose
the lift 7 2- [ 7 2(0)], where [•] denotes the integer part of •), and Proposition
1.2.2 allows us to find the period-2 points of / . These arguments obviously extend
to points of period q> 2.
An alternative approach is to recognise that if 7 , (0)e[/, I + 1) then (1.2.13) is
1.3 Flows and differential equations 11
replaced by
This point of view often has the advantage that f , f 2, . can be presented
on the same diagram (see Figure 1.10) without ending up with a confusion of
curves in the vicinity of y = x and y = x + 1.
The lift / of f: S1 -> S1 not only provides a means of conveniently finding fixed
and periodic points, it can also allow us to determine their stability. If (1.2.13a)
is satisfied at x*, then the orbits of points near to x* under / can be obtained by
moving between y = /(x ) and y = x as in Figure 1.11. The fixed point x* is stable
(unstable) if
|D /(x * )|< l (>1), (1.2.21)
(see any first course in Numerical Analysis). The stability of 0* = jt(x*) is clearly
the same as that of x*. When (1.2.13b) is satisfied, we can either replace f by
/ — 1, so that x* is then represented by an intersection with y = x, and proceed
as above or construct paths for the orbits of / by using y = f(x) and y = x + 1.
The stability of the fixed point is still given by (1.2.21).
that:
for each i,je Z. It is said to be an action of the group Z on M or, more precisely,
the Z-action generated by f (see Chillingworth, 1976). In this section we consider
the action of the group R on M; such R-actions are called flows on Af.
(a) = (1.3.2a)
Observe that (1.3.2a and b) imply that (<p,) 1 exists and is given by <p_,. Since
ip e C \ it follows (see Exercise 1.3.1) that ip,.M -*M is a diffeomorphism for each
teR .
Let us pursue the analogy with diffeomorphisms a little further. We define the
orbit or trajectory of ip through x to be {<p,(x)|teR} oriented in the sense of
increasing t. It can be shown (see Exercise 1.3.2) that there is one and only one
trajectory of tp passing through each point xeA i. If <p,(x*) = x* for all te R then
x* is said to be a fixed point of the flow. Fixed points of flows can be stable,
asymptotically stable, neutrally stable or unstable in the sense of Liapunov. Precise
definitions are obtained by the transcription f" t-+ tp, and m eh -> te U in Definition
1.2.3 and the comments following it.
The orbit of a fixed point is just the point itself. If x is not a fixed point it is
said to be ordinary or regular. The trajectory through an ordinary point gives rise
to an oriented curve on M and tp has periodic points if this curve is closed.
Clearly, if q>t{x) = x the orbit returns to x after time t. If T is the least, positive
time for which this occurs, x is a periodic point with period T. It is easily shown
(see Exercise 1.3.3) that if a closed orbit has one point with period T, then every
point of y is periodic with period T. Thus, T is also called the period of y.
The set of all trajectories of a flow is called its phase portrait. Since each trajectory
corresponds geometrically to an oriented curve or point on Af, a valuable pictorial
representation of the flow is obtained by sketching or plotting typical trajectories.
Some examples are shown in Figure 1.12. Notice that the caption to this figure
does not specify tp„ instead a differential equation is given. How are flows related
(1.3.3)
Proposition 1.3.1 i»i(xo) is the solution of x = X(x) which passes through \ 0att = 0.
V fc { (i))-* (0 ,№ ))
£
= x « w ). (1.3.4)
(c)
1.3 Flows and differential equations 15
Thus, f(t) is a solution of x = X(x) and, since <p0 = idM, £(0) = tp0(x0) = x0, as
required. □
Notice that if X(x*) = 0 then <p,(x*) = x* is the solution of x = X(x) passing through
x*. Moreover, if q>,(x*) = x* for all t then (1.3.3) implies X(x*) = 0. We conclude,
therefore, that x* is a fixed point of tp, if and only if X(x*) = 0. Such points are
referred to as singular points of the vector field X.
Proposition 1.3.1 means that every flow on M corresponds to an autonomous
differential equation. Unfortunately the converse is not true. This is because there
are autonomous differential equations with solutions that cannot be extended
indefinitely in t. For example, x = x2 has general solution
= (1.3.6)
1 —x0t
provides a local flow for x = x2. When xo > 0 , (1.3.5) implies t e ( —o o ,x j‘) in
(1.3.6). It is easy to verify that <p, satisfies (1.3.2) provided t,s and t + s all belong
to ( —oo, Xq *)• The same function tp, can be used when x0 < 0 provided t is restricted
to the interval (x ^ 1, oo). Equation (1.3.6) obviously provides the trivial solution
when x 0 = 0. This local flow is sufficient to characterise the solutions of x —x 2 in
a neighbourhood of the origin of the t, x-plane. For example, for |x0| < e, (1.3.6)
certainly gives the solutions to x = x2 for f e ( —e^ l , e “ ‘). Flows of this type are
frequently used implicitly when local properties are discussed (e.g. the saddle-node
singularity in Example 2.7.4).
With the above proviso in mind, differential equations, vector fields and flows
merely provide alternative ways of presenting the same dynamics. These alternatives
have arisen for historical reasons; applications frequently lead to differential
equations; local analysis is usually presented in terms of vector fields; and global
analysis uses the language of flows. We hope the reader will become familiar with
all three possibilities.
Invariant sets are said to be positively (negatively) invariant if the orbits of their
elements remain within them for m e Z + (Z~) or < > 0 (t< 0 ).
Clearly, the orbit of any point is an example of an invariant set. It follows
therefore that fixed points, cycles and closed orbits are all invariant sets. However,
they are rather special in two main ways.
(i) They are minimal in the sense that they do not have any proper subsets
that are themselves invariant. For example, the circle '■€ is an invariant
set for both of the flows shown in Figure 1.14. In contrast to the flow
shown in (a), the circle # in (b) has proper subsets, Pt, P2, T, and r 2,
that are invariant under the flow.
(ii) They exhibit periodicity. This is particularly important for applications
where such sets frequently correspond to observable phenomena.
More subtle forms of recurrence than periodicity can occur in dynamical systems
and the following definitions allow us to describe them.
The set of non-wandering points for f (tp) is called the non-wandering set, 0(f)
(fi(ç>)). It is easy to see that fixed points and periodic orbits lie in Q (see Exercises
1.4.2 and 1.4.3), however, points exhibiting milder forms of recurrence are also
present. For example, consider an irrational rotation of the circle, S1. No point
of the circle is periodic, but the orbit of any point x ultimately approaches x
arbitrarily closely. Thus, every point of S' is a non-wandering point and fi = S1.
The structure of fi will be examined more closely in § 3.6, but we can recognise
some important subsets of it by formalising the idea that fixed points and closed
orbits frequently attract or repel the trajectories of phase points not contained in
them.
The set of all ( limit points of x is known as the limit set of x, denoted
(co- Ico-
f-Lcr(x)
by These sets are invariant under f (<p). Let 2 = f"(y), m e Z (z = <p,(y),
Im x )
te R), where y satisfies Definition 1.4.2. Then Lim f" '+M(x) = z (Lim <pll+,(x) = z)
i -* co i-»oo
so that z and y belong to the same limit set of x.
Notice that a- and co-limits sets are subsets of 0 for any x. Recall if y then
Figure 1.14 The circle ¥ is an invariant set for both of the flows shown.
However, in (a) ¥ has no proper subsets that are themselves invariant;
while in (b) % is the disjoint union of the invariant sets P ,, P2, r 1; T2.
18 I Diffeomorphisms and flows
there exists a neighbourhood Kay such that fm(V )n V is empty for all m > 0.
However, y eL m(\) implies P 'M e V for / > N, say, and hence there is z = fN(x) e V
such that P", ' N(z)e V for i > N. Thus P (K ) n V cannot be empty for all m and y
must lie in ft.
Example 1.4.1 Find LJx) and L J x ) for (a) x = 0; (b) x # 0, when <p is the flow
on R2 induced by
Solution, tp has a unique, attracting closed orbit y given by r(t) = 1, with period
T = 2k, and an unstable fixed point at the origin (see Figure 1.15).
(a) x = 0
Note <p,(0) = 0 for all t therefore
(b) x?£0
Let y = (cos 0O, sin 60)ey and let tf be the sequence of t > 0 at which the orbit of
x crosses the radial line from 0 through y. Then Lim <pu(x) = y and y is an co-limit
|x |< 1,
(1.4.5)
r
1.4 Invariant sets 19
However, for |x| > 1, the Lim <p,,(x) does not exist for any sequence t, such that
j~*oo
t; -> —oo as i -» oo and therefore L„(x) is empty. □
Example 1.4.2 Let the flow <phave the phase portrait shown in Figure 1.16. What
are Lra(x) and L Jx) for xeA , B ,C respectively? What feature do all three co-limit
sets have in common?
Let r A and TB be the trajectories of the flow which form the séparatrices of the
saddle point P0. Observe that
dA = r Au P 0, ôB = r „ u P 0, (1.4.6)
and it follows that all three co-limit sets are unions of fixed points and the trajectories
joining them. □
This theorem states that the types of limit sets illustrated in Examples 1.4.1 and
1.4.2 are the only compact ones that can occur in flows on the plane. It is one of
the few theorems which gives the existence of a global feature of a phase portrait.
Figure 1.16 Phase portrait of the flow required for Example 1.4.2. The
points P q , 2 are fixed points. The open sets A, B have boundaries dA,
dB, respectively. C is the complement of the closure of A u B .
20 l Diffeomorphisms and flows
Definition 1.4.3 A limit cycle is a closed orbit y such that either y E L Jx)o r y E L„(x)
for some x^y.
Theorem 1.4.1 has the important corollary that a non-empty, compact set A which
is positively or negatively invariant contains either a limit cycle or a fixed point.
This result can be useful in demonstrating the existence of limit cycles (Arrowsmith
& Place, 1982, pp. 147-51).
1.5 Conjugacy
We now turn to the equivalence relations which allow us to recognise when two
diffeomorphisms or two flows exhibit the ‘same’ behaviour. These equivalence
relations lie at the heart of topological or qualitative theory.
h f = g.h. (1.5.1)
Topological conjugacy of two flows <p„ÿ,: Ai -» Ai is defined in the same way with
(1.5.1) replaced by = for all feR .
Definition 1.5.1 means that h takes each orbit of f (or <pt) onto an orbit of g (ÿt)
preserving the parameter m (t), i.e.
f and <px is confined to this interval and has the same orientation for both maps
(N.B. sign(x„+1 - x.) = sign(/(x„) - x„) = sign(x), neZ).
Let x0, }'0e(xf, xf+1) and consider the orbit of x0 under / and the orbit of y0
under <pj. Let P„ = /" (x 0) and Q„ = <p,(y0), neZ . Observe (see Figure 1.19) that
and (1.5.4)
V i - [Qn> Q » + 1 ] - * [ Q n + i t 6 » + 2 ] >
hoiyo) = x0 + ( y - ro )( ^ t °- (1-5.5)
K(y)=f"ho<P-n(y)- (1-5-6)
Clearly, hn: [Q„, 8„ +,] -> [P„, Pn+,] and, what is more,
Figure 1.19 Orbits of the points x0 and y0 under / and <px, respectively.
It is convenient to define P ,= /"(x 0) and Q, = i>B(>o). for n e l.
1.5 Conjugacy 23
1x* y = xf
h„(y) for ye[Q„,Q„+ j], n= Z (1.5.8)
x?+1 Z= 1
is a homeomorphism.
Finally, it is easy to verify that h exhibits the conjugacy of / and cp,. If
xe [xf, x,*+ j] then x 6 [Qn, Qn+J for some n and
h-<Pi{x) = K +l-q>l(x)
= f " * l h0 '(P-n-i-<Pi(x)
=f K( x )
=f-h(x) (1.5.9)
as required. □
'y = x
24 1 Diffeomorphisms and flows
as required. Thus, when h exhibits the conjugacy of <p and i/t, the derivative map,
Dh, transforms the vector field X(x) into Y(y) with y = h(x).
An important example of C'-conjugacy of flows occurs in the qualitative study
of local phase portraits in the neighbourhood of an ordinary point. Let x0 be an
ordinary point of the flow tp: R x R" -> R" of the vector field X: R" -* R".
For convenience, we will assume that H has normal X(x0) in the following
discussion. Observe, (see Figure 1.21) that there is a neighbourhood, V, of x0 such
that any point x e V can be written as x = <p„(y), where yeS . In other words, we
can use the trajectories of the flow to define new coordinates on V.
These new coordinates are best related to local coordinates at x0, therefore, let
x h x - x0 so that x0 is at the origin of both sets of coordinates. Now suppose
we choose a basis in R" which has X(0) as its first vector. Then the first coordinate
of every point y e S is zero and S defines a neighbourhood, S, of the origin in R"~'
(see Figure 1.21(b)). Each point of S can be specified by £ e R" ~ 1 and every point
x of V can be written as
By definition of tp, h: R" -> R" is a C’-function. What is more, h|S is the identity
and D„h(0) = X(0), by (1.3.3). Thus Det Dh(0) 0 so that IT 1 exists and is C1
by the Inverse Function Theorem. In the new coordinates, the trajectories of the
flow are simply lines of constant $ (see Figure 1.21(c)), i.e.
W « . «) = (» + »,«)■ (1.5.14)
1.5 Conjugacy 25
S c R " -'
(«
X(0)
26 I Diffeomorphisms and flows
(1.5.15)
(1.5.16)
(1.5.17)
and <j/, is C'-conjugate to ipr The arguments presented above essentially constitute
a proof of the ‘Flow-box’ Theorem.
Theorem 1.5.1 (Flow-box) Let x0 be an ordinary point of the flow ip. Then in every
sufficiently small neighbourhood of x0, (p is Cl-conjugate to the flow ^(t, x) = x + ie,,
where e] is a unit vector parallel to the x ,-axis.
The above examples emphasise that in order to prove two flows or diffeomorphisms
conjugate, we must construct an appropriate map satisfying (1.5.1). It is often a
great deal easier to recognise when no such map exists. For example, consider
two flows: ip, with an isolated fixed point and i/r, with no fixed points at all. The
fixed point is a trajectory of ip, and, therefore, if ip, and ijr, are topologically
conjugate, there is a homeomorphism which takes a trajectory of ty, onto the fixed
point. However, every trajectory of contains more than one point and can only
have a single point image under a non-injective map. This contradiction proves
that ip, and tji, are not topologically conjugate. This result has an obvious extension:
a necessary condition for two flows to be C°-conjugate is that they have the same
number of fixed points. Here an easily recognisable property of the flows (namely,
the number of fixed points) allows us to conclude that they are not conjugate.
Another, perhaps less trivial example of this approach, is afforded by diffeo
morphisms on the circle. Let us begin by considering pure rotations.
A property that distinguishes rational and irrational rotations is their rotation
number. This quantity can be defined for any homeomorphism / : S1 -» S'.
(1.5.18)
where f is a lift of f.
As our notation suggests, it can be shown that p {f) is independent of the point
x occurring in (1.5.18). A proof of this fact can be found in Nitecki (1971, pp. 33-4).
1.5 Conjugacy 27
R y(x) = x + y (1.5.19)
Proposition 1.5.1 A diffeomorphism f \ S' -> S' has periodic points if and only if its
rotation number, p (f), is rational.
Proof. If f has a periodic point then, given a lift, / , of / , there exists x* e R such that
/"(**) = x* + p, (1.5.20)
for some integers p and q. It follows that f**(x*) - x* + ftp, and therefore
f ^ x* ) - x * _ n p _ p
(1.5.21)
nq nq q
Hence p (f) is rational.
To prove the converse, suppose / has no periodic points then,
f '( x ) * x + p, (1.5.22)
or
f'( x ) - x * p , (1.5.23)
for any integers p, q and any x e R. Since gq(x) = 7 , (x) - x satisfies gq(x + 1) = gq{x),
Rr
x --------------------------------- - x + a
Rr
0 = x mod I Rr (6) = (8+a) mod 1
= (x + a) mod 1
28 I Diffeomorphisms and flows
for each x, (1.5.23) means that there exists s> 0 such that either
or
For a proof of Denjoy’s Theorem the interested reader should consult Arnold
(1983, pp. 105-6) or Nitecki (1971, pp. 45-9). This important result means that
every orbit of / is dense in the circle provided f e C 2 and p ( f) is irrational. If
f$ C 2, then more complicated phenomena, such as invariant Cantor sets, can
occur (see §6.4.1 and Nitecki, 1971).
the orbits of the latter are parametrised by a continuous variable t. This allows us
some additional freedom in the mapping of orbits onto orbits.
Definition 1.6.1 Two flows, <p, and <//„are said to be topologically (or C°) equivalent
if there is a homeomorphism, h, taking orbits of (p, onto those of preserving their
orientation.
equations
r = r(l-r), 0 = 2, ( 1.6.2)
where (r, 0) are polar coordinates, have similar phase portraits. Both have an
attractive closed orbit y with r(t) = 1 and an unstable focus at the origin. However,
the closed orbit has p e r io d s in (1.6.1) and period-7t in (1.6.2). Thus, if h : y - ,y
preserves the parameter t, it must fail to be a bijection. Thus (1.6.1) and (1.6.2)
are not topologically conjugate, but they are topologically equivalent. Observe that
the time rescaling ti-*2t transforms (1.6.2) into (1.6.1).
