Part 1
Part 1
”
James C. Maxwell
“. . . le souci du beau nous conduit aux mêmes choix que celui de l’utile.”
Henri Poincaré
Preface
ix
x Preface
been repressed.
xii Preface
2 Chronologically, this discovery came after the creation of “relativity theory” and “quan-
tum mechanics”.
3 The possibility of applying hyperbolic attractors to nonlinear dynamics remains prob-
ẋ = −σ(x − y) ,
ẏ = rx − y − xz ,
ż = −bz + xy ,
material of this chapter also has reference value, beginners may call on it
when needed.
In Chap. 2 we examine the behavior of trajectories in a neighborhood of a
structurally-stable equilibrium state. Our approach here goes back to Poincaré.
Using this approach we classify the main types of equilibrium states. Special
attention is given to equilibria of the saddle types, and, in particular, to leading
and nonleading (strongly stable) invariant manifolds. We also give sufficient
attention to the asymptotic representation of solutions near a saddle point. As
mentioned, our methods are based on Shilnikov’s boundary-value problem. In
addition, we prove some theorems on invariant manifolds. We would like to
stress that along with well-known theorems on stable and unstable manifolds
of a saddle, some rather important results which we will need later are given
here. In the last section of the chapter some useful information concerning
Poincare’s theory of resonances for local bifurcation problems are presented.
In Chap. 3 we discuss structurally-stable periodic trajectories. Our con-
sideration is focused on the behavior of trajectories of the Poincaré map in a
neighborhood of the fixed point. As in the case of equilibria we investigate an
associated boundary-value problem near a saddle fixed point and prove a the-
orem on the existence of its invariant manifolds. Sections 3.10–3.12 and 3.14
are concerned only with the properties of periodic trajectories in continuous
time.
Invariant tori are considered in Chap. 4. More specifically, we study a non-
autonomous system which depends periodically, as well as quasi-periodically,
on time. This class of non-autonomous system can be extended to higher
dimensions by adding some equations having a specific form with respect to
cyclic variables. To prove the existence of an invariant torus in such a system,
we use a universal criterion, the so-called annulus principle which is applicable
for systems with small perturbations. In the case of a periodic external force,
the behavior of the trajectories on a two-dimensional invariant torus may be
modeled by an orientable diffeomorphism of a circle. In relation to this we
present a brief review of some related results from the Poincaré–Denjoy the-
ory. We complete this chapter with a discussion of an important problem of
nonlinear dynamics, namely, the synchronization problem associated with the
phenomenon of “beats” in modulations.
The final two chapters, Chap. 5 and 6, are dedicated to local and global
center manifolds, respectively. We re-prove in Chap. 5 a well-known result
that in a small neighborhood of a structurally unstable equilibrium state, or
xviii Preface
sometimes lead to subsequent confusion when more subtle details of the be-
havior of the trajectories are desired. The essence of our proof is a technique
(based on the reduction of the problem to a theorem on strong stable invariant
manifold) for making a series of coordinate transformations which are robust
to small, smooth perturbations of the system. We will use this special form in
the second part of this book when we study homoclinic bifurcations.
Last but not the least we would like to acknowledge the assistance of our
colleagues in the preparation of this book. They include Sergey Gonchenko,
Mikhail Shashkov, Oleg Sten’kin, Jorge Moiola and Paul Curran. In particu-
lar, Sergey Gonchenko helped with the writing of Sections 3.7 and 3.8, Oleg
Sten’kin with the writing of Section 3.9 and Appendix A, and Mikhail Shashkov
with the writing of Sections 6.1 and 6.2. We are also grateful to Osvaldo Garcia
who put the finishing touches to our qualitative figures.
We would also like to acknowledge the generous financial support from a US
Office of Naval Research grant (no. N00014-96-1-0753), a NATO Linkage grant
(no. OUTR LG96-578), an Alexander von Humboldt visiting award, the World
Scientific Publishing Company, and a special joint Professeur Invite Award (to
L. Chua) from the Ecole Polytechnique Federal de Lausanne (EPFL) and the
Eidgenoissische Technische Hochschule Zurich (ETH).
Leonid Shilnikov
Andrey Shilnikov
Dmitry Turaev
Leon Chua
Contents
Preface ix
xxi
xxii Contents
Bibliography 381
Index 389
Chapter 1
BASIC CONCEPTS
def dx
ẋ = = X(x) , (1.1.1)
dt
where x = (x1 , . . . , xn ), X(x) = (X1 , . . . , Xn ). We assume that X1 , . . . , Xn are
Cr -smooth (r ≥ 1) functions defined in a certain region D ⊆ Rn . In the theory
of dynamical systems it is customary to regard the variable t as time and the
region D as the phase space, which may be bounded or unbounded, or may
coincide with the Euclidean space Rn . A differentiable mapping ϕ : τ 7→ D,
where τ is an interval of the t-axis, is called a solution x = ϕ(t) of system
(1.1.1) if
ϕ̇(t) = X(ϕ(t)) , for any t ∈ τ . (1.1.2)
1
2 Chapter 1. Basic Concepts
(a)
Fig. 1.1.1. The projection of an integral curve onto the phase space D may be an unclosed
trajectory (a) or, for example, a periodic trajectory (b).
4 Chapter 1. Basic Concepts
(b)
For a proof of this theorem we refer the reader to the book Theory of Dy-
namical Systems on a Plane by Andronov, Leontovich, Gordon and Maier [6].
The trajectory L corresponding to a periodic solution ϕ(t) is called a
periodic trajectory.
Any other trajectory which is neither an equilibrium state nor a periodic
trajectory is an unclosed curve. It follows from Theorem 1.1 that an unclosed
trajectory has no points of self-intersection.
Note that any two solutions which differ from each other only in the choice
of the initial time t0 correspond to the same trajectory. Vice versa: any two
distinct solutions corresponding to the same trajectory are identical up to a
time shift t → t + C. It follows that all solutions corresponding to the same
periodic trajectory are periodic of the same period.
In the case where the solution corresponding to a given trajectory L is
defined for all t ∈ (−∞, +∞) we will say that L is an entire trajectory. Any
trajectory which lies in a bounded region is an entire trajectory.
From the view point of kinematics, the point ϕ(t) is called a representa-
tive point and its trajectory is called the associated motion. Moreover, for any
1.1. Necessary background from the theory of ODEs 5
trajectories other than equilibrium states, one can introduce a positive direc-
tion of the motion which points in the direction of increasing t. At each point of
such a trajectory this direction is determined by the associated tangent vector.
To emphasize this we will label all trajectories with arrowheads.
Along with system (1.1.1) let us consider an associated “time-reverse”
system
ẋ = −X(x) . (1.1.5)
The vector field of system (1.1.5) is obtained from that of (1.1.1) by reversing
the direction of each tangent vector. It is easy to see that each solution x = ϕ(t)
of system (1.1.1) corresponds to a solution x = ϕ(−t) of system (1.1.5) and
vice versa. It is clear also that systems (1.1.1) and (1.1.5) have the same phase
curves up to a change of time t → −t. Thus, the time-oriented trajectories of
one system are obtained from the corresponding trajectories of the other by
reversing the direction of the arrowheads.
Consider next the system
Generally speaking, not all trajectories may be continued over the infinite
interval τ = (−∞, +∞). In other words, not all trajectories are entire tra-
jectories.1 Examples of entire trajectories are equilibrium states and periodic
trajectories. From the point of view of dynamics, the entire trajectories, or
those which may be defined at least for all positive t over an infinite interval of
time, are of special interest. The reason is that, despite the importance of the
information revealed by transient solutions over a finite interval of time, the
most interesting phenomena observed in natural science and engineering ob-
tain an adequate explanation only if time t increases without bounds. Systems
whose solutions can be continued over an infinite period of time were named
dynamical systems by Birkhoff. An abstract definition of such systems which
takes into account their group properties, will be presented in the following
section.
1. ϕ(0, x) = x .
2. ϕ(t1 , ϕ(t2 , x)) = ϕ(t1 + t2 , x) . (1.2.1)
3. ϕ(t, x) is continuous with respect to (x, t) .
Observe that in the case of an unclosed trajectory any point of the trajec-
tory partitions the trajectory into two parts: a positive semi-trajectory and a
negative semi-trajectory.