If Definition 1.6.1 is satisfied with h e C \ 1, then the stronger relationship
between q>and <1?can be emphasised by saying that they are Ck-equivalent. If two
flows ip and <1/ are C‘-equivalent (!c> 0) then their vectors fields X(x) and Y(y)
are also said to be (^-equivalent. This terminology is frequently used because
flows are often described implicitly in terms of their vector fields. For example, in
applications one is often provided with a model differential equation but no explicit
form for its solutions.
When k > 1 there is a Ck-diffeomorphism, h, such that
(cf. (1.5.11)), where er:R"->R takes only positive values corresponding to the
reparametrisation of the time. Recall the vector field of is given by
= cr(y)Y(y) (1.6.4)
where cr(y) = i y(0) is a positive scale factor altering the magnitude but not the
direction of Y(y).
Example 1.6.1 Show that the vector fields Jx and J 0x, with
O
+
Since /?> 0 , the flows defined by (1.6.6) and (1.6.8) are topologically equivalent
with h equal to the identity. In other words, they have identical trajectories and
differ only in the speed at which they are described.
Elimination of t from (1.6.7) and (1.6.8) gives
flu
0 = ( 0 -0 „ ) + 0 „. (1.6.9)
Equation (1.6.9) defines a map taking the trajectory of (1.6.8) through (r0,0 O)
onto the trajectory of (1.6.7) through (R0, 0 O) (see Figure 1.25). For r, r0 > 0, this
map is 1:1, continuous and preserves orientation (indeed it preserves t itself);
Figure 1.25 Illustration of the effect of the map (1.6.9) on the orbit of
(1.6.8) through (r0, 60). The result is the orbit of (1.6.7) passing through
(*o.»o)-
32 1 Diffeomorphisms and flows
R = /■"«, 0 = 0, (1.6.10)
Example 1.6.2 Use the map r' —r, 0' = 0 —ln r (r > 0) to demonstrate that the
vector fields J 0x, where J 0 is given in (1.6.5), and x are topologically equivalent.
The map h: R2 -> R2 is continuous and has continuous inverse, r = r', 0 = O' + In r',
r' > 0. Since h(0) = 0, h takes the fixed point trajectory of the flow of J 0x onto
that of x. For x ^ 0, h is differentiable so we can check its effect on the flow by
transforming the differential equation x = J 0x or, in polar coordinates, r = r, 0 = 1.
We find
which is just the polar form of * = x. Of course, h is not differentiable at the origin
so that (1.6.11) is only a homeomorphism of the plane. Hence J 0x and x are
topologically equivalent. □
When two flows are topologically equivalent we say they are of the same topological
type. The results obtained in Examples 1.6.1 and 1.6.2 play an important role in
the classification, up to topological type, of all linear vector fields on R2 (see
Arrowsmith & Place, 1982, p. 58). The matrix J in (1.6.5) is the real Jordan form
of any 2 x 2 real matrix, A, with complex eigenvalues a ± i/I, a > 0, i.e. there is a
real non-singular matrix M such that M - ! AM = J. It follows (see Exercise 1.5.6)
1.7 Poincaré maps and suspensions 33
that the flows of Ax and Jx are linearly conjugate. Examples 1.6.1 and 1.6.2 show
that all such vector fields are topologically equivalent to the vector field x.
The complete classification of linear vector fields on R2 is summarised in Figure
1.26. Each point of the (Tr A, Det A)-plane represents a similarity class of real,
2 x 2 matrices. The striking feature is that the vast majority of points in Figure
1.26 correspond to vector fields of stable, unstable or saddle type. Such linear
vector fields are said to be hyperbolic (see §2.1) and Figure 1.26 suggests that
hyperbolic behaviour is ‘typical’ for linear vector fields on R2. The point to note
is that, without a suitable equivalence relation, the idea of what is typical has no
meaning. We will return to the question of typical or generic properties of flows
and diffeomorphisms in §3.1.
Figure 1.26 Topological types of all linear vector fields on the plane.
Each point in the (Tr A, Det A)-plane corresponds to an equivalence
class of linear vector fields. Details of the derivation of this diagram are
given in Chapter 2 of Arrowsmith & Place (1982). The differential
equation x = x has Tr A = 2, Det A = 1 and is therefore unstable.
Det A
STABLE J, UNSTABLE
jU
SADDLES
34 1 Diffeomorphisms and flows
to <pr!, i.e. the maps are of different topological type. We will have cause to return
to time-r maps of flows in Chapter 5.
Another, more significant, way of obtaining a diffeomorphism from a flow is to
construct its Poincare map. Let <p be a flow on M with vector field X and suppose
that E is a co-dimension one submanifold of M satisfying:
(i) every orbit of ip meets I for arbitrarily large positive and negative times;
(ii) if x e E then X(x) is not tangent to E.
Then E is said to be a global (cross) section of the flow. Let y e E and t(y) be the
least, positive time for which y „y)(y)eE.
D efinition 1.7.1 The Poincare (or first return) map for E is defined to be
Exam ple 1.7.1 Obtain the Poincare map, P, of the flow defined by
where (r, 0) are plane polar coordinates, taking E to be the half-line 0 = 0. How
does P change if E is taken to be the half-line 0 = 0„?
Solution. The phase portrait of (1.7.2) is shown in Figure 1.15. E is the positive
x-axis in the plane and (1.7.1) can be written
where x(r) = 2k and (-)r denotes the radial component of •. We, therefore, conclude
that P takes the form (1.7.5) with x replaced by r, the radial distance along 0 = 0O.
□
By construction P: E -> E is a diffeomorphism and dim E = dim M - 1. In contrast
to time-t maps we, therefore, expect these diffeomorphisms to reflect the properties
1.7 Poincaré maps and suspensions 35
of flows in one higher dimension. For example, the Poincare map P(x) in (1.7.5)
has a fixed point at x = 1 (observe x* = P(x*) implies (1 - x*)(l —a) = 0, which
is only satisfied for x* = 1). Furthermore, if x ^ 1, then P{x) $ x so that x = 1 is
an attracting fixed point. This fixed point in P clearly corresponds to the stable
limit cycle in the phase portrait of <p (see Figure 1.15).
Another example is afforded by the flow on the torus, T 2, defined by
where 9 and <pare as shown in Figure 1.27. The equations (1.7.7) have solutions
Figure 1.27 Diagram showing how the coordinates, 0 and <p, used in
(1.7.7) are defined.
36 1 Diffeomorphisms and flows
observation. It means that any result that can be proved for diffeomorphisms
should have a counterpart for flows in one higher dimension (see Smale, 1967).
The following explicit definition is given on p. 59 of Arnold & Avez, 1968.
where x e M , 0 e [O ,1] and [•] denotes the integer part of • , defined on a compact
manifold M by identification of (x, 1) and (f(x), 0) in the topological product
M x [0,1], is called the suspension of the diffeomorphism f : M -> M.
A B
38 I Diffeomorphisms and flows
representative of this equivalence class which clearly exhibits the connection with
the diffeomorphism f. When looked at from this point of view it is easier to
understand how the nature of f (whether or not it is deformable to idM) affects
the resultant manifold on which the suspension is defined.
where
for all re R. The transformation t 1—♦r/y, X(x, i)b->yX(x, yf), with y = TjIn, allows
(1.8.1) to be written as the autonomous system
defined on M x K, where
for all 0e R (see Exercise 1.8.1). It is then convenient to identify 0 4- 2nm, m eZ,
with 0 to obtain a differential equation on M x S \ where 0 is the circular
coordinate. This procedure is illustrated in Figure 1.30 where some possible
solutions of (1.8.1 and 2) are shown. Observe that the solutions are not necessarily
periodic (see Exercise 1.8.2). However, it is easily verified that if £(t) is a solution
of (1.8.1 and 2) then so is ?(t + T) (see Exercise 1.8,2). i.e. advancing a solution
by one period of the vector field also gives a solution.
Figure 1.30 helps us to associate this ‘period advance map’ of the non-
autonomous system with the Poincare map P9: -> of (1.8.3) defined on the
global section, £ e = M x {0} of M x Sl. It is worth noting that Pe and P„., 0 ^ 0 '
are topologically conjugate (see Exercise 1.8.4). Thus, in discussing topological
properties it is sufficient to consider P = P0. Conversely, we can associate the
solutions of the non-autonomous system with the suspension, on M x S1, of the
Poincare map, P, which is itself a diffeomorphism on £ 0 = M x {0}.
There is a complete correspondence between the properties of the Poincare map,
P, and those of its suspension. For example, P has a fixed point x* if and only if
its suspension has a closed orbit of period In, i.e. if and only if the non-autonomous
system has a periodiq solution of period T. Figure 1.30 shows a 2-cycle of P along
with the corresponding solution of (1.8.1,2) with period 2T. Furthermore, a
periodic solution of (1.8.1,2) is stable (asymptotically stable), in the sense of
Liapunov, if and only if the associated periodic point of P is stable (asymptotically
stable). The following example shows how this last result can be applied.
1.8 Periodic non-autonomous systems 39
0= — mod I k
T
40 ] Diffeomorphisms and flows
Exam ple 1X 1 Find the period advance map for the non-autonomous system
where a>(t) = w(t + T), t e R. Obtain the Poincare map P and show that Det P = 1.
Hence, deduce that the null solution of (1.8.5) is stable (in the sense of Liapunov)
if |Tr P| < 2 and unstable if |Tr P| > 2.
Solution, The second-order equation (1.8.5) can be written in the first-order form
x = A(t)x, ( 1. 8. 6)
(1.8.7)
A ,,,= U ° < - > ] ‘ I»)-
The solutions of (1.8.6) form a two-dimensional vector space (see Exercise 1.8.3).
The solution {(f) satisfying £((„) = x 0 can be written in the form
where the columns of Q(f) form a basis for the solution space of (1.8.6). Q (t) is
called a fundamental matrix for the problem (see Jordan & Smith, 1977) while
<p(t, r0) is known as the state transition matrix (see Barnett, 1975). Now observe that
Thus ip(t + T,t): I ,- » E ,+T (see notation in Figure 1.30) is the period advance
map at t. Clearly, if { and if are solutions of (1.8.6) then
<p(t + T, t)(a((t) + bq(t)) = a<p(t + T, t)Z(t) + b<p(t + T, t)if(t),
a,b^U , and the period advance map is linear for any t. Moreover,
v (t+ T ,t) = Q(t + T K r \ t )
since Q(t) = A(t)Q(f). It follows (see Exercise 1.8.6) that, if W(t) —Det(^(t, 0)), then
fV(t) = Tr(A(t))lT(t) = 0 for (1.8.6). Hence W(t) = 1T(0) = Det(v (0,0)) = 1 and, in
particular,
W(T) = Det(«p(T, 0)) = Det P = 1. (1.8.12)
The null solution of (1.8.6) corresponds to the fixed point of P at the origin.
The stability type of the null solution is the same as that of the fixed point and
the latter is determined by the eigenvalues, A, 2, of P. Since D e t P = l , the
characteristic equation of P is A2 —(Tr P)A + 1 = 0 and
If |Tr Pj < 2 then (Tr P)2 < 4 and the eigenvalues are complex with A2 =Af =
exp(i)3) (since AtA2 = 1), where tan /?=[4 —(TrP)2] 1,2/T rP . Let u + iv, u, veR2, be
the eigenvector of P with eigenvalue A,. Then the matrix K = (v-u) is such that
cos ft —sin p
K>PK (1.8.14)
,sin ft cos P.
Figure 1.31 When |Tr P| < 2, the orbits of points x / 0 under P lie on
ellipses as shown. Observe that, for any xeN', P mxeN for all meZ.
Thus, x = 0 is stable in the sense of Liapunov (see Definition 1.2.3).
x2
42 l Diffeomorphisms and flows
Here the orbits of P lie on hyperbolae and, as Figure 1.32 shows, the x = 0 is
an unstable fixed point. □
Example 1.8.1 suggests that periodic perturbations of the frequency, a>, of a
harmonic oscillator can de-stabilise the equilibrium point with x = 0. This is
essentially what a child on a swing achieves by appropriate movements of weight,
in order to build up the amplitude of the oscillations of the swing. A simple example
illustrating how this instability can be achieved is given in Arnold (1973, pp. 205-6).
This phenomenon is known as parametric resonance.
(1.9.1)
Figure 1.32 For k > 1 the hyperbolae x,x2 = c, c / 0, are invariant curves
for the map Dx, where D is given by (1.8.15). The origin is a hyperbolic
saddle point and therefore for every N' £ N, there exists xeN' for which
P"x $ N. for some me 2 +. Hence the saddle point is unstable in the sense
of Liapunov.
1.9 Hamiltonian flows and Poincaré maps 43
H = //(q, p) is the Hamiltonian for the system and the equations (1.9.1) are known
as Hamilton’s equations. The state of the system at time t is specified by
The configuration, q(t), of the system is given by the n generalised coordinates q,(t)
and p(t) consists of the n conjugate generalised momenta p((t). A system with n
degrees of freedom is often called an n-F system.
In general, q, and p, change with t but H does not. Observe
W : M -» R
Wfi
R R
=H b }'
44 1 Diffeomorphisms and flows
overlap map hai (see Figure 1.3). Thus, if x = (qi , . ..,q „ p t, . . . , p„)T in Ua and
y = (<2i, • • •, <2„, P i y -, PJT in Up represent the same point on M, then
and
Of course, we require that the vector fields X„a and XHfl give rise to the same
dynamics on the overlap between two charts and this imposes constraints on the
manifold itself. To make the dynamics on W, and Wp agree on W ,n Wp, we demand
that
t o%. dU. t o . dH \
dqyh ’ " " dq .’ dP i" " ' d p j
SQ i dQ t d Q i\
№ . .
dqt dq„ dP l dp„
top t o p dH p dH p\
(1.9.7)
\<>Qi d Q „ 'd P " S P j
SP„ t o .
\dqi 3 p J
[ D A ( x ) ] T = SXWi(x) (1.9.8)
and
/0 -I\
with S = I I and 1 equal to the n x n unit matrix. Operating from the left
[ D A ?(x)]TSDxhap(x)XH>(x) = [ D A / x ^ S X ^ x ) ) , (1.9.10)
Clearly, (1.9.5) is satisfied if and only if the overlap map is such that
[ D A /,(x)]TSD,ha/I(x) = S, (1.9.13)
for each x e h j W .n Wf ).
[Dh(x)]TSDh(x) = S (1.9.14)
A differentiable manifold for which all the overlap maps satisfy (1.9.13) is said to
be a symplectic manifold. The theory of symplectic manifolds provides a coordinate
free approach to Hamiltonian mechanics (Abraham & Marsden, 1978; Arnold,
1968).
It is important to realise that (1.9.13) is sufficient to ensure that the form (1.9.1)
of Hamilton’s equations is valid on both Ua and U^ (see (1.9.8,9)). The arguments
involved in obtaining (1.9.13) are not confined to overlap maps. Consider the
effect of a coordinate transformation, h, on a Hamiltonian system defined on R2".
If we demand that the equations of motions of the new coordinates be derived
from the transformed Hamiltonian by applying (1.9.1), then we can conclude, by
precisely the same steps as we have used above, that h must be symplectic. However,
preservation of Hamilton’s equations in this sense is the property that defines
canonical transformations in Classical Mechanics. Thus symplectic and canonical
transformations are one and the same thing.
A property that distinguishes a Hamiltonian flow, tpf, from other flows of even
dimension is that tp* preserves volumes of phase space.
Theorem 1.9.1 (Liouville) Let tp, be the flow induced by x = X(x) and fi(t) be the
volume of the image, tp,(D), of any region D of its phase space. I f div X = 0, then
tp, preserves volume, i.e. il(f) = 0(0) for all t.
To illustrate the ideas behind the proof of Theorem 1.9.1 we will assume that D
and tp,(D) both lie in the same chart. Since tp, is a diffeomorphism for each t, we
can regard it as a change of coordinates in phase space. With notation in Figure
1.34,
(1.9.16)
46 I Diffeomorphisms and flows
dH Det(Dp,(x)) - 1 U2n
= ft(0) = Lim j ( d2"x, (1.9.19)
dt r=0 J f l\
since fl(0) = j DAql , . . . , dp„. However, observe that if DX(x) has eigenvalues q,{x)
then
= n a + t » , i( x )+ o ( t2)),
i
Of course,
Figure 1.34 The flow map ip, takes D at time zero (see (a)) to <p,(D) at
time t (see (b)). Since <p, is a difleomorphism, this transformation can be
regarded as a change of coordinates from (qt ....... q„, p, , . . . , p„) = xT to
1.9 Hamiltonian flows and Poincare maps 47
The above arguments do not depend on the initial time being zero and (1.9.22)
can be generalised to
d
div X(x) = i — ( — (1.9.24)
,-id q ,\d p t dp,
Hence (pf preserves phase space volumes. This result highlights, in a geometrical
way, the very special nature of Hamiltonian flows. In general, even dimensional
flows may expand volumes in some parts of phase space and contract them in
others. Clearly, (1.9.24) imposes a global restriction on <p?. The volume-preserving
nature of Hamiltonian flows is also reflected in the nature of the transformations
that relate them to one another. It can be shown (see Arnold, 1968, p. 222 and
Exercise 1.9.5) that (1.9.14) implies Det(Dh(x)) = l so that symplectic trans
formations preserve volumes of phase space. However, it is perhaps worth noting
that D et(D h(x))s 1 only implies h is symplectic when h: R2 -> R2 (see Exercises
1.9.5 and 1.9.6).
It is reasonable to consider to what extent Hamilton’s equations can be simplified
by symplectic transformations. Let h: (q, p) -> (Q, P) and H(Q, P) = W(h“ 1(Q, P)).
In particular, the transformed equations will be simpler if the new Hamiltonian
is independent of one of the generalised coordinates. For example, suppose H does
not depend on Q„, then
dH
(1.9.25)
dQ„
and
• dH
Q, = r - = « i(/ i . - - . U (1.9.28)
dPi
48 1 Diffeomorphisms and flows
Qi=Pu Pi = ~«>i9i.
(1.9.30)
< <h = P2> P2=~«2<i2-
The Hamiltonian //(q, p) is given by
W(<ii.«2;Pi.P2) = 5 £ ( P t + t » f q f ) (1.9.31)
1.9 Hamiltonian flows and Poincaré maps 49
q , = A t c o s ( « , i + i/i ) ,
(1.9.32)
Pi - 4
— w f/ ( s i n ( c o ,i + t h ) ,
i(fe R, i = 1, 2. The aim is to construct the Poincare map in such a way that one
pair of conjugate variables (q2, p2, say) are removed. Thus we argue that by
restricting to the Hamiltonian shell H(q, p) = h0 > 0 we can express p2 in terms
of qi,Pi and q2. Since q2 is periodic with period ln/(o2, the orbit of a phase point
in the plane q2 = 0 returns to q2 —0 after time 2n/a>2 (see Figure 1.35). Therefore,
the Poincare map P defined on the section q2 = 0 is given by
l a cos ( co, Ti + —
/4, 2nl +»)!
V L «2J
2nl
—0)i A l Sin^iÜ!^ t + — \ + n2
0)2J
cos2np + — s i n 2 n u |/ (j I\
P2i+(oiqi = C, (1.9.34)
with 0 < C < 2h0, are invariant curves. These closed invariant curves correspond
to invariant tori in the flow on the H = h0 shell.
The important thing to notice about (1.9.33) is that Det P = 1. This means (see
Exercise 1.9.5) that the Poincare map, constructed in the manner described above,
is area-preserving. That this is also the case when the system is non-integrable
follows from the Poincare-Cartan invariant (Arnold, 1968, pp. 233-40 or Arnold
& Avez, 1968, pp. 230-2). A derivation of this invariant for 2-F systems requires
^■92 =0
50 l Diffeomorphisms and flows
Figure 1.36 (a) Tubular region to which Stoke’s Theorem is applied for
1-F systems. The vector field curl v is tangent to the surface at every point
of the tube so that curl v■dS = 0. The closed curves )>! and y2are obtained
by taking sections transverse to the tube of flow flines. (b) Dissection of
the tube shown in (a) used to obtain (1.9.35).
Figure 1.37 If y2is given by r = r(u), then r(u) = (q,(u), 0, p,(u), p2(u), t(u)),
where p2(u) is determined by H(q, p) = h0. The curve, y2, obtained by
projecting y2 onto t = 0 (see (a)), is the image of y, under the Poincare
map P (see (b)).
(o) (b)
1.9 Hamiltonian flows and Poincaré maps 51
j c u rlv d S = 0 = j v d r - f v d r , (1.9.35)
Js Jy, Jyi
where dr = (dq, dp, dt). Thus
j p dq —H dr = j pdq —H dt (1.9.36)
4
pl dql + p2 dq2- H dt= \ Pi dg, + p2 dq2 - H dt,
Figure 1.38 Some typical orbits of the Henon map (1.9.40) for cos a = 0.8.
(1.9.37)
Two fixed points can be seen: one elliptic (see §6.5) and one saddle-like.
What appear to be closed curves are each the orbit of a single point, i.e.
the orbit is confined to what is topologically an invariant circle. For
small numbers of iterations of (1.9.40) individual points of these orbits
can be distinguished moving around the origin (cf. Exercise 1.9.9). As
the number of iterations increases, the plotted points merge into what
looks like a closed curve. Individual orbit points are more apparent in
the vicinity of the saddle point. (After Henon, 1969.)
52 1 Diffeomorphisms and flows
where V! and y2are closed curves bounding a tube of the flow in the five-dimensional
extended phase space with coordinates (qlt q2, p,, p2, t). Now let y2 consist entirely
of points such that H = hg,q 2 = t = 0. Suppose we follow the lines of the flow </>,"
until we return to q2 = 0. Although H remains at h0 and q2 returns to zero, the
Figure 1.39 A selection of plotted orbits of (1.9.40) for cos a = 0.4. Orbits
of points near to the saddle point become highly irregular. Successive
iterates still move around the fixed point at (0,0) but they are no longer
confined to a closed curve. Instead they appear to spread over a
two-dimensional region in an erratic manner. Eventually, they are pulled
away along the unstable manifold of the saddle and, left to themselves,
will cause an overflow error in the computer doing the plotting. On the
other hand, orbits of points near the origin still appear to be confined
to invariant circles. Between these extremes, a new feature called an
island chain can be seen. The ‘islands’ themselves are formed around the
points of an elliptic periodic orbit, here of period six. The ‘straits’ between
successive islands contain a hyperbolic periodic orbit also of period six.
The orbits of points near to the elliptic periodic points move from island
to island, returning to an invariant circle surrounding the initial elliptic
point at every sixth iteration. Some information to help the reader to
observe island chains is given in Exercise 1.9.9. (After Henon, 1969.)
1.9 Hamiltonian flows and Poincaré maps 53
H d t = \ p2 dg2 = 0 (1.9.38)
Jn J y<
for i = 1,2, so that (1.9.37) becomes
Pi d?i = (1.9.39)
1.
where y2 is the projection of y2 onto t = 0. Now y2 is the image of y, under the
Figure 1.40 Analogous plots to those shown in Figure 1.39 but with
cos a = 0.24. Observe that a five-fold island chain is the dominant feature
here. In fact, (see §6.5) island chains of all periods occur but only a few
are easily visible. The orbits looking like séparatrices of the hyperbolic
periodic points are deceptive (see Figure 1.41). (After Hénon, 1969.)
54 1 Diffeomorphisms and flows
where a is a real parameter and t e Z (see Hénon, 1969). This map is not constructed
Figure 1.41 Two orbits of (1.9.40) for cos a = 0.22. The first is the orbit
of a point near an island centre giving invariant circles around the five
elliptic periodic points. In the present context, it serves only to indicate
the position of the islands. The remaining points are all generated by
iterating a single initial point. Once again, the iterates spread out, in a
stochastic manner, over a two-dimensional region in the neighbourhood
of what appeared to be séparatrices in Figure 1.40. (After Hénon, 1969.)
x
1.9 Hamiltonian flows and Poincare maps 55
A
56 1 Diffeomorphisms and flows
Exercises
1.1 Introduction
1.1.1 Let M be the unit circle in the complex plane. Explain how the map it: R -* S1
given by xi-*exp(ix) can be used to define a set of charts on the circle S'. Define
explicitly two charts which form an atlas on S'. What is the differentiability of
the overlap maps for this atlas?
1. 1.2 Let M be a (7-manifold. Show that if f: M -»M is a C*-map, k < r, and the points
x0 and f(x„) are in overlapping charts (L/„, h„), (Vr hy) and (Uf, hp), (Ut, h4)
respectively, then the map
ly f-h ; (El.l)
is (7 at h„(x0) if and only if
M V (E1.2)
is C‘ at hf(x0). What does this imply about the differentiability of f at x0?
1.13 (a) Find an atlas of the torus T2= {(x mod 1, y mod l)|(x, y)eR2} containing
four charts by using the local diffeomorphism n:R2 -* T2 given by (x,y)i~>
(xmod 1, ymod 1).
(b) Use stereographic projection on the unit sphere to obtain an atlas consisting
of two charts. Construct the overlap map between these two charts.
2.5 Consider an orientation-retiming homeomorphism f :S‘ ->Sl. Show that any lift
/: R -» R of / will have the property that / is strictly decreasing and f(x + 1) =
f(x) - 1. Show that / always has two fixed points. Is f 1 orientation-reversing or
-preserving?
1.4.6 Use the Poincare-Bendixson Theorem to show that the Van der Pol oscillator
x + e{x2 —l) i + x = 0, £ < 0, (El.9)
has at least one stable limit cycle for sufficiently small values of e.
1.5 Conjugacy
1.5.1 Show that:
(i) f(x) = 2x and y(x) = 8x are topologically conjugate on R but not differentiably
conjugate;
(ii) f(x) = 2x and g(x)=-2x are not topologically conjugate by showing
conjugacy preserves orientation of a map;
(iii) f(x) = 2x and g{x) = \x are not topologically conjugate by investigating the
nature of the Fixed point at the origin in the two cases.
1.5.2 Prove that q>: x ^ x 2"*1, neN, is a topological conjugacy of the diffeomorphisms
f(x) = 2x and g(y) = 22"+'y on R. Why is there no differentiable conjugacy when
n>0.
1.5.3 Let f, g:R“-*R" be C‘-conjugate (Ic^l) diffeomorphisms by h: R"-> R” and
suppose f(0) = 0. Prove that the Jacobian matrices of f at 0 and g at h(0) are
similar. What does this imply about the eigenvalues of Df(0) and Dg(h(0))?
Show that f{x) = ax and g(x) —fix are not C*-conjugate for a # /1. When are
/ and g not C°-conjugate?
1.5.4 Let f,g be diffeomorphisms on R given by
f{x) = X + sin X (E1.10)
and
g{x) = x + hk(x). (El.11)
where
x3 Xs . x Zi+I
-----1-----. . . _j_(— J) ----------------- (E1.12)
3! 5! (2fc + l)l
k e l* . Prove that / and g are not topologically conjugate for any k.
1.5.5 Find the number of period-2 points of the diffeomorphisms / and g of the circle
S', where the lifts are:
f{x) = x + 0.5 + 0.1 sin 2irx; (El.13)
g(x) = x + 0.3 + 0.1 sin 2nx. (El.14)
Show that / and g are not topologically conjugate.
1.5.6 Prove that the two linear systems x = Ax, y = By, x, yeR", have flows which are
linearly conjugate if and only if the matrices A and B are similar.
1.5.7 Let tp, $ be flows on R" and suppose that <p has a fixed point at the origin. If <p
and f are C‘-conjugate, k > 0, by h: R" -> R", show that the vector fields X and
Y of <pand ^ are such that the matrices DX(0) and DY(h(0)) are similar. Compare
your answer with that of Exercise 1.5.6 and comment on why the converse of the
above is not true in the non-linear case.
1.5.8 Find the rotation number of the circle homeomorphism / with lift /: R -* R given
Exercises 59
by:
(i)f(x ) = x + î;
(ii) /(x) = x3 + J, 0 ^ x < 1, /(x + l)= /(x ) + 1;
(El. 15)
ÎK
by locating fixed or periodic points of /.
are topologically equivalent. Hence prove that all linear flows, exp(At), for which
A has pure imaginary eigenvalues are topologically equivalent to exp(J0t). Why
are these flows not all topologically conjugate?
6.2 It is often easy to recognise why two flows are not topologically equivalent by
noting key features in their phase portraits. Describe, for each diagram in Figure
El.l, a distinguishing feature preserved by topological equivalence which is not
shared by the others. Explain your answers. Why is the invariant circle in (b) not
a limit cycle?
6.3 Show that the topological types of a saddle and node are different by considering
the séparatrices of the saddle.
6.4 Consider the system x = X„(x), a e R,
x = l, y =a (El.16)
on R2 and the flow on T2 induced by the map n: R2 -» T2 where n(x, y) =
(x mod 1, y mod 1). Show that the phase portraits of the system (El.16) for a (a)
rational and (b) irrational are not topologically equivalent.
6.5 Let if and if' be topologically conjugate flows on the manifold M and and iji’
be topologically conjugate flows on the manifold N. Prove that the product flow
60 I Diffeomorphisms and flows
n+ X
a , , . .. , a„+,, not all zero, such that £ a,^(0) = 0. Use uniqueness of solution
f=i
»+1
to show that £ a,■£,(() = 0.
i=i
<c) Show that the vector space of solutions is n-dimensional.
.8.4 Define the state transition matrix q>(t, t0) of the system x = A(t)x, (x, t)e R" x R,
A(t + T) = A(t). Prove that
(i) <i>(t + T, t0 + T) = t0), ,F1
(ii) f(f,0 )-' = 1K0,t). 1 '
Use these results to show that
q>(t + T, t) = <p(0, t)- V(T, 0H»(0, t). (El-19)
What does this imply about the family of Poincare maps P,o: E#0-* £6o, 0O6 [0, 2ji]
of the system
x = A(0)x, 0= 1, (El.20)
where A(6 + 2a) = A(0), and = {(x,0o)|xeR2}?
8.5 The state transition matrix is useful where A is independent of ( or aperiodic.
8.6 Let q>(t, t0) be the state transition matrix of the system x = A(i)x and define W(t) =
Det(</>(i, 0)). Prove
(i) <f(t + h,0) = q>{t, 0) + hk{t)q>{t, 0) + 0(h2); (E1.21)
(ii) W(t + h) = Det(l + hk (t) + 0(/i2))W(t)
show that
x = A(f)x + B(t) (El.26)
has solution
when x = x0 at t = t0.
Find the solution of
■(! (E1.28)
\1 ‘ 0/
V i \cos
“ 'tj)
when x(0) = x0.
2.8.6 Use blowing-up techniques to determine the topological types of the singularities at
the origin in the following systems:
(a) x = y + x3, y = x3; (E2.42)
(b) x = x2 —y2, y = —2xy. (E2.43)
Explain why the technique fails to determine the topological type of the singularity
at (x, y) = 0 in the system x = y + x3, y = - x 3.
3
Structural stability, hyperbolicity, and
homoclinic points
Definition 3.1.1 A linear flow, exp(At): IR" -+ R", (or iiffeomorphism, A) is said to
be structurally stable in L(R") if there is an s-neighbourhood of A, Nt(A )cL (R n),
such that, for every Be NC(A), exp(Bt) (or B) is topologically equivalent (conjugate)
to exp(At) (or A).
The following result shows that, in this linear case, the structurally stable systems
can be completely characterised.
Proof. A linear flow exp(At) is hyperbolic if all the eigenvalues of the matrix A
have non-zero real parts (see Definition 2.1.2). The eigenvalues of any £-close
matrix B differ from those of A by terms 0 (e) (see Exercise 3.1.1). Thus, by making
e sufficiently small we can ensure that the eigenvalues of B are near enough to
those of A for their real parts to be non-zero. Moreover, A and B will then have
the same number, ns(n„), of eigenvalues with negative (positive) real parts. Theorem
2.1.2 then implies that exp(At) and exp(Bt) are both equivalent to the flow of
x = —x, y = y, where xeR "s and yeR "0. Thus A is structurally stable.
Conversely, suppose the flow exp(At) is not hyperbolic. Then A has at least one
eigenvalue with zero real part. However, B = A + e! is hyperbolic for almost all
e ^ 0 and can be made arbitrarily close to A by taking s sufficiently small. Thus, the
non-hyperbolic flow exp(At) is not structurally stable. Hence, if a linear flow is
structurally stable it must be hyperbolic (see Figure 3.1).
The proof of Proposition 3.1.1 for diffeomorphisms follows similar lines; it is
considered in Exercise 3.1.2. □
Proof. As we have already noted in the opening remarks to this chapter, SF(R")
must be open. More precisely, by Definition 3.1.1, each AeSF(R") lies in an
^neighbourhood, iVt(A), of topologically equivalent flows. Since Ne{A) is open,
each of its elements must also be structurally stable, i.e. Ne(A) E SF(R"). Hence
SF(IR") is open because it is a union of open sets.
It follows from Proposition 3.1.1 that SF(R") = HF(R")( e L(R")) the set of linear
transformations which give rise to hyperbolic flows on R". However, HF(W) is a
dense subset of L(R"). To show this, suppose A 4 HF(R"). Then there exist linear
transformations, B, arbitrarily close to A for which exp(Br) is hyperbolic (e.g.
(a) (b)
-*3 *3
(c) (d)
3.2 Local structural stability 123
B = A + el, with e small). Thus, every element of L(R") is arbitrarily close to points
of HF(W) and HF(R”) = SF(R") is a dense subset of L(R"). Hence, structural
stability of flow is a generic property of L(R"). □
The analogous result for hyperbolic, linear diffeomorphisms is dealt with in Exercise
3.1.2.
3.2 Local structural stability (Hirsch & Smale, 1974, Chapter 16)
An important outcome of the discussion in §3.1 is that hyperbolic fixed points of
linear flows and diffeomorphisms persist under sufficiently small linear perturbations.
This result can be extended to the local flows or diffeomorphisms defined in the
neighbourhood of hyperbolic fixed points of non-linear systems. Of course, we
must choose an appropriate class of perturbations and specify when they are small.
Let U be an open subset of R" and Vec‘(l/) be the set of all C ‘-vector fields
defined on U. The magnitude of each vector field X eV ec‘(l7) is taken to be its
C'-norm, || X j| i, where
dX‘(x)
11X1,= supP i £t \X'(x
I i+ I (3.2.1)
V li=l u=l dX:
Thus, if X(x) = (X‘(x ),. . . , X"(x))T then ||X|| i is ‘small’ when X'(x) and dX‘(x)/dxJt
i , j = 1 , . . . , n, are ‘small’ for all x e U. We now define
Figure 3.2 The flow in (a) is an allowed perturbation of the centre, (b),
in L(R2) but it is not an element of CL(R2). For a sufficiently weak
spiral, (a) can be made arbitrarily close to (b) in L{R2).