In the case where the mapping ϕ(t, x) is a diffeomorphism4 the flow is
a smooth dynamical system. In this case, the phase space D is endowed
with some additional smooth structures. The phase space D is usually cho-
sen to be either Rn , or Rn−k × Tk , where Tk may be a k-dimensional torus
S1 × S1 × · · · × S1 , a smooth surface, or a manifold. This allows us to set up
| {z }
k times
a correspondence between a smooth flow and its associated vector field by
defining a velocity field
dϕ(t, x)
X(x) = . (1.2.2)
dt t=0
By definition, the trajectories of the smooth flow are the trajectories of the
system ẋ = X(x). In this book we will study mainly the properties of smooth
dynamical systems.
Discrete dynamical systems are often called cascades for simplicity. A
cascade possesses the following remarkable feature. Let us select a homeo-
morphism ϕ(1, x) and denote it by ψ(x). It is obvious that ϕ(t, x) = ψ t (x),
where
ψ t = ψ (ψ(· · · ψ (x))) .
| {z }
t−1 times
ẋ = X(x, t) ,
Let us denote by W the set of wandering points. The set W is open and
invariant. Openness follows from the fact that together with x0 any point in
U (x0 ) is wandering. The invariance of W follows from the fact that if x0 is
a wandering point, then the point ϕ(t0 , x0 ) is also a wandering point for any
t0 . To show this let us choose ϕ(t0 , U (x0 )) to be a neighborhood of the point
ϕ(t0 , x0 ). Then
Hence, the set of non-wandering points M = D\W is closed and invariant. The
set of non-wandering points may be empty. To illustrate the latter consider a
dynamical system defined by the autonomous system
ẋ = X(x, θ) ,
θ̇ = 1
in phase space Rn+1 , x = (x1 , . . . , xn ). Observe that (1.2.4) holds here since
θ(t) = θ0 + t increases monotonically with t. Hence, every point in Rn+1 is a
wandering point.
It is clear that equilibrium states, as well as all points on periodic trajecto-
ries, are non-wandering. All points on bi-asymptotic trajectories which tend to
equilibrium states and periodic trajectories as t → ±∞ are also non-wandering.
Such a bi-asymptotic trajectory is unclosed and called a homoclinic trajectory.
The points on Poisson-stable trajectories are also non-wandering points.
If for any T > 0 there exists t such that t < −T and (1.2.5) holds, then the
point x0 is called a negative Poisson-stable point. If a point is positive and
negative Poisson stable it is said to be Poisson-stable.
are called Poincaré return times. Two essentially different cases are possible
for an unclosed P -trajectory:
1. The sequence {τn (ε)} is bounded for any finite ε, i.e. there exists a
number L(ε) such that τn (ε) < L(ε) for any n. Observe that L(ε) → +∞
as ε → 0.
In the first case the P -trajectory is called recurrent. For such a trajectory
all trajectories in its closure Σ are also recurrent, and the closure itself is
a minimal set.8 The principal property of a recurrent trajectory is that it
returns to an ε-neighborhood of the point x0 within a time not greater than
L(ε). However, in contrast to periodic trajectories, whose return times are
fixed, the return time for a recurrent trajectory is not constrained.
In the second case, the closure Σ of the P -trajectory is called a quasi-
minimal set. In this case, there always exist in Σ other invariant closed sub-
sets which may be equilibrium states, periodic trajectories, or invariant tori,
6 See
the proof in [14].
7 In
the case of flows the set of times during which a P -trajectory passes through U ε (x0 )
consists of infinitely many time intervals In (ε), where tn (ε) is chosen to be one of the values
in In (ε).
8 A set is called minimal if it is non-empty, invariant, closed and contains no proper
Fig. 1.2.1. The flow on a torus can be represented as a flow on the unit square. The slope
of all parallel trajectories is equal to ω2 /ω1 . Gluing the opposite sides of the square gives a
two-dimensional torus.
etc. Since a P -trajectory may approach such subsets arbitrarily closely, the
Poincaré return times can therefore be arbitrarily large.