124 3 Structural stability, hyperbolicity and homoclinic points
Proposition 3.2.1 Let XeVec l(U) have a hyperbolic singularity at x = x*. Then
there exists a neighbourhood V of x* in V and a neighbourhood N of X in Vec‘(U)
such that each Y e N has a unique hyperbolic singular point y* e V. Moreover, the
linearised flow, exp(DY(y*)i) has stable and unstable eigenspaces of the same
dimension as exp(DX(x*)t).
Notice the perturbed fixed point y* does not, in general, coincide with x*. However,
given any 5 > 0 ,N can be chosen such that |y* - x*| < <5for all Y e N. Furthermore,
by using Proposition 3.2.1 in conjunction with Hartman’s Theorem (Theorem
2.2.3) we can deduce that there are neighbourhoods of x* and y* on which X and
Y are topologically equivalent. Proposition 3.2.1 gives equality of dimension of
the stable eigenspaces of the linearised flows exp(DY(y*)t) and exp(DX(x*)t). It
follows that these linear flows are topologically equivalent (see Theorem 2.1.2).
The Hartman Theorem states that there is a neighbourhood of x* (y*) on which
the flow of x = X(x) (y = Y(y)) is topologically conjugate to exp(DX(x*)f)
(exp(DY(y*)t)). Thus the local, C°-equivalence of the flows of X and Y follows from
where <p, and are the flows of X and Y respectively. U"> and 1/,» g U denote
the neighbourhoods on which Hartman’s Theorem is valid. There is, therefore, a
sense in which <pt: U -» R* is structurally stable in a neighbourhood of x*. Namely
that for every pair (Y, y*) with Y e N there is a neighbourhood (/*'. g i/ of y* such
that is C°-equivalent to <p,\U$. Thus, one might say that <p, is locally
structurally stable at x*. Alternatively we can observe that (3.2.2) means that the
topological type of the fixed point (see opening paragraph of Chapter 2) is preserved
under all sufficiently small C1-perturbations and say that the type of fixed point
is structurally stable.
Hyperbolic fixed points of diffeomorphisms also persist under sufficiently small
C'-perturbations. Let DifT1(t/) be the set of C'-diffeomorphisms f: l/( g R " ) -*■ R"
with the C’-norm. Then the following result parallels Proposition 3.2.1 for flows.
(3.3.1)
The above precautions are not enough to focus attention on the structural
instabilities occurring in the interior of D2. Unfortunately, instabilities associated
with the behaviour of the vector fields at the boundary of the disc can still occur.
In the absence of further constraints, vector fields that are tangent to 5D2 are still
present in Vec‘(D2). Topologically distinct C 1-perturbations of such vector fields
are illustrated in Figure 3.3. These unwanted structural instabilities can be excluded
by confining the discussion to those vector fields that are transverse to dD2. Clearly,
a vector field that satisfies this requirement is a member of one of two disjoint
126 3 Structural stability, hyperbolicity and homoclinic points
subsets of Vecl(D2): it either points into or out of D2 at every point of dD2. Let
VeciJD2) be the set of vector Fields defined on the disc D2 in the manner described
above and such that X(x) points into D2 for every x in dD2. Then the following
theorems generalise Propositions 3.1.1 and 3.1.2 for linear vector fields.
Notice that items (i) and (ii) in Theorem 3.3.1 simply ensure local structural stability
of the fixed points and closed orbits in the flow of X. It is really only item (iii)
that involves a global property of the flow.
Theorem 33.2 The subset of vector fields in Vec‘„(D2) that are structurally stable
is open and dense in Vecfn(D2).
Obviously, parallel results could be stated for those vector fields that point out
of D2 at every point of dD2. Indeed, the flows of vector fields in Vec'„(D2) and
Vec‘„,(D2) are in 1:1 correspondence by time reversal.
By stereographic projection (see Figure 3.4(a)), the unit disc D2 is homeomorphic
to a closed cap, C2, based on the south pole of the sphere S2. The boundary
condition on dD2 can then be replaced by considering flows on S2 with a single
repelling fixed point in S2\C 2. This is conveniently placed at the north pole (see
Figure 3.4(b)). Provided that this additional fixed point is hyperbolic, a structurally
stable vector field on D2 is also structurally stable on S2. This follows from the
work of Peixoto (1962) who extended Theorems 3.3.1 and 3.3.2 to vector fields
on two-dimensional, compact manifolds. Let M be a two-dimensional, compact
manifold without boundary and let Vec*(M) be the set of C1-vector fields on M
*0 *o
(a) (b)
3.3 Flows on two-dimensional manifolds 127
with the C'-norm. This norm is defined by imposing the C'-norm on each of the
charts of a finite atlas for M. Then, Peixoto’s result may be stated as follows.
Theorem 3.33 (Peixoto) A vector field in Vec*(M) is structurally stable if and only
if its flow satisfies:
(i) all fixed points are hyperbolic;
(ii) all closed orbits are hyperbolic;
(iii) there are no orbits connecting saddle points;
(iv) the non-wandering set consists only of fixed points and periodic orbits.
Moreover, if M is orientable the set of structurally stable C‘-vector fields forms
an open dense subset of Vec '(Ai).
Here orientable simply means that two distinct sides of M can be recognised. The
sphere, torus, pretzel, etc. are all examples of orientable manifolds.
The statement of Theorem 3.3.3 contains an additional condition (iv) that does
not appear in Theorem 3.3.1. Its role can be illustrated as follows. Consider the
irrational flow on the torus T 2. This flow satisfies items (i)-(iii) vacuously. There
are no fixed points, closed orbits or saddle connections. However, it fails to satisfy
item (iv) because the non-wandering set is the whole of T 2. Therefore, the irrational
flow is not structurally stable on T 2. Clearly, there exist e-C 1-close rational flows
for which every orbit is closed. Some examples of structurally stable flows on S2
and T 2 are shown in Figure 3.5.
It should perhaps be noted that, since M is compact, flows on it can only have
finitely many fixed and periodic points if they are all hyperbolic. By the Hartman
Figure 3.4 (a) The stereographic projection from the north pole N of
the sphere. The circle # projects onto the boundary of D2. (b) A vector
field in a neighbourhood of N which cuts # transversely.
128 3 Structural stability, hyperbolicity and homoclinic points
Example 3.3.1 Show that the vector field, X, of the differential equation
x = 2 x - x 2, y = - y + xy (3.3.2)
is not structurally stable on any compact subset of the plane with the line segment
joining the singular points of X in its interior.
Solution. The system (3.3.2) has saddle points at x* = (0,0) and y* = (2,0). On
the x-axis, y = 0 and so there is an orbit connecting these hyperbolic fixed points.
The phase portrait is shown in Figure 3.6(a). Let D be any compact set containing
the common separatrix of x* and y*. Notice that the stable manifold of x* is
the line x = 0 while the line x = 2 is the unstable manifold for y*. This means that
there are points x in the boundary of D for which X(x) points both into and out
of D. Therefore, we are (technically) not able to apply Theorem 3.3.1. However,
consider the one-parameter family of systems
The vector field of (3.3.3) can be made e-C 1-close to that of (3.3.2) on any compact
subset D of the type described above by taking n sufficiently small. The phase
portrait for (3.3.3) with yt > 0 is shown in Figure 3.6(h). The points (0,0) and (2,0)
are saddle points for all real y. However, for non-zero y, y / 0 on the x-axis
between these points. Moreover, the stable separatrix at (2,0) is tangent to y = f/rx.
Therefore, for ft > 0 there is no saddle connection between the fixed points. Hence,
the flows for y = 0 and ft > 0 are topologically distinct. The vector field in (3.3.2)
is consequently structurally unstable on every compact subset of the plane
containing the saddle connection. □
It is perhaps worth noting that, in view of the role played by the boundary condition
on dD2 in Theorem 3.3.1, we can say that a vector field that fails to satisfy the
conditions (i)-(iii) will certainly be structurally unstable independently of the
boundary condition on dD2. Of course, the converse is not true unless the vector
field is transverse to the boundary.
Example 3.3.1 illustrates a useful way of giving meaning to structural instability
on the plane. Some other examples are shown in Figure 3.7. However, we must
not be misled into believing that if a vector field is structurally stable on arbitrarily
large compact subsets of the plane then it is structurally stable on the whole plane.
The following is a counterexample to this erroneous conjecture.
Example 3.3.2 Show that there are arbitrarily large compact subsets of the plane
on which the system
Figure 3.6 Phase portraits for (a) (3.3.2), (b) (3.3.3) with ft> 0.
130 3 Structural stability, hyperbolicity and homoclinic points
is structurally stable. Verify that the topological type of (3.3.4) on the whole plane
is changed by the addition of the perturbation (0, p) to (x, y), however small the
value of p / 0.
Solution. Figure 3.8(a) shows y as a function of y. It follows that (3.3.4) has fixed
points at (y,y) = (0, p), p eZ . These fixed points are alternately stable nodes and
saddle points as shown in Figure 3.8(h). Suppose D is a compact subset of R2 whose
boundary intersects the y-axis at (0, y,), (0, y j with y, < y, and y„ y„ # p for
any integer p. If D is such that y(y,) > 0 and y(y„) < 0 then (3.3.4) is structurally
stable on D by Theorem 3.3.1. Clearly, arbitrarily large D can be constructed in
this way.
Consider the family of vector fields defined by x = X^x), where
Notice that the topological type of the flow of X0 in Example 3.3.2 changes when
/i departs from zero, even though it has only hyperbolic fixed points and no saddle
connections. Since the plane is not compact, infinitely many hyperbolic fixed points
can occur. If we wish to maintain contact with Theorem 3.3.1, we must specify a
boundary condition at infinity. However, if we require that all vector fields point
inward on the boundaries of all sufficiently large discs, we are led back to Theorem
3.3.3 via stereographic projection. In other words, we can obtain structurally stable
flows on R2 from structurally stable flows on S2, however, they will have only
finitely many hyperbolic fixed points and their behaviour at infinity will correspond
to having a hyperbolic fixed point at the north pole of the sphere (see Figure 3.9).
(a) («
132 3 Structural stability, hyperbolicity and homoclinic points
(0 (d)
3.4 Anosov diffeomorphisms 133
all of which are hyperbolic. Moreover, the structurally stable diffeomorphisms form
an open dense subset of Diff'fS1).
Recall (see Proposition 1.5.1) that if / has periodic points then its rotation number
p ( f) is rational. Therefore, the structurally stable diffeomorphisms on S1 have
rational rotation number (the converse of this statement is not true (see Exercise
3.4.1)). I f / is structurally stable with rotation number p( f) = p/q, in lowest terms,
then its dynamics are very simple. It has an even number of period-^ cycles with
stable and unstable periodic points alternating around the circle (see Figure 1.23).
It was hoped that generalisations of the behaviour described in Theorems 3.3.3
and 3.4.1 would not only characterise structurally stable systems in higher
dimensions but would also prove to be generic. To this end, ‘Morse-Smale’ vector
fields and diffeomorphisms were defined (Chillingworth, 1976, p. 231 and Nitecki,
1971, p. 88). Unfortunately, it was found that, while such systems were structurally
stable, their properties did not characterise structurally stable systems in higher
dimensions. In particular, it was shown that there were structurally stable
diffeomorphisms on manifolds of dimension n > 2 (corresponding to vector fields
on manifolds of dimension n + 1 > 3, by suspension) whose non-wandering sets
contained infinitely many periodic points. The Anosov diffeomorphisms of the torus,
T", are a subset of DifTl (T") that exhibit this behaviour. We can describe a
diffeomorphism f on T" in terms of a ‘lift’ in much the same way as we did for
diffeomorphisms of the circle (see § 1.2.2). In this case, the lift f is a diffeomorphism
on IR" which satisfies
where l(k)eZ" (see (1.2.10) et seque). All lifts of diffeomorphisms on T" must satisfy
(3.4.4).
We will begin by describing a special subset of the Anosov diffeomorphisms
that are known as the Anosov automorphisms. The lift of an Anosov automorphism
f is a hyperbolic, linear diffeomorphism, A: IR" - » R", which is such that:
Det A = + 1. (3.4.5b)
134 3 Structural stability, hyperbolicity and homoclinic points
Together (3.4.5a and b) ensure that A " 1: Z" -* Z" and it follows that both A and
A~‘ satisfy (3.4.4). Moreover, given that f(jt(x)) = n(Ax) and g(w(x)) = rt(A '1x),
then
%(*(*))) = f(*(A~ *x)) = jt(AA_1x) = rt(x) (3.4.6)
and
Thus, f 1: T"-+ T" exists and has lift A -1. Finally, observe that f(x) and f _1(x)
are differentiable because Ax and A ~‘x are obviously so and n is a local
diffeomorphism. Hence f is a diffeomorphism on T".
Hyperbolic, linear diffeomorphisms on IR" are often referred to as automorphisms
because they are isomorphisms of the group IR" with itself. The diffeomorphism f
defined above is called an automorphism to distinguish it from other Anosov
diffeomorphisms whose lift is not linear. A general definition of Anosov difleo-
morphisms can be found in Arnold (1973, p. 126) or Nitecki (1971, p. 103) but,
thanks to the following result due to Manning (1974), it will be sufficient for us
to consider only the automorphisms here.
Theorem 3.4.2 Every Anosov diffeomorphism f of T", such that ft(f) = T", is
topologically conjugate to some Anosov automorphism of T".
Coupled with Theorem 3.4.2, the following result shows that such Anosov
diffeomorphisms have complicated non-wandering sets.
*2
3 -X
r*
2 11
11
1 i
.J1
I1
—
f
-3 - 2 -1 0 2 3 X
-1 x(x )
-2
-3
(a ) (b)
3.4 Anosov diffeomorphisms 135
Proof. Let 6 be a periodic point of f, then F{6) = 0, for some positive integer q.
Suppose xeR " satisfies n(x) = 0, then
Now, A* —I is an integer matrix and therefore (A1- I ) ' 1 has rational elements.
Hence x has rational coordinates.
Conversely, if Os T" has representative x = ((p[0>/ r ) , . . . , (p(„0|/r))T, where pS0),
r e l with r ^ 0, then, for any k eZ ,
Pf
A‘x = (3.4.11)
for some integers p\k>, . . . , ptk). However, there are at most r" points on T" that
can be represented in this way and, therefore, there is a q> 0 such that
jt(A*x) = Jt(x). □
Proposition 3.4.1 not only implies that f has infinitely many periodic points; it
also shows that the periodic points are dense in the torus. All these points lie in
the non-wandering set 0 of f and, since fi is closed (see Exercise 1.4.2), we conclude
that ft = T".
The final piece of the argument against Morse-Smale systems was provided by
Mather (1967).
Theorem 3.4.3 (Mather) The Anosov diffeomorphisms on T" are structurally stable
in Diff^T").
A= (3.4.12)
and (3.4.13)
Figure 3.12 Segments of the stable and unstable manifolds of the fixed
point, x(0), of f on T2. Segments AB and A'B' of £“ and Es for A have
been mapped onto the torus using Jt(x) = (x mod 1, y mod 1). All inter
sections of W’ and W‘, except at the origin, are transverse homoclinic
points. Note M,U(M,S) always has +ve(-ve) slope on B,.
3.5 Horseshoe diffeomorphisms 139
exhibited by them can be shown to occur in any map that has transverse homoclinic
points (see §3.7).
y
c D
x
(*)
B A
—1 r r k 1
Figure 3.14 (a) Construction of U Q o ^ Q i)' stretching/contraction yields two strips within the rectangle R;
folding yields four horizontal strips for Q(2) = f ( 2 (1,)r>Q. Images of Q 0 ( Q i ) at each stage are shown. (6)
Illustration of Qtn> for n — 1,2,3. Qln> consists of 2" disjoint horizontal strips whose width, 2/5", rapidly
decreases with increasing n; indeed, the eight strips of Ql3} (shown in black) are barely resolved in this diagram.
142 3 Structural stability, hyperbolicity and homoclinic points
A = n e (n) (3.5.4)
neZ
then A is a Cartesian product of two Cantor sets which is itself a Cantor set.
Proof. Let xeA, then xeQ*"’ for all n e l . Now if xeQ[ neN, then (3.5.3)
implies that f(x) 6 Q{~<n' 1)(n I(Q) £ Q(~ ~ 1,). If x e Q("\ n e Z +, then observe th at:
S t' A ,'
(a )
3.5 Horseshoe diffeomorphisms 143
Together (i) and (ii) imply f(x)ef(Q<n,)n Q = (2<"+11. Thus, if x e QM, for all n e Z,
then f(x)eQ ("+1) for all neZ. Hence f(x)eA.
Similar arguments (see Exercise 3.5.2), with the roles of (3.5.1) and (3.5.3)
reversed, show that A is invariant under 1 and, therefore, f(A) = A. □
Co' Do'
Aa"B
Co Do
Bo' Ao
Bo f-'
C,
C /D ,'
4,
Do Co B| " A|
(.v,>-)h-(|.v, 5y ) replace on Q
B / A ,
(h)
(c)
144 J Structural stability, hyperbolicity and homoclinic points
Q (f(Q) £ Q)- Moreover, it does not have an obvious connection with diffeo-
morphisms on compact manifolds without boundary. However, a diffeomorphism,
g: S2 -* S2 can be constructed such that f is the restriction of g to a subset of the
sphere. The first step in the construction is to extend the map f to a capped square
Q' as indicated in Figure 3.16. The extension f is constructed in such a way that
f |F has a unique, attracting, hyperbolic fixed point. This means that once a point
is mapped into F its orbit subsequently remains in F. The map f can, in turn, be
extended to a closed disc D2 of suitably large radius. The extension g': D2 -* D2
is taken to be such that g'(D2) takes the form shown in Figure 3.17. The
diffeomorphism g: S2 -» S2 is finally obtained by identifying the disc, D2, on IR2
with a cap, C2, on the sphere (by stereographic projection, see §3.3) and adding
a unique, repelling, hyperbolic fixed point in S2\C 2.
Since g is a global diffeomorphism on S2, both g and g~ 1 are defined for all
points of the sphere. However, its construction ensures that g coincides with f or
Figure 3.16 (a) The capped square Q' = GuQuF; (b) the extension
f: Q' Q' is such that G' = f(G) and F' = F(F) are both subsets of F.
Q Q
Figure 3.17 (a) The disc D1 containing Q'\ (b) image of D2 under g'
(shaded) with g'|8' = f'.
(«) (h)
S J Horseshoe diffeomorphisms 145
f when restricted appropriately. Here we have taken the notational liberty of not
distinguishing between these restrictions on the sphere and their representatives
on IR2 via stereographic projection. This distinction does not play a significant
role in our discussion and, once noted, should not lead to any confusion. The
unstable fixed point in S2/C 2 means that the ordinary points of this set move
towards C2 under g. As we have shown in Figure 3.17(b), g|(D2\Q ') is again a
contraction and, therefore, g essentially delivers points to Q'. On Q', g behaves in
the same way as f . We already know that f' has an invariant Cantor set A arising
from its restriction, f, to Q, but what of the points S2\A = A'? The following
proposition provides part of the answer to this question. It states that those points
of Q' that do not lie on the infinite set of vertical line segments, f j G (~"),
neN
are eventually swept into F.
Proposition 3.5.2 The orbits under g of points in Q\( Q{ B)) ultimately approach
neM
the stable fixed point of g in F.
Proof. Figure 3.16 shows what happens to the various parts of Q under a single
application of g|Q ' = f'. The left (L) and right (R) fifths of Q are mapped, together
with G and F, into F. Since F contains a unique, attracting fixed point, the orbit
of any point in F approaches this point asymptotically. Points in the mid-fifth,
Af, of Q suffer the same fate after one more iteration. Such points are mapped
into G by f and into F by f'2. Only points in P0u P , remain in Q (in fact in
Qo'jQi) after one application of f . In other words, points in Q\Q(0) enter F after
at most two iterations of g.
Let us focus attention on the partition of the square Q provided by Q(~ l), rather
than Q(0). Observe that points in Q<0)\(Q ,0)riQ (_1)) are mapped into L, M, R after
one iteration of g |Q = f (see Figure 3.15(b)) and thence into F after two or three
iterations. Thus we conclude that all points in (Q\Q<0)) ^ (Q(0)\(Q <0>o Q<_ 1>)) =
2\(Q <0>r'G ( 1 ) ) enter F after at most three applications of g. Similarly, if
we consider the partition of Q provided by Q(_2), we conclude that points
in G<-1)\Q (_2, = (G(0)n Q ,- 1>)\(Q(0>n Q <-1)r\Q (~2)) have images under g in
2
G<0l\(G <0,riQ <~ u). Therefore, if x e Q \ Q(~n>then gk(x )e F for k> 4. Thus, we
n=0
difficult to show that Q{~"' is invariant under g (see Exercise 3.5.3). Bearing
146 3 Structural stability, hyperbolicity and homoclinic points
in mind that p) Q{~n) is a set of straight line segments parallel to the y-axis and
iteN
that g involves a contraction along that direction, it is not surprising that points
in ( p | (2<-",)\A have orbits that approach A asymptotically (see Exercise 3.5.3).
neN
It must be emphasised that these orbits are not confined to a single vertical line
subset of p)
neN
Let us now turn to the dynamics of points in S2\Q'. Unlike f 1, g 1 is defined
for all x eQ (see Figure 3.18). Of course, g_ l |Q<l) = f~ ‘ i s as illustrated in Figure
3.15 but g~1( 6 \6 <1,) £ S 2\(2’ = (2'c. This means that points in Q\Qm are the
images under g of points lying outside Q'. We have already discussed the fate of
such images under forward iterations of g. Since A n (8 \Q (1)) = 0 , points in
(6\Q <1,)r i (P) Q( n>) have orbits approaching A asymptotically; while those in
Q
3.5 Horseshoe diffeomorphisms 147
we are able to extend the invariant set p | Q{ ") onto the whole sphere. The
neN
resulting set of points is called the inset, in(A) of A, i.e.
Exercise 3.5.3). It follows that there is a set of points, out(A) £ G u Q u F, the outset
of A, whose orbits approach A under reverse iterations of g, i.e.
The role of the set A in the dynamics of g is clearly analogous to that of a saddle
point in simpler diffeomorphisms. The inset and outset of A generalise the stable
and unstable manifolds of the saddle. We will return to sets possessing this more
general hyperbolic structure in §3.6 but now we must consider the dynamics of g
on A.
Proposition 3.5.3 The left shift a: E -* E has periodic orbits of all periods as well
as aperiodic orbits.
a V ) = o*, (3-5.8)
q e Z +. If q is the least, positive integer for which (3.5.8) is satisfied then a* is said
to be of period q. It is not difficult to see that (3.5.8) will be satisfied if and only
if a* = <7*+,, for all ne Z. It is then easy to find periodic points of a with any given
period, q. The required sequence, cr*, is generated by repetition of a block of
symbols of length q that is itself not composed of repetitions of any of its sub-blocks.
For example, the point
satisfies a '4(er*) = o* but has period-7 because a 1(o*) = o*, also. It is eqvally
straightforward to show that a has aperiodic orbits. For instance,
■ 4 3 2 1
which contains symbol blocks of the type shown for all n e l * , is such that there
is no q e l* such that a, (<r) = cr.
Proposition 3.5.4 There is a topology in which the periodic points of a are dense in E.
There is a natural way of defining how close two symbol sequences are to one
another. Given two sequences in £, we can obtain the length of the largest symbol
block, centred on the binary point, on which they agree. The larger the size of this
block the closer the two sequences are deemed to be. We are then able to define
the limit of a sequence of elements in £. A sequence {ai",|}“=0 S E is said to tend
to <re£ as m -*oo, if, given Ne1*, there exists Mel* such that o("> =o „for
—(N —1) < n < N, when m > M . Clearly, if <r,",>-* o as m -* oo then o{m) and a
agree on increasingly large central blocks. For example, the sequence (r<m,defined by
- ( m - 1) s $ n iS m ,m e Z +
rr<">= (3.5.12)
otherwise,
To justify Proposition 3.5.5 we must show that a has an orbit on I that approaches
every point of £ arbitrarily closely. Let c e E be such that <r_„ for ne № is given by
3.5 Horseshoe diffeomorphisms 149
All possible blocks of all lengths are included in{o _ „} „;o„, n e Z +, can be chosen
arbitrarily. The orbit of o under a contains {ccm(cr)|me N}. Now, by construction
<t contains any given symbol block of length N in its left hand half. After sufficiently
many applications of a this block will be centrally placed about the binary point.
Since N is arbitrary, any element of I can be approximated arbitrarily closely by
some point on the orbit of a under a.
In view of the rather special construction used above to obtain a sequence o
whose orbit under a is dense in E, the reader may feel that such sequences are in
some sense rare or atypical. This is not the case. In fact, most binary bi-infinite
sequences contain any prescribed block of symbols (see Hardy, 1979) and therefore
have a dense orbit under a. The particular example chosen above is carefully
ordered purely to make the argument more convincing.
Having established some properties of the left shift we must reveal
our motive for examining the dynamics of this map: namely to obtain a symbolic
description of the dynamics of the horseshoe diffeomorphism on A. Before doing
this, it is worth noting that the validity of Propositions 3.5.3-5 does not depend
upon the binary nature of the sequences in I . Similar results can be derived for
sequences of m-symbols, { 0 ,1 ,..., m - 1} say (see Exercise 3.5.5). Binary symbol
sequences allow us to deal with the horseshoe map of §3.5.1. However, there are
more sophisticated maps of this type (see Exercise 3.6.5) whose ‘symbolic dynamics’
involve sequences of m symbols with m > 2.
Q '^ e g H P o W i F , ) (3.5.15)
with g2(P0) n g 2(Pi) = 0 (see Figure 3.14(a)). In general, for n e l +,
QM ^ ( P 0) ^ ( P i ) (3-5.16)
and g"(P0) n g"(P,) is always empty because P0n P , = 0 and g is a diffeo-
morphism. Thus a horizontal strip of Q(n) lies either in g"(P0) or g’tPj). We allocate
the symbol 0 to a strip of Q(n) if it is a subset of g"(P0) and the symbol 1 if it lies
in g"(Pi) (see Figure 3.20). Obviously, these symbols alone do not provide a unique
description of each horizontal strip in Qw for n > 2, however, they can be used
to obtain one. For example, two strips of Q(21 have been allocated the symbol 0
but they are distinguished by the fact that one lies in g(P0) (i.e. strip 0 of Q(1))
and the other lies in g(P,) (i.e. strip 1 of Q(1)). Hence the strips in Q,2> can be
e (2)a
n QM ■
g - V O n P o n g l P O n g V , ) : rf2, = { 1 0 1 1 }
3.5 Horseshoe diffeomorphisms 151
uniquely labelled by giving two symbols: the first specifying a strip in Q,u so that
the second uniquely determines a strip in Qt2) (see Figure 3.20). Similarly, the
strips of Q(3) can be uniquely labelled by starting from the unique labelling of the
strips in Q(2) and appending the symbols allocated to Q(3). It follows that the strips
of Qw are uniquely specified by a set of n of the symbols (0,1}. Similar arguments
can be carried through for Ql~n) by considering the images of P0 and P, under
powers of g“ 1 (see Exercise 3.5.8). A vertical strip of Q{~"\ ne№, is allocated the
symbol i if it is a subset of g “"(Pf), i = 0,1. For ne Z +, unique labels for the strips
of Q,_">are obtained by appending these allocated symbols to those of the strips
in (see Figure 3.21). Notice we have appended symbols on the left so
that the order in the strip label matches that of the negative integers. Finally, we
Figure 3.20 Coding of strips in Q(") for (a) n = 1; (b) n = 2; (c) n = 3. The
symbol allocated to each strip is shown on the left and the unique code
for the strip is given on the right.
0 «- g(P0)
1 «- g(P.)
-0 1 g(Po)r>g2(P,)
0^ •00 g(Po) n g2(P„)
0 ns .10 g(Pi)r>g\P0)
\' 11 g(fl) r , g 2(P,)
(b)
can construct the symbol blocks representing the squares occurring in A(iy). Each
such square is the intersection of one of the vertical strips of (2t-(iV~ 1)) with one
of the horizontal strips of Q(m. If the vertical strip has label o _ (N_ „ , . . . , <r_ u o0
and the horizontal strip has label ol, . . . , o N, the symbol block representing the
square is taken to be <r(iV) = { a - ^ ..,) ,.. .,<r_„ <T0-<r,,. . a*,}. Thus, for example
{11.01} and {10.11}, respectively, represent the top right hand and bottom left
hand squares in the illustration of A(2) given in Figure 3.19(h). It is not difficult
to show (see Exercise 3.5.11) that the square represented by the symbol block o{N)
Figure 3.21 Coding for strips of Q1'" ’ for n = 0,1, 2: (a) Q<01; (b) 5 ,_1);
(c) Q' 2). Unique labels for the strips are given above and allocated
symbols below. Notice that, to match the negative integers, symbols are
appended to the left rather than to the right.
g''(ft) A f t g~'(ft) H ft
(ft)<~>ft \ / g -'(ft)r« ft
ft ft
\10 \00 01/ 1/1
0 1
(o) (b)
1001 1001
3.5 Horseshoe diffeomorphisms 153
is given by
( H g ^ jW (3-5.17)
\n= (JV-l) 1
In the limit N -* oo, the above construction assigns a unique, bi-infinite binary
sequence with each point of A. Moreover, (3.5.17) allows any such sequence to
be converted to a unique point of A. We have therefore constructed a bijection
h:£-A.
Proof. The nested nature of vertical and horizontal strips defining A means that
sequences that are close, in the sense that they agree over large central blocks,
map under h to points of A that are geometrically close together. Similarly if two
points of A are geometrically close, the symbol sequences agree over a large central
block because it is only for N sufficiently large that such points are distinguished
in A(iV,. Thus h is a homeomorphism.
Let <re£ and
x g" ( P J (3.5.18)
Then
—
Co
g(h(ff)) = g(x) = g (| H g"(P„,
_neZ
nQ ,
nQ ,
D g " +1( ^
_«eZ
Og r>Q,
.neZ
= H«(<r)). (3.5.19)
invariant sets like A are often referred to as chaotic sets (see § 3.6) because of the
presence of such orbits.
While the dynamics of g|A is very complicated, we must not forget that the
dynamics of g|Ac suggests that A is, in some sense, hyperbolic. In the following
section, we consider how such sets fit into a general theoretical framework.
Let us return to the hyperbolic nature of the invariant set A of the horseshoe
diffeomorphism g: S2 -*■S2. In fact, A is said to have a hyperbolic structure or to
be a hyperbolic set for g. Our aim in this section is to explain this statement and
to introduce an important theorem about diffeomorphisms whose non-wandering
set, ft, has a hyperbolic structure.
It is helpful to review our previous encounters with hyperbolicity (see §2.1 and
2.2). The striking feature is that, thus far, we have only had to consider hyperbolic
fixed points. Non-trivial hyperbolic sets such as a hyperbolic periodic orbit, or a
normally hyperbolic invariant circle, are defined in terms of a hyperbolic fixed
point of a related map (P or §■ in §2.2). We are then involved with the local
behaviour of a map at a fixed point in a Euclidean or Banach space. In such cases,
the hyperbolic nature of the fixed point is given in terms of the eigenvalues of the
derivative map (DP or OS' ). It is not possible to use this approach to characterise
the hyperbolicity of the invariant set, A, of the horseshoe diffeomorphism. However,
it is useful to consider why such an approach fails. There are two problem areas.
(i) The horseshoe diffeomorphism is defined on a manifold (the sphere) and not
a Euclidean space. This means that the generalisation of the derivative map to
this situation must be considered.
(ii) The complexity of A is such that it is not possible to formulate the problem
in terms of a fixed point of some related map. For example, A contains aperiodic
orbits which do not correspond to a fixed point of g1, for any q 6 Z +. Thus, having
introduced the appropriate generalisation of the derivative map, our definition of
hyperbolicity must allow for the fact that x and g(x) are different points in A.
Let us begin by considering how the results of §2.2 can be applied to a
diffeomorphism f:M - * M when M is an n-dimensional, differentiable manifold
that is not a subset of IR" (see Figure 3.22). The derivative map, D f(x * ): IR" -> IR",
used to discuss the hyperbolic fixed point, x*, of f: IR" -* R" in §2.2, is replaced
by the tangent map T fx.: TMX>-* TMX., where TAi„. is the tangent space to M
at x * . Recall (see §1.3) that TMX can be defined, for any x e M , in terms of
equivalence classes of curves on M with the same tangent vector at x . To see the
connection with the behaviour of f near x * , let ij((), with t e / s I R , O e / and
if(0) = x*, be a parametrised curve on M passing through x*. To find the tangent
vector at x*, we need to differentiate ij(l) with respect to t and (see § 1.1) this can
only be done by using a local chart, (Ux, h .) say, containing x*. The local
representatives, f, ij and C q, of f, tj and f-if, respectively, in (U a, h j (or, more
3.6 Hyperbolic structure and basic sets 155
at t = 0. The vectors (f„* ?«)(()) and q„(0) are a-representatives of elements of TM„.
in ((/„, h j. Strictly speaking, they lie in the tangent space to Ux at x* = qa(0) but,
as TUj* is a replica of R", this distinction is not always apparent. The derivative
map is the local representative of the tangent map Tf,.. As the opening
remarks to § 2.2 suggest, x* e M is said to be a hyperbolic fixed point of
if x* is a hyperbolic fixed point of f4 in the sense of Definition 2.2.1, i.e. if Dfa(x*)
has no eigenvalue with unit modulus. What is more, if we assign a metric to TV
then (see Exercise 2.1.2) hyperbolicity of x* corresponds to imposing bounds on
|Df4(xJ)"v| for all v in the stable and unstable eigenspaces of Df„(xJ). Of course,
it is only by making such an assignment that we can define a metric on TMX..
Indeed, if x* lies in the overlap of two charts (Ua, ha) and (Up, h^), then we are
only allowed to choose metrics on T l/X -. and T t/X-. that, for all v e TMX>, satisfy
I'ali = where va (v^) is the a- (/^-representative of v (see Figure 3.23). The
common value defines ||v||x. for any v in TMX>. Compatible metrics, such that
||v||x is positive definite, defined at all points x of all overlaps of an atlas provides
a Riemannian structure for M (see Exercise 3.6.2). If M is equipped with a
Riemannian structure then we can express the hyperbolic nature of x* in a
coordinate-free way by requiring that:
(i) TMX. = £ x. © £ “., where E jf’ is the stable (unstable) eigenspace of Tfx.;
(ii) there exist c, C > 0 and 0 < p < 1 such that, for every n e Z +,
generalise (3.6.1>-(3.6.5) as follows. Let x, and f(x,) belong to charts (U„ h„) and
(Up, h^), respectively, then
(f.*) = h,.(f.<,) (3.6.6)
with
(f^ K O -W W O ). (3.6.7)
Item (iii) is trivially satisfied if A is a periodic orbit, since the points x eA are
isolated. However, it is an important technical restriction for invariant sets
containing a dense orbit or a dense subset of periodic orbits.