The simplest example of a flow all of whose trajectories are Poisson stable
is a quasi-periodic flow on a two-dimensional torus T2 defined by the equations
ẋ1 = ω1 ,
(1.2.6)
ẋ2 = ω2 ,
ẋ1 = ω1 f1 (x1 , x2 ) ,
ẋ2 = ω2 f2 (x1 , x2 ) ,
9 This is called a quasi-periodic trajectory.
12 Chapter 1. Basic Concepts
where
ρ(x, A) = inf kx − x0 k .
x0 ∈A
The study of any phenomenon which exhibits dynamical behavior usually be-
gins with the construction of an associated mathematical model of a dynam-
ical system in the form (1.2.1). Having a model in an explicit form allows
us to follow the evolution of its state as time t varies, since the initial data
1.3. Qualitative integration of dynamical systems 13
lim ϕ(tk ) = x∗ ,
k→∞
I. Equilibrium states.
Figure 1.3.1 shows examples of limit sets of type III where the equili-
brium states are labeled by O. Using the general classification above, we
may enumerate all types of positive semi-trajectories in planar systems:
1. equilibrium states;
2. periodic trajectories;
(a)
(b)
(c)
Fig. 1.3.1. Examples of two ω-limit homoclinic cycles in (a) and (c), and of a heteroclinic
cycle in (b) formed by two trajectories going from one equilibrium state to another.
16 Chapter 1. Basic Concepts
(a)
(b)
Fig. 1.3.2. (a) An ω-limit cycle. (b) A limit cycle which is both ω-limit and α-limit for
unclosed trajectories in its neighborhood.
Fig. 1.3.2. Such a periodic trajectory is called a limit cycle in the theory of
two-dimensional systems.
The corresponding situation in the case of higher dimensions is much more
complicated. In this case, in addition to equilibrium states and periodic tra-
jectories, the limit set may be a minimal, or a quasi-minimal set of various
topological types, such as a strange attractor in the form of a smooth, or a
non-smooth manifold, or a fractal set with a local structure represented as a
direct product of a disk and a Cantor set and ever more exotic sets.
1.3. Qualitative integration of dynamical systems 17
Let us now turn to the problem concerning the study of the totality of
trajectories. In fact, characterizing a dynamical system means topologically
(or qualitatively) partitioning the phase space into the region of the existence
of trajectories of different topological types. We usually refer to this problem
as “constructing the phase portrait”. This problem poses the question: When
are two phase portraits similar? In terms of the qualitative theory of dynamical
systems we can answer this question by introducing the notion of topological
equivalence.
This definition implies that equilibrium states, as well as periodic and un-
closed trajectories of one system, are respectively mapped into equilibrium
states, as well as periodic and unclosed trajectories of the other system. The
topological equivalence of two systems in some sub-regions of the phase space
is defined in a similar manner. The latter is usually used for studying local
problems, for example, in a neighborhood of an equilibrium state, or near pe-
riodic or homoclinic trajectories. The definition of topological equivalence of
two dynamical systems gives an indirect definition of the qualitative structure
of partition of the phase space into the regions of the existence of trajectories
of topologically different types. Such structures must be invariant with respect
to all possible homeomorphisms of the phase space.
Let G be a bounded sub-region of the phase space and let H = {hi } be a
set of homeomorphisms defined on G. We can introduce a metric as follows
hi L = L .
It is clear that all equilibrium states and periodic trajectories are special
trajectories. Unclosed trajectories may also be special. For example, all tra-
jectories of a two-dimensional system which tend to saddle equilibrium states
11 See Sec. 2.5 for details.
18 Chapter 1. Basic Concepts
L2 = hm(ε) · · · h1 L1 .
This theorem, together with its proof, occupies a significant part of the book
Theory of Dynamical Systems on the Plane by Andronov, Leontovich, Gordon
and Maier [6]. This theory provides not only a mathematical foundation for a
theory of oscillations of two-dimensional systems but also gives a recipe for the
investigation of concrete systems. In particular, the investigation proceeds in
the following order: First, classify the equilibrium states, and then all special
trajectories such as separatrices tending to saddle equilibria and trajectories
approaching limit sets of type III, either as t → +∞, or as t → −∞. This
entire collection of special trajectories determines a schematic portrait called
a skeleton which allows one to partition the phase space into cells, as well as
to study the behavior of the trajectories within each cell.
Unfortunately, this does not work when we examine systems of higher
dimensions. The set of special trajectories in a three-dimensional system may
12 The set S can be considered as a factor-system with respect to the above equivalence
relation.