It is not difficult to accept that the invariant set A = f) QM is a hyperbolic set
for the horseshoe diffeomorphism g. Observe (see §3.5.1) that the set of vertical
line segments, f ) Q' "1, on the square Q give rise to curves on S2 analogous to
neN
the stable manifold of a periodic orbit. Similarly, the horizontal line segments,
P| Qm, lead to the analogue of the unstable manifold of the periodic orbit. At
neZ *
each point x eA we can identify tangents to these curves to obtain Ex and £ “.
Moreover, this splitting into El and £ x depends continuously on x, because, for
158 3 Structura! stability, hyperbolicity and homoclinic points
any x, x 'eA , ££ and £*., (or £ “ and £ “.) are tangent to diffeomorphic images of
parallel line segments on the square Q. Finally, the contraction on E?x and expansion
on £J satisfy (3.6.9) and (3.6.10) with p>% and c = C = 1, so that A has a hyperbolic
structure.
In §1.4, we noted that fixed points and periodic orbits are invariant sets that
frequently appear to attract or repel the orbits of points not contained in them.
What is more, they are rather special in so far as they have no proper subsets that
are themselves invariant. The following theorem for diffeomorphisms whose
non-wandering set, ft, has a hyperbolic structure, provides the theoretical basis
for these observations.
Each ftj is closed, invariant and contains a dense orbit of I. Moreover, the splitting
of ft into basic sets is unique and M can be decomposed as a disjoint union
M = [) in(ftf), (3.6.12)
i=l
where
in(ft,)= {xeA f|f"(x)-»fti,m -* oo} (3.6.13)
Diffeomorphisms with hyperbolic non-wandering set, ft, and periodic orbits dense
in ft are usually referred to as axiom-A diffeomorphisms (see Chillingworth, 1976,
p. 240; Nitecki, 1971, p. 189). Clearly, any diffeomorphism whose non-wandering
set consists of a finite number of fixed points or periodic orbits is axiom-A.
Moreover, fixed points and periodic orbits are closed, invariant sets that trivially
contain a dense orbit, i.e. they are basic sets.
Theorem 3.6.1 does not merely give a decomposition of ft. Equation (3.6.12)
states that every xeJVi belongs to the inset (or equivalently, the outset) of one
and only one basic set. This means that the wandering points move between the
basic sets approaching those that are attracting asymptotically. Some simple
examples are illustrated in Figure 3.24.
The horseshoe diffeomorphism, g, on the sphere is a more substantial example.
The non-wandering set, ft, of this diffeomorphism consists of the invariant set
A= 0 and the two fixed points; one stable and one unstable, ft has a
neZ
hyperbolic structure and Proposition 3.5.4 shows that the periodic points of g are
dense in ft, so Theorem 3.6.1 applies. Obviously, each fixed point is a basic set
3.6 Hyperbolic structure and basic sets 159
(ft, unstable; ft2 stable, say) but can A be decomposed into a number of basic
sets? Propositions 3.5.5 and 3.5.6 show that A contains a dense orbit of g. This
means that further decomposition of A is out of the question and, since it is also
closed (it is a Cantor set) and invariant, the only remaining basic set (ft3) is A
itself. Basic sets of this type are referred to as chaotic sets (see §3.5.3 and Exercise
3.5.11).
The Anosov automorphisms provide another illustration of Theorem 3.6.1
involving a chaotic basic set. Recall that the periodic points of these maps are
dense in T" and the non-wandering set is the whole torus. Moreover, in the
and E“, are given by (1, (1 —5l/2)/2) and (1, (1 + 51/2)/2), respectively, at every
periodic point x, (see Exercise 3.4.2). Continuity requires that this be so for each
x e ft because the periodic points are dense. The splitting of the tangent space is
Figure 3.24 Illustrations of Theorem 3.6.1 where the basic sets are fixed
points and periodic orbits, (a) f: Sz S2 has non-wandering set, ft,
consisting of two basic sets, ft, and ft2 - both fixed points. All wandering
points have a-limit set ft, and co-limit set ft2. (b) The basic sets ft, and
ft2 are unstable fixed points, ft3 is a saddle-like 4-cycle and ft4 is a stable
4-cycle. The dynamics of the wandering points are shown schematically
on the right, (c) The basic sets are all fixed points in this case: ft, is
unstable; ft2 is stable and ft3, ft4 are saddle-like. Once again a schematic
representation of the dynamics of the wandering points is given.
160 3 Structural stability, hyperbolicity and homoclinic points
therefore trivially continuous, being the same at every point of ft. Hyperbolic rates
of contraction and expansion follow from the hyperbolicity of A (see Exercise
3.6.4). Hence ft has a hyperbolic structure. In this case, there is only a single basic
set ft, = £1= T 2 and it follows from Theorem 3.6.1 that the toral automorphism
must have an orbit which is dense in the torus.
A further example is the transformation, f, of the solid torus, T = S* x D2, shown
in Figure 3.25. The torus is treated rather like a solid rubber ring. It is stretched
(with consequent loss of cross-sectional area), twisted and folded to fit inside itself.
Repeated application of this transformation results in longer and longer tori,
wrapped around T increasingly many times. If the disc D2 is a cross-section of T,
then D2n is a Cantor set. Thus, the «-limit set of f is locally the
product of a Cantor set and a one-manifold. This example has the important
property that the chaotic basic set is an attractor.
The set A = Q Qin) in the horseshoe diffeomorphism has only a one-dimensional
inset. This means that most orbits are not asymptotic to A and this makes A
Figure 3.25 (a) Illustration of a transformation f of the solid torus T
which has an attracting chaotic set. The image, f(T), of T under f is
shown shaded, (b) Intersections of successive images of T under f with
a cross section D2 of the torus. Notice that f- 1 is not defined on the
whole torus, however, only forward iterations are required to observe
the attracting set. The mapping f is sometimes called the ‘spinning
diffeomorphism’.
(b )
3.6 Hyperbolic structure and basic sets 161
Figure 3.26 The Duffing attractor (see Guckenheimer & Holmes, 1983,
pp. 82-91 & 191-3). The Dulling equation can be written in the form
x = y, y = x —x3 + f.(a cos 0 —by), 0=1.
This system is periodic in 0 and the phase space can be taken as
M = R2 x S'. Every surface 0 = constant is a global Poincare section
so that the system behaviour is completely described by the Poincare map
PIf#. Numerical approximations to Pt 9appear to have a chaotic attracting
set - the Euler approximation is shown in this diagram for ea = 0.4 and
eb = 0.25. The structure of Pt i is discussed in greater detail in §3.8.
f f i ' " “ :V \
, / /
f.'
V- ï
V.:
*; ’ .. \y ‘
162 3 Structural stability, hyperbolicity and homoc/inic points
Figure 3.27 The Henon attractor (see Henon, 1976). The map f producing
this attractor is defined by
(x, y)^*(y~“x2 + 1, bx),
where a, beR. The attracting set for a = 1.4 and 6 = 0.3 is shown. It
arises from the repeated folding and stretching brought about by the
action of f. When magnified the attractor is found to consist of many
curves, of similar shape to those resolved above, occurring very close
together. This ‘braided’ nature of the attractor appears to be repeated
on all scales.
y
3.6 Hyperbolic structure and basic sets 163
Figure 3.29 The Lorenz attractor (Lorenz, 1963). It must be pointed out
that there is a theorem corresponding to Theorem 3.6.1 for flows so that
strange attracting sets can also arise in flows that are not the suspension
of a diffeomorphism. The Lorenz equations
x= 10(y~x), y = x(28 - z) - y, z = xy —(8/3)z,
have fixed points at (±6(2I/2), ±6(21/2), 27). The system does not have
a global section so the projection onto the xz-plane is shown. The orbit
generated by using the Euler method with step length of 0.005 and initial
point (x, y, z) = (0.1,0,0) is plotted. The projected orbit switches between
revolving about (x, z) = ( + 6(2I/2), 27) and (x, z) = ( ~6(21,2)l 27) in an
apparently random way.
z
X
- 6/2 6/2
that are not resolved on the scale of Figure 3.27. Further magnification shows that
each of the latter ‘curves’ has a similar structure, and so on. The attractor is said
to have a ‘braided’ nature which is repeated on all scales. The reader will recall
that the chaotic basic set A of the horseshoe diffeomorphism has this property
(see Exercise 3.5.9). In fact, a theoretical connection with the basic set of the
horseshoe diffeomorphism can be made in some cases: namely when homoclinic
points occur.
fw(x{)
166 3 Structural stability, hyperbolicity and homoclinic points
the roles of Wl and Wu are reversed and, for some N ' e l * , = R0, where
R, and R0 intersect as shown in Figure 3.32. Clearly, if p = N + N ', f P{R0) = R t
and we would expect P to exhibit horseshoe-like behaviour, i.e. be conjugate to
a left shift on two-symbols. The homoclinic point referred to in Theorem 3.7.1
would in this case be x+ = f " r (xi).
Theorem 3.7.1 means that f exhibits all the complexity of the left shift or. X -►X
discussed in §3.5.2. In particular, in every neighbourhood of a transverse
homoclinic point of f, there is a periodic point. By Theorem 3.7.1, the transverse
homoclinic point x* e A and PI A is topologically conjugate to the left shift a: X -►X.
However, by Proposition 3.5.4 the periodic points of a are dense in X. Hence,
periodic points of P |A are dense in A and, therefore, there is a periodic point
of f arbitrarily close to xf. Thus there are infinitely many periodic points in any
neighbourhood of xf.
It is important to realise that Theorem 3.7.1 employs sufficient conditions to
ensure the existence of A. As Smale has pointed out (see Smale, 1963), we might
expect a similar result to hold with weaker constraints on f. Figure 3.32 suggests
that the key requirement is that the stable and unstable manifolds of a hyperbolic
fixed or periodic point intersect transversely. With this in mind, the following
example shows that the remarkable phenomena described above really do occur.
Let us examine the planar map
* i= x + y„ y, = y + k x ( x - 1), (3.7.1)
numerically, for 0 < k < 4. This map has fixed points at (x, y) = (0,0) and (1,0) for
all values of k. The fixed point at (0,0) is non-hyperbolic. The linear approximation
to (3.7.1) at (0,0) is conjugate to an anticlockwise rotation through angle 0, where
Linearisation at (1,0) shows that this fixed point is a hyperbolic saddle point with
£(i,o) and E"1-0) given by
u = m{ —fe—[fc(4 + k)]1/2}/2 and t>= u{-fe + [/c(4 + k)]1/2}/2, (3.7.3)
respectively, where (m, i>) are local coordinates at (1,0). It is not difficult to then
use a microcomputer to plot successive images of, say, one hundred points lying
in a small interval of Ej\t0, close to (1,0). The result of such a calculation can be
quite spectacular (see Figure 3.33(a)). Of course, a suitable interval along ^U.O)
can be iterated, using the inverse map
(*)
168 3 Structural stability, hyperbolicity and homoclinic points
Figure 3.34 (a) Plot of the stable and unstable manifolds of the saddle
point at (1,0) of (3.7.1). Since the map is orientation-preserving, the
image of the manifold loop S0 must be one of loops St, i = 1 ,2 ,..., with
the same orientation as S0. It is not difficult to see that the orientation
of the loops S; is opposite to that of S0. In fact, for (3.7.1), the image of
S, is S,+, (see Figure 3.34(c)) but this is not the case in general. For
example, the image of Sf under the square of (3.7.1) is Si+2. (b) The
images of S0 under iterates of the inverse of (3.7.1) are the regions S_,
which wrap further around the fixed point at (0,0) as i increases, (c)
Numerical plot of the orbit of the point P = (0.64, —0.094) under (3.7.1).
It sweeps around (0,0) twice, passing near to the saddle point on each
occasion, before arriving in S0 at the fifteenth iteration. Subsequent
iterates are carried away to infinity under the influence of the saddle
point. Note that, since the manifold loops become extremely narrow and
close together, the number of revolutions of the orbit about (0,0) before
expulsion to infinity can depend sensitively on the choice of initial point.
(c)
3.7 Homoclinic points 169
S(, i = l , 2 , . . . , with the same orientation as S0, and not one of Sf, i = 0 ,1 ,2 ,..
for which the orientation is reversed. Thus points in S0 are ultimately swept off
to infinity under the influence of the saddle point at (1,0). Similarly, points in S0
are swept around the fixed point at (0,0) and fed back into the vicinity of the
saddle point once again. The role of this movement about (0,0) in the dynamics
of (3.7.1) is best understood by considering images of S0 under powers of the
inverse map. The pre-images of S0 are a subset of the regions S_ i= 1 , 2 ,. .. ,
shown in Figure 3.34(b). Observe that, as i increases, these regions stretch further
around (0,0). Indeed, for each N e Z +, there is an i(N) such that S_i(iV) wraps
around (0,0) N times. It follows that there are points in S_,(W) whose orbit makes
N trips around (0,0) before it appears in S0 and subsequently sweeps out to
infinity. It is not difficult to confirm these ideas numerically. An orbit exhibiting
this behaviour is shown in Figure 3.34(c).
Similar orbits were shown in Figure 1.39 and 1.40 for the Henon area-preserving
map. This is no coincidence. Henon has shown (see Henon, 1969) that every
quadratic, area-preserving, planar map, with rotational linear part at the origin,
is conjugate to the form (1.9.40). It is easily verified that the derivative of the map
(3.7.1) has unit determinant for all (x, y )e R 2 (see Exercise 3.7.5)). Thus (3.7.1)
and (1.9.40) must exhibit the same dynamics. For our present purpose, (3.7.1) has
the advantage that the saddle point remains at (1,0) for all k, so that 0) and
£(] 0) are easily calculated.
In the above discussion, we have assumed that the stable and unstable manifolds
that intersect one another come from a single fixed point x*. Recall that Theorem
3.7.1 includes the case where the stable and unstable manifolds involved are
associated with a fixed point of f*. Similarly, if x* is a periodic point of period
greater than one, then, for example, the unstable manifold of x* may intersect
transversely with the stable manifold of f(x*) (see Figure 3.35(a)). Once again, the
manifolds oscillate wildly because images of homoclinic points are homoclinic
points. Given that the unstable manifold of f(x*) also intersects the stable manifold
of x* transversely, then consideration of the images under f of a suitable
parallelogram, R, again indicates that some power of f behaves like a horseshoe
map (see Figure 3.35(h)).
This construction is also relevant to quadratic, area-preserving maps of the
plane. Suppose x* has period-^ and homoclinic points arise in the manner described
above at each point of the periodic orbit, i.e. in the above argument x* t-* f<m“ 1,(x*)
and f(x*)H f"'(x*), m' = m mod q, for m = 1........ q. Then we obtain a chain of
homoclinic tangles as shown in Figure 3.36. In this case, the orbit of a point such
as P in this figure could sweep around the whole periodic orbit before being fed
back into the vicinity of x* at a different point, F . Because of the massive stretching
along the unstable manifold at each periodic point, the position of F depends
sensitively on that of P.
There is evidence of this kind of behaviour in the maps (1.9.40) and (3.7.1). The
‘two-dimensional’ orbits shown in Figures 1.41 and 1.42 are associated with a
hyperbolic periodic orbit, they are generated by iterating a single point and their
170 3 Structural stability , hyperbolicity and homoclinic points
extent is similar to that of the expected homoclinic tangles (see Gumowski & Mira,
1980, p. 303). In this situation, there is a good reason (see Figure 6.17) why orbits
of this kind do not escape from the influence of the periodic orbit. Therefore, the
plotted iterates of a single point appear to Till out the two-dimensional region in
an apparently random way.
Figure 3.35 (a) Illustration of the unstable manifold of the periodic point
x* intersecting the stable manifold of f(x*) transversely at x( and hence
at infinitely many other homoclinic points. (i>) The parallelogram R is
iterated forward to R, and in reverse to R2. The map from R2 to R, is
horseshoe-like.
3 j8 The Melnikov Junction 171
* = f0(x) (3.8.1)
Figure 3.37 Phase portrait for x = f0(x). The origin is a hyperbolic saddle
point and r is a homoclinic saddle connection.
172 3 Structural stability, hyperbolicity and homoclinic points
with e e R + and f,(x, 6) = f,(x, 6 + In). For sufficiently small e, it follows from
Proposition 3.2.2 that (3.8.3) also has a hyperbolic periodic orbit, yt, close to y0.
However, the invariant manifolds, 1F“(ye) and IP fyJ, need not intersect to form
a cylinder (see Figure 3.38). The Melnikov function is related to the ‘distance’
between these two manifolds.
Let x0 e R2 be a point of the saddle connection T in the unperturbed system
(3.8.1). Take a perpendicular section L to the saddle connection at x0. We use
the point x0 and the section L in the 6 = 0o-plane, I #o, as follows. Consider the
perturbed system and the intersections of yt, Wu(yt) and WTyJ with This is
equivalent to studying the Poincare map Pe #o: I #o -* Zfo of the flow (3.8.3). P£„o
will have a hyperbolic saddle point, x*-9o, near to x = 0, with stable and unstable
manifolds, H/ “'5(x*flo) = W “'s(y£) n I #o, which are close to T on (see Figure 3.39).
The distance between lF u(yt) and W s(yt) on Z#o is calculated along L. Observe
that this distance will, in general, change with 0O, since e > 0 implies that the
curves ^ " ( x j , ) and IF^x*^) will be 0o-dependent. Obviously, for the special case
e = 0, the distance would be zero for all values of 60.
Of course, the manifolds 1F"'5(x*9o) may intersect L many times, however, on
Figure 3.38 The manifolds W^y,.)and W'(yt) for (a) r. = 0 and (b) e > 0.
3.8 The Melnikov Junction 173
each curve there will be a unique point of intersection A*’*, closest to x0 (see Figure
3.39). Let (x“-s(i; 0O, e), t), f eR , be the unique trajectory of (3.8.3) passing through
/!“•’ at t = 0o, i.e. A“-’ is the point xu'!(0o; ®o> £)6 ^«0- We then define the
time-dependent distance function,
Ae(t, 0O) = f0(x0(t - e0)) A [x“(i; 0O, e) - x5(t; 0O, e)], (3.8.4)
where x0(f) is the homoclinic trajectory of (3.8.1) with xo(0) = xo. In (3.8.4) the
wedge product is defined by a a b = aib2 —a2bl where a ,b e R 2 have Cartesian
coordinates (al( a2) and (bt, b2), respectively. It follows that A£(t, 0O) is |f0(x0(t —0O))|
times the component of the vector [x“(i;0 o, e) —x!(f;0o, e)] perpendicular to
f0(x0(t - 0O))- The latter vector is, of course, tangent to T at x0(i - 0„). Thus,
At(0o, 0o)/|fo(xo)l is the distance between M/U(y£) and H/,(yE) measured along L on
V
We can obtain a useful form for A£(0O,0 O) by studying (3.8.4) more closely. Let
and
where x“, x‘, are first variations with respect to e. Thus, (see Exercise 3.8.1)
Figure 3.39 The intersections of y£, lT“(y£) and VT!(y£) with I 0o for e = 0
and e > 0.
A V \xU )
W“(%)rMe„
W 'W n Z t.
IV'IxU)
174 5 Structural stability, hyperbolicity and homoclinic points
We can obtain differential equations for A“ and A'. It can be shown that, since
* o (i-0 o ) = M xo( f - 0 o)).
A i(t, 0„) = e[Tr(Df0(xo(t - 6»o)))f0(x0(t - 0O» a x\(t, 60)
Here we have noted that A“(—oo, 0O) = 0 because x0(—oo) = 0 = f0(0). A similar
calculation leads to
and therefore,
so that
Proposition 3.8.1 I f M (0O) has simple zeroes, then, for sufficiently small e > 0,
W/u(x*So) and H/ *(x*6o) intersect transversely for some 0O6 [ 0 ,2n). On the other
hand, if M(ff0) is bounded away from zero, then M/U(x*9())r'i VT’fx*^) = 0 for all 0„.
x0. Of course, all the Poincare maps P£9o, 80 e [ 0 , 2n), are topologically conjugate
(see Exercise (1.7.3)) and, consequently, !Pu(xJSo) and lP s(x*-#o) must intersect
transversely for all 0„ 6 [0, In) (although, obviously, not always near to x0 (see
Figure 3.40). Equally, if M(0O) is bounded away from zero, then, for sufficiently
small e, so is At(0o, 0O). This, in turn, means that transverse homoclinic points do
not occur on L for any 0oe[O, 2ti). As Figure 3.40 shows, this conclusion does
not depend on the choice of x „ e r through which L passes. Hence, there are no
homoclinic points.
Example 3.8.1 Show that the Poincare map of the Duffing equation
a,b> 0, has transverse homoclinic points, for sufficiently small values of e, provided
a 4 cosh(?i/2)
(3.8.18)
b > 3(21/2)ti
so that f0(x) = (y, x —x3)T. The differential equation x = f„(x) has a hyperbolic
saddle point at x = 0 and two further fixed points at x = ( + l ,0 ) T. It is a
Hamiltonian system with
tf(x,y) = ^ y 2 - x 2 + y j (3.8.20)
and the level set of H{x, y) = 0 consists of two homoclinic orbits, T f , and the saddle
point at x = 0 (see Figure 3.41). It can be shown (see Exercise 3.8.2) that the
trajectories passing through (x, y) = (+ (2li2), 0) at t = 0 are given by
Figure 3.41 The phase portrait for the planar system x = f0(x), f0(x) =
(y, x - x’)T. Stable and unstable manifolds of saddle point x = 0 coincide
to form a pair of homoclinic orbits Fg. The level set H(x, y) = 0 is
r o- u { 0 } u r o+.
5.8 The Melnikov Junction 177
which satisfies f, (x, 0) = f,(x, 0 + 2k). It follows (from (3.8.15)) that the Melnikov
function for the homoclinic orbit To is
4b
M(0o) = 21/2asin(0o) I sech(i) tanh(t) sin(i) di —y . (3.8.26)
■r
The integral occurring in (3.8.26) can be evaluated using the method of residues
(see Exercise 3.8.4)) and we finally obtain
Clearly, if (3.8.18) is satisfied M(0O) has simple zeroes and, by Proposition 3.8.1,
transverse homoclinic points must occur. On the other hand, if the reverse
inequality is satisfied, M(0O) is bounded away from zero and Proposition 3.8.1
implies that there are no homoclinic points. □
There is one remaining possibility for the system (3.8.17): namely that
4 cosh(?i/2)
3(21,z)
In this case, M(0O) has a double zero at 0O= 3n/2. This corresponds to lP“(x*i3lt/2)
and W',(x*3l/2) meeting tangentially rather than transversely. As before, the orbit
of such a homoclinic point under Pc,3„/2 consists entirely of tangential intersections
178 5 Structural stability, hyperbolicity and homoclinic points
y
Figure 3.43 (After Ueda, in Guckenheimer & Holmes, 1983, p. 90.) Comparison of the shape of the unstable manifold of the saddle point
and the attracting set for the Poincare map of the Duffing equation (3.8.17): when (a) ea = 0.40, eb = 0.25; (b) ea = 0.30, eb = 0.20.
- 1.0 x - 1.0
180 3 Structural stability, hyperbolicity and homoclinic points
Exercises
3.1 Structural stability of linear systems
3.1.1 Consider a real, n x n matrix, A, with eigenvalues A,,. .. , l„ that are not necessarily
distinct. Let B, with eigenvalues /i,, ■■.,p„ be e-close to A in L(R"). The spectral
variation of B with respect to A is defined by
Sa(B) = max [min (|A, —gj)]. (E3.1)
1 i
(a) Assume that A can be diagonalised and show that
S*(B)<e.
(b) Suppose A cannot be diagonalised and show that
SA(B) <(«)**■
provided e*< 1/n.
(c) Deduce that {pt, .. .,/<„} -* {A,,. .. , 2„} as £ -►0 for any AeL(R").
3.1.2 Let SD(R") be the subset of structurally stable linear diffeomorphisms in L(R").
Show that a linear diffeomorphism is structurally stable if and only if it is
hyperbolic. Hence, or otherwise, show that SD(R") is open and dense in L(R").
Exercises 181
1.3 Let S be the subspace of L(R") defined by ■ l yt 0 or 1 >. Show that every
{G 3 h - 4
linear diffeomorphism in S is structurally stable within S but not within L(R2).
1.4 Consider the subspace, 0 (R 2), of L{R2) defined by {A|ATA = I, A eL (R 2)}. Show
that no element of 0(R 2) is structurally stable in L(R2).
3.3.3 Consider the system on the cylinder C = {(0, r)|O ^0<2n, reR}, given by
has a saddle point at «(0), where n: R2 -> T2 is the map given in (3.4.2). Find the
equation of the séparatrices of A at 0 and show that they have irrational slope.
What are the implications of this on the torus? Show that the point on the torus
given by «(x*), where = Anosov auto(131,2~ l)/2(13li2), 1/131'2), is a transverse
homoclinic point. How does the Anosov automorphism considered in this question
differ from that given by (3.4.12)?
3.5.4 Suppose g: Q R2 is the horseshoe map and take x0, with coordinates (x0, y0),
to be a point of Q\A. Let d = d(x0, A) = min {|x0 —x|} be the horizontal distance
(Jc.y)eA
from x0 to A. Find the maximum value, N(d), of n such that f"(x0)eQ.
3.5.10 Use the form for xe A given in Exercise 3.5.9 to find the coordinates in Q of the
fixed and period-2 points of the horseshoe map f: Q R2.
3.5.11 Let f:Q -»R2 be the horseshoe map. In the notation of §3.5.3, the set
i*
A'*1= D Q("' consists of 22* connected components, each one being a square
»=-(N-1)
region of side 2/5* lying within Q.
(a) Prove that (3.5.17) uniquely associates a symbol block <r(l,,= ..., <r„}
with each connected component, k(ct,*)) say, of A1*1. Locate k(o-<w>) for the
following symbol blocks aim:
(i) {11}; (ii) {11-00}; (iii) {010-101}.
(b) Let i}tw, = .. .,!)„} and v,m = ..., v„} be two symbol blocks
of length 2N. Explain how to choose a point xeA in k(i/(**) such that
f2N(x)e ic(v<N>). Hence show that there exist points in jc(r/<w>) whose orbit under
f2N visits every connected component of A1*1in any desired order.
(c) What restriction must be imposed on the elements of the sequence a if the
orbit of the point x = h(<r) e A is to remain in a particular component,
say, for k applications of f? What is the maximum number of connected
components that can be reached from k(o‘<w’) in k iterations of f?
(d) What aspect of the dynamics of f|A do the observations (a)-(c) reflect?
3.5.12 Recall that the horseshoe map f: Q - * R2 satisfies ( x, y ) ( / , ( x ) , / 2(x , >’) ) . Verify
this property for f| A by considering the left shift a: X -»X that is conjugate to f.
Show that /,: [ —1,1] -» R has a repelling invariant Cantor set.
3.5.13 The Baker’s transformation B: T2 -> T2 is defined by
x ,= 2 x m o d l, y, = ^(2x-x, +y)mod 1
for (xmodl, ymodlje T2 (Arnold and Avez, 1968, Appendix 7).
Describe the effect of this transformation on the rectangles P0 = [0, j) x [0,1)
and = [j, 1) x [0,1). Show that every point x e T2can be written in the form
x = h(<r)= pj
It — — 00
where <s= „ is a bi-infinite sequence of {0,1}. Use this result to show that
3.6.5 Consider the diffeomorphisms g,, g2: S1 -> S2 defined in Figure E3.2. Outline
arguments to show that g,-, i = 1,2, has an invariant Cantor set A, £ Q such that
gilAj is conjugate to a shift on m(i>symbols, where m(l) = 3 and m(2) = 4. What
are the basic sets of g, and g2? For both maps, draw schematic diagrams illustrating
the dynamics of the wandering points. Is g, | At conjugate to g2|A2?
3.6.6 Find which of the following diffeomorphisms of the torus T 2 satisfy the
hypotheses of Theorem 3.6.1 and describe their basic sets:
/3 4
(a) f,(«(x)) = «(Ax), A = l 5
Figure E3.2 The restrictions of g! and g2 to the capped square Q' are
shown in (a) and (6), respectively. On each component of Qf0’, i = 1, 2,
g; is assumed to be linear. The map g, is a contracting diffeomorphism
on both F, and G, with hyperbolic fixed points P, e f j = gi(F,)
and f>2eG', = g1(G1). Observe that G2 = g2(G2) c F2 and g2|F 2 is a
contracting diffeomorphism with hyperbolic fixed point P3e F’t = g2(F2).
On S2\Q\ both g, and g2 have a single repelling hyperbolic fixed point
P0. It may be assumed that Theorem 3.6.1 applies to both g, and g2.
W gi
A Q D
(h) g2
Exercises 187
(b ) f2(*(x)) = j i ( x + b), b =
/ sin(27tx)'\
(c) f3(>t(x)) = n(<p,(x)), where <p, is the time-one map of the system
\sin(2xy)/
6.7 Obtain the Henon attractor using a microcomputer to plot the iterates of the map
x, = y - 1.4x2 + 1, yi=0.3x, (E3.9)
with initial value (1,0). Choose x and y scales such that the square Q =
{(x, y)||x|< 1, |y| < 1} fills a large portion of the screen. Observe the braided
nature of the attractor by magnification and shifts of the origin.
7.3 Find the eigenvalues of the linearisation of (3.7.1) at the fixed point (x, y) = (1,0)
and verify that both are positive for ic > 0. Show that the eigendirections are given
by y = {—k±[k(k + 4)],/2} (x- l)/2. For <c= 1.5 take an interval of approximate
length 0.0001 containing 100 points on the appropriate branch of the unstable
manifold of the saddle fixed point. Plot 15 iterates of each point under (3.7.1)
to obtain a numerical approximation to the unstable manifold at (1,0). Use the
inverse map (3.7.4) to complete the homoclinic tangle shown in Figure 3.33.
Modify the program to exhibit the image of each successive iteration of the interval
separately. Observe the repeated stretching and folding around the origin.
7.4 Use the program developed in Exercise 3.7.3 to study how the extent of the
homoclinic tangle depends on k. Plot the tangle for k = 0.4,0.8,1.2,1.6 and 2.0.
Comment on your results.
7.5 Show that the derivative map of (3.7.1) has unit determinant for all (x, y)eR2.
Given that there exists a linear conjugacy between (3.7.1) and (1.9.40), find a
relation between k and a. Modify the program used in Exercise 1.9.9 to generate
orbit plots for (3.7.1) corresponding to Figures 1.38 and 1.39.
7.6 Let y(t), t e l e R, define a closed curve, y, in the plane oriented with increasing
i. Suppose that r(s), s 6 J e R, defines a segment of a planar curve, T, that intersects
y transversely. Assume that the point of intersection is given by x0 = y(0) = T(0)
and that T(s) lies inside y for s > 0. Verify that y(0) a T(0) determines the
orientation of y.
Let f: R2 -» R2 be a diffeomorphism and show that
(f-'y)(0) a (f.'r)(0) = Det Df(x„)[y(0) a T(0)] (E3.10)
gives the orientation of f-y.
188 3 Structural stability, hyperbolicity and homoclinic points
3.7.7 Let f: Q -* R2 be the horseshoe map and x* be a periodic point of period q > 1,
on the invariant Cantor set A. Use symbolic dynamics to construct a point xf of
A which is homoclinic to the periodic orbit containing x* (cf. Figure 3.35). Can
the transverse nature of the homoclinic point be detected by the symbolic
dynamics?
3.8.5
i —oo
sech(i) tanh(r) sin(tot) di = itco sech(jtro/2) by using contour inte-
gration on a rectangle in C with vertices at (+ R, 0), (± R, in) and letting R-> oo.
The Melnikov function given in (3.8.15) can also be used to indicate the separation
of stable and unstable manifolds of two different saddle points in a Hamiltonian
system (the so-called heteroclinic case). Show that the saddle connections, ,
between the fixed points (—Jt, 0) and (it, 0) of the system
x = y, y= —sin(x) + c(a —by) (E3.15)
with £ = 0, are given by
, (*o(t)> F'o(f)) = ( ± 2 arctan(sinh(f)), ± 2 sech(t)).
Calculate the Melnikov function for (E3.15) along these orbits and show that
M(0o) = + 2ax —8b, (E3.16)
for To , To, respectively. Explain why M(60) is constant.
Exercises 189
m ) = l f TunmcosW
(E3.18)
to ( cosh(nco/2)
Describe the regions of the (a, b)-p!ane for which transverse heteroclinic points
occur.
HINTS FOR EXERCISES
Chapter 1
1.1.1 W, = {exp(ix)|a < x < b}, W2 —{exp(ix)|c < x <d} such that IF, u VP2 = S',
(b —a),(d —c)< 2n. C”-overlap maps.
1.1.2 l y f - h '1= (hi -hJ" l)(h4-f*h“1)(hJ-h“1). Composition of two C*-maps is C* and
overlap maps h^-hj 1 and ly h “1 are C* since r'ak. Differentiability of f is
independent of charts.
1.1.3 (a) Pick open subsets A, B, C, D of R2 such that {/t(/t), /t(C), n(D)) is an
open covering of T2and restrictions of it to A, B, C, D are homeomorphisms.
(b) W, - S 2\N, N the north pole; W2 = S2\S, S the south pole, h,(h2) is the
stereographic projection from N(S) poles. Overlap map (r, <p)t—► (4/r, <p),r ^ 0.
1.2.1 Arnold, 1973, pp. 163-5.
1.2.2 (a) yes; (b) no; (c) circle map not homeomorphism.
1.2.3 Fixed points x = 0, all other points period-2.
1.2.4 Plot y= ]\x).
1.2.5 / an orientation-reversing homeomorphism on R implies it is strictly decreasing.
7(x+ 1) = f(x) - 1 as in proof of Proposition 1.2.1. Fixed points of / only at
intersection of y =f(x) with y = x and y = x + 1 (see Proposition 1.2.2).
1.3.1 Definition 1.3.1 implies ip, is C1 for all feR. <p,~1=ip-,-
1.3.4 X(x) = x - x 2.
1.3.5 (a)x = x3; (b)x = x, y = y2.
1.4.1 Minimal: (a) S'; (b) {x, Rw,(x),. . . , RJ,"‘(x)}, x eS 1.
General: (a) S1; (b) S = C/uRp/,(C/)u- • • R ^'ff/), U ^ S ', closed.
1.4.2 (a) Show fic is open.
1.4.5 Séparatrices connecting n = l, 2, 3, 4 saddle points enclosing unstable focus.
Consider Hamiltonian system with desired saddle connection and introduce
dissipation in the region bounded by the séparatrices, e.g.x = —2y{ 1 - x2) + /jxB(x ),
ÿ = 2x(l —y2) + fiyB(y), fi > 0,
exp[ —x2/(l —x2)], |x |< l
0 1x15=1.
Hints for exercises 395
dr/dfl = -t:r sin2 0(1 - r1 cos2 0) + 0(e2); r(0)- r(2/t) = -ctir^l - —^ + 0(c2).
Construct positively invariant set containing no fixed points.
1.5.1 (i) Similar construction to Example 1.5.1 for topological conjugacy with reference
intervals [1,2] for / and [1,8] for g.
1.5.4 Conjugacy preserves fixed points.
1.5.5 Use Proposition 1.2.2. Plot y = / 2(x). Conjugacy preserves periodic points.
1.5.6 Use (1.5.11).
1.5.7 Differentiate (1.5.11) and set x = 0.
1.5.8 If x* periodic with period-q, p(f) - 1 Lim(/I,,(x*) - x*)/nq Imod 1.
\«-* CO /
1.6.3 Consider separatrix of the saddle which is of opposite stability to that of the node.
1.6.4 Consider the lifted flow <j>,(x, y) = (x + i, y + at) on R2. Periodic orbits are given
by <Pt(x>y) = (x + m, y + n), T^O, m, neZ.
1.6.5 Recall that ( i p x ^),(x,y) = (q>,(x),^,(y)), x e M , ys N. Consider the lifted flows
q>,(x) = x + i, q>',(x) = x + 21/2t, I£,(x) = <A,'(x) = q>,{x).
1.6.6 Arrowsmith & Place, 1982, §2.3.
1.7.1 <p,(x) = xe/[xe - x + 1].
1.7.2 Fixed point x* must lie in I but X(x*) = 0.
1.7.3 Show that P2(q>to(x)) = q>t(i(P,(x)), xeS,.
1.7.4 Cylinder. Two limit cycles: stable x = 0; unstable x = 1.
1.7.5 (a) Möbius band; (b) Klein bottle.
1.8.2 x(f) = C exp(f - cos t).
1.8.4 IfQ(t) is a fundamental matrix so is Q(t + T)andQ(t) = Q(t + 7’)Q"l(t0+ T)Q(t0).
All P8o are conjugate.
/exp(>.,x) 0
1.8.5 (a) <p(!,t0) = , T= t - t 0;
\ 0 exp(A2t)/
^ -e x p f-tA
(b) x(t)= _
V,2exp(t) - 1/
1.8.7 Use polar coordinates. Null solution is stable.
cos t
1.8.8 x(t) =
,sin t ' siuV
cos l) (\- s m
0 t,'
1.9.1 Sketch level curves of //(x ,,x 2).
1.9.3 The generic case has non-zero eigenvalues.
1.9.4 Hamilton’s equations in plane polars are i- = i~' dH/cOJ)= - r " 1dH/dr. Examine
extrema of H as a function of r for various fixed values of 0.
d(r,0) 2 fl(T,0)
(|) ---------= _ ; (h ) ---------
1.9.5
S(x,y) r o(x, >■)
396 Hints for exercises
Chapter 2
2 . 1.1 If the Jordan form of L is not diagonal examine the powers of blocks of the form
/.I + N, JVy = &,j. ,. Observe (N % = 8Q . k, U k n - 1, N" = 0.
2. 1.2 (a) p = max{|J.,|........|x,|}, (b) Pick N > 3 such that p = JV,/W|A| < 1.
3 + 51,2\
2.1.3 (i) A|E": - — lu, orientation-preserving expansion;
3 —5 1,r .
A |P : v h v, orientation-preserving contraction;
/-1 -i o\
2.1.4 Real Jordan form of A is | 0 . A|£* is a rotational contraction.
\ 0 0 2/
2 . 1.6 x = Ax is linearly conjugate to y = Ay, A = Show that y i = Xiyi is
topologically conjugate to z, = sign(A,)Zj, i = 1, 2, 3, and use Exercise 1.6.5.
2.1.7 Use Theorem 2.1.2.
/0 1\
2 .2.2 D y,(0) = exp(DX(0)) = expl I has eigenvalues cosh (l)± sin h (l). Show that
2.2.5 Vector field is symmetric under clockwise rotation by n/2. A fixed point x* of </>,,
with topological type given by Dy,(x*) = exp(DX(x*)), becomes a periodic point
off = i>1-R .,/2.
2.2.6 Let x, = <pjx0) and define SJ, = Use (low box coordinates to prove
P^: So and P„:S0 -*S0 are C'-conjugate and result then follows from
Exercise 1.7.3.
2.2.7 Introduce cylindrical polar coordinates and recognise closed orbit for
r = (xf + = 1, z = x3 = 0. The Poincare map <p2„ defined on the plane
0 = constant has a fixed point at (r, z) = ( 1,0). Hyperbolicity follows from
D</>2*(L0) = exp(DX(l, 0)) and Hartman’s Theorem.
2.3.1 (i) Solve quadratic for y and expand square root.
(ii) Use (i) to obtain y, and substitute into expansion for (I + y,)~'.
2.3.4 Let h2(y) = ( + a 23’i.V2 + write down LAh2(y) and show that a, and
\biy2t +b2yiy2 + b3yl/
bit i = 1, 2, 3, can be chosen such that LAh2(y) = X2(y). Find at = a2= a3 = bl =
b2 = 1, b3 = 0.
2.3.5 Use resonance condition to show that is the only resonant term.
Hr.
2.4.1 (1) ad —b e ^ 0; (2) ad —b e = 0, a + d ^ 0; (3) ad —b e = a + d = 0, a1 + b 2 + c2 +
d1 ^ 0. cod{S,) = 0; cod(S2) = 1; cod(S3) = 2. Linear vector fields satisfying (2.4.1)
and (2.4.3) have codimension 1 and 2, respectively.
2.4.4 c = b, d = 2a.
2.4.5 Consider the types of Jordan block which give rise to non-hyperbolic linear
systems. Show that each type of block satisfies a resonance condition for all r > 2.
2.5.2 Observe that c > 0 implies f(x) > (< ) —x for x sufficiently small and positive
(negative).
2.5.3 Use (2.5.8) for complex form with n = 2. Observe that /„ , = 0 for m2 = 0 implies
no z-dependent terms arise. Alternatively, use (2.5.8) with n = 1 and a single
(complex) variable z (see Exercise 2.5.2). Note A,+ 1= A.
2.5.4 a = exp(3ia)/[l - exp(4ia)]. Note exp(4ia) £ 1 for a £ 2np/q, q = 1, 2, 3, 4.
398 Hints for exercises
-Tt/2
2 .6.1 has eigenvalues + i but
0 )■
2 .6.2 Note:
(i) if AB = BA then exp(A + B) = exp(A) exp(B);
(ii) N" = 0.
2.6.3 Let S_1MS = J, find InJ from Exercises 2.6.1 and 2. exp(lnJ) = J implies
L = Sln JS~'.
Alternative implies state transition matrix (itself a particular fundamental
matrix) ip(t, 0) = U(t) exp(Ct), make change of variable x = U(i)y and show that
y = Cy. Thus alternative implies Theorem 2.6.1 with A = C and B(i) = U(r).
Theorem 2.6.1 implies x = A(f)x has solutions x(t) = B(f) exp(At)y0 = </>(£, 0)xo =
Q(i)Q " 1(0)x„ for any fundamental matrix Q(t). Let y0 = Q ' ' (0)xo to obtain the
statement given in the question and y0 = x0 for B(t) = <p(t, 0) exp( - At).
2.6.4 If Jc is a Jordan block corresponding to a complex eigenvalue 22 of M2 then the
complex linear transformation that reduces Jc to the real Jordan form J R,
transforms In Jc into a real matrix, i.e. J K has a real logarithm.
If P = <p(2iz. 0) then P2 = </>(4it, 0) = exp(4nA), for real A, by the first part of
the question. Show that B(t) = <p(t,0)exp(- At) is 4tt-periodic in t.
2.6.5 Q(t) = (x+ ■x_), v(t,0) = Q(t)Q(0)-',
„ /cosh2tt sinh2tt\ f /0 l\]
* ” ■ (« 1 . oj}'
2 .6.8 (a) z|z|2, z|z|4; (b) z|z|2, z|z|4, z4 exp(2it).
2.6.9 For A, = 0, i = 1, 2, (2.6.14) implies resonance only if v = 0, i.e. all time-dependent
0 O'
2.7.1 A|ECgiven by(a)(jj J );(b )(J J); (c) . (b) gives unbounded motion.
.0 0.
2.7.2 Decompose R" into the direct sum P © Ec 0 E“ and consider restrictions of
exp(At) to £‘ and £“.
2.7.3 C®, unstable.
2.7.4 No, origin is hyperbolic node. Maximum differentiability given by [h/a].
2.7.5 «2j = 0, a3j = 0, a4j= - 4 \ a5j = 0, a6j-= 2(6, +')(1 - ( j f “ ); f a^i1 converges
for i = 4 if fi<$ and for i = 6 if ft <i. J~°
2.7.6 For C 540 centre manifold is non-analytic.
SO
2.7.11
2.8.1 Polar blowing-up gives: (i) saddles at 0 = 0, n, r> (<) 0 for r > 0 and 0 = 0 (n);
(ii) 0 = 0 unstable node, 0 = it stable node.
2.8.2 Singularities on r = 0 circle are; (a) 0 = 0 unstable node, 0 = Jt/4 saddle, 0 = ji/2
unstable node, 0 = n stable node, 0 = 5ji/4 saddle, 0 = 3/t/2 stable node; (b) 0 = 0
unstable node, 0 = ji/4, ji/2 saddles, 0 = it stable node, 0 = 5ji/4, 3ji/2 saddles.
2.8.3 Repeated blowing-up along positive y-axis gives further saddle-node singularities.
2.8.4 Division by |«|* and |u|* is necessary to prevent orientation reversal.
2.8.5 Positive x-blow-up, unstable node; negative x-blow-up, stable node. (cf. Exercise
2.8.1 with a - - 1 , b = 2.)
2.8.6 (a) Do polar blow-up, investigate resulting singularities at 0 = Jt/2, 3ji/2 with
further polar blow-ups. Obtain non-hyperbolic saddle.
(b) Polar blow-up gives six hyperbolic singularities. Obtain ‘monkey’ saddle.
Note that the unfoldings of the vector fields considered in this question appear
in Section 5.6 (see (5.6.2) (q = 2) and (5.6.14) (q = 3)). The reader may like to
confirm that the underlying singularity for q = 5 (see (5.6.34) and (5.6.35)) is a
focus, while q = 4 (see (5.6.21)) admits a variety of singularity types.
Chapter 3
3.1.1 Recall:
(i) the spectral radius, p(A), of A is the maximum of the absolute values of the
eigenvalues of A; (ii) the spectral norm, u(A), of A is the positive square root of the
largest eigenvalue of ATA; (iii) p(A) ^ <r(A); (iv) <r(A) ^ ||A||, where HAH = £ |ay|.
0
(a) Let M 'AM = D, D = [2,i5y]. Consider det(M" 'BM —pi), with p an
eigenvalue of B that is not equal to 2, for any i = l , . . . , n , and show
that d(D” 'C |)> 1 , where D,, = D —pi and C, = M _1CM. Observe
maxQAi-pl"1] ^ “1 implies min[|2, - p|] c.
I i
subspaces on which the restriction of A and B is not the identity. Hence construct
a conjugacy for A and B. A is not structurally stable in L(R2): consider
3.1.4 Show that A eO (R2) is a rotation so that every circle, centre x = 0, is invariant
under A. Observe B = (1 - c)A,e > 0 , has no invariant circles. Prove that conjugacy
preserves invariant circles.
3.2.1 (a) Use the Implicit Function Theorem and Exercise 3.1.1
(b) Use Exercise 3.1.1.
(c) DX(0) = DX(0).
3.2.2 (a) |r;l < e/4; (b) |f)| <f./32; (c) h/| <f./13. Use D<p2„(r0) = exp(2rcDA'r(r0)) (see
Exercise 2.2.2), where X r = r and Xr(r0) = 0.
3.2.3 (a) |<5| < e/(2 + R); (b) |i| < e/(R2 + 2R). The fixed point x* is not hyperbolic.
3.3.1 Theorem 3.3.1 is:
(a) applicable, non-hyperbolic closed orbit;
(b) applicable, non-hyperbolic fixed point at (x, y) = (1,0);
(c) not applicable, use perturbation (<5B( y),0), <5e R.
3.3.2 (a) Theorem 3.3.3(i) fails; (b) Theorem 3.3.3(iii) fails.
x = sin(2ttx), v = <5sin(2ny) ) . i(a)for|<5| < e/(1 + 2n)
> is e-C -close to <
x = 1, v = 2 + 5 sin [2 n (y -2 x )]J T(b) for |<5| < s/( 1 + 6x).
3.4.2 Recall that n: R2 -> T1 is a local diffeomorphism and differentiate irif’fx)) = f ’(it(x))
with respect to x to show that Tf’fBfx)) (see §3.6) and DT’(x) are conjugate.
DUIx) = A*, for all x, and A’ is hyperbolic. H/!-u(f(7t(x*)) = jt(A'x* + £*•"), where
q~1
P(E ") are the stable (unstable) eigenspaces of A. W' u = y tFs’"(f (ji(x*)).
i=0
x-coordinate, show that |F(x) —f"(x')| -> 0 as n -* oo. Eliminate xeQ\(~) Q’- "1.
neN
3.6.3 Let (l/j, h,), i = 1, 2, be overlapping charts and assume that h12 = h2h," Ul -*U2
is C*. Show that D(h,fh, ')(hj(x)) and D(h2flt2 ‘)(h2(x)) are similar. If f(x*) = x*
the eigenvalues of the tangent map Tfx. can be unambiguously defined to be those
of its local representatives. This means that a fixed point on M is hyperbolic if
all its local representatives are hyperbolic in the sense of Definition 2.2.1.
3.6.4 Chart T2 with n and define || 7TJ(v0)|| = |Df”(x)v,|, v ,e TT2
, and v,= Tn~'v,. A
has eigenvalues X, = (3 + 51,2)/2, X2 = (3 —51,2)/2. Take /i = |A,|~1= |A2|, C = 2
and c = {.
3.6.6 (a) f, is an Anosov automorphism; Theorem 3.6.1 applies; T1 connected implies
there is only one basic set 0, = T2.
(b) f2 has no periodic points but il= T 2 (note f2"(x, y) = (x, y + 2n(3',2|)mod 1,
neZ ); Theorem 3.6.1 is not satisfied.
(c) Four hyperbolic fixed points, Pu ...,P i -, Theorem 3.6.1 applies; il =
{p 1)p 2)p 3, / ,4-
3.7.2 Recall: a Cantor set is a closed, uncountable set with empty interior such that
every point is an accumulation point.
3.7.5 k = 4 sin2(a/2).
3.7.6 Note (f-T)(0) points into the image of y under f and Au a Av = Det(A)(u a v).
Chapter 4
4.1.1 (a) Terms of order r ^ 3 are not removed as in (4.1.14).
(b) Linear terms are not removed when ji, ytO. Transformed system is not an
unfolding of y= - y2.
4.1.2 (a) Take X{p,x) non-versal and Y(v, x) versa!. Show X ~ Y but Y ~X . For
example X(p, x) = x2, T(v, y) = v + y2.
(b) (i) X ~Y:hin0,n 1,x) = x - i t l/2,<f>in0,iti) = n0- n 2J4;
Y~X:h(v0, x) = y, </>(v0) = (v0, 0).
(ii) Let x = x*(jiB, ft,) be fixed point of x = X(fi0, ft,, x). Then
X ~Y : hOi0,/i1;x) = x-x*(;i0,^ 1),
V(Po,Pi) = {Pi-lx*(n0,iil)2, -3x*0io,ti,));
v. (2v\ vj
,Vj, y) = y - —, «>(vo,v1) = l - ^ - + v0v1,v0 + v1- -
Y~X:h(v0
4.1.3 Let x = ay, then y = T(i), >’), T(0, y) = ay2, becomes x = X(i(, x) = aT(fl,a_lx) so
Hints for exercises 403
6.8.2 (a) Use repeated root of (6.8.5) in (6.8.8) and expand in powers of v2. (b) Use the
Implicit Function Theorem on (6.8.8). Substitute r = r(to, v)(l +<r) in (6.8.4).
6.8.4 Introduce metric distance between two circles as maximum radial displacement.
Show that associated functional equation which maps circles using N, is effectively
the second component of (6.8.12).
6.8.5 Use induction on k and show that |N‘(z) + exp(2jripk/q)c,z, ~I|2 = |Nk(z)|2 up to
order |z|’.
6.8.6 Express in polar coordinates, take logarithms and separate real and imaginary
parts. Introduce local coordinate rL and consider Taylor expansion.
6.8.7 Use Theorem 5.4.2 and (5.4.14). Resonance tongue with tip at (0, v2). Show that
(v,, v2) near (0, v2) and (Re A, Im A) near (cos(2ttp'/q'), sin(2ap'/q')) are related by
a local diffeomorphism (cf Figures 5.5 and 5.6).
6.8.8 Use generalisation of (6.8.5) and (6.8.8).
REFERENCES