13 The finiteness condition of S is rather general, holding for a wide class of planar systems.
1.3. Qualitative integration of dynamical systems 19
already be infinite, or may even form a continuum. The same situation applies
to cells. Thus, the problem of finding a complete topological invariant in this
case seems to be quite unrealistic. This is the reason why we must reconcile to
the concept of a relatively-incomplete classification based on some topological
invariants which apply only to certain cases. Nevertheless, the basic approach
for studying concrete high-dimensional systems remains the same as in
the two-dimensional case; namely, it begins by examining the equilibrium
states and the periodic trajectories. We will consider this “comprehensive”
local theory in Chaps. 2 and 3, respectively.
Chapter 2
ẋ = X(x) (2.1.1)
It follows from the definition that the coordinates of the equilibrium state
can be found as the solution of the system:
X(x0 ) = 0 . (2.1.2)
21
22 Chapter 2. Structurally Stable Equilibrium States of Dynamical Systems
of (2.1.1) is polynomial there are standard algebraic methods for the evaluation
of the number of equilibrium states.
From the point of view of numerical simulations the determination of all
isolated solutions of system (2.1.2) (or equivalently of all stationary states of
(2.1.1)) in a bounded subregion of Rn is a relatively simple task for small n.
However, the number of equilibrium states of a system of higher dimension may
be very large and therefore searching for all of them becomes problematic.
The study of system (2.1.1) near an equilibrium state is based on a standard
linearization procedure.
Let a point O(x = x0 ) be an equilibrium state of system (2.1.1). The
substitution
x = x0 + y (2.1.3)
shifts the origin to O. With the new variables the system may be written as
ẏ = X(x0 + y) , (2.1.4)
ẏ = Ay + g(y) , (2.1.6)
where
∂X(x0 )
A= ;
∂x
A is a constant (n × n)-matrix and g(y) satisfies the condition
∂g(0)
g(0) = = 0. (2.1.7)
∂y
In the general case, the last term in (2.1.6) is of a higher order of smallness
(with respect to the usual norm) than the first term. It is apparent that the
behavior of the trajectories of system (2.1.6) in a small neighborhood of the
origin is governed primarily by the linearized system
ẏ = Ay . (2.1.8)
main source of such systems was the theory of automatic control, in particular,
the control theory of steam engines. The central problem of linear dynamics
in that period was the search for the most effective criteria of stability for
stationary states.1
The stability of an equilibrium state is determined by the eigenvalues
(λ1 , . . . , λn ) of the Jacobian matrix A which are the roots of the character-
istic equation
det |A − λI| = 0 (2.1.9)
where I is the identity matrix. The roots of the characteristic equation are also
called the characteristic exponents of the equilibrium state. The equilibrium
state is stable when all of its characteristic exponents lie in the left half-plane
(LHP) on the complex plane. Moreover, any deviations from equilibrium de-
cay exponentially with decrements of damping proportional to the values Re λ i ,
(i = 1, . . . , n). Thus, the primary problem of constructing a simple and effec-
tive criterion of the stability of an equilibrium state was in finding some explicit
conditions in terms of the entries of the matrix A such that it would allow one
to determine, without having to solve the characteristic equation, when all of
its eigenvalues lie in open LHP.
Here, we present the most popular algorithm called a Routh–Hurwitz cri-
terion. Let (a0 , . . . , an ) be the coefficients of the polynomial det |λI − A|:
det |λI − A| = a0 λn + a1 λn−1 + · · · + an .
We construct an (n × n)-matrix:
a1 a3 a5 ··· 0 0
a0 a2 a4 ··· 0 0
0 a1 a3 ··· 0 0
à = 0 a0 a2 ··· 0 0 (2.1.10)
.. .. .. .. .. ..
. . . . . .
0 0 0 ··· an−1 0
0 0 0 ··· an−2 an
and find the minors ∆1 = a1 , ∆2 = a1 a2 − a0 a3 , . . . , ∆n = det Ã. Here, ∆i is
the determinant of the matrix whose entries lie on the intersection of the first
i rows and the first i columns of the matrix Ã.
1 The necessity for studying nonlinear nonconservative systems emerged only in the first
part of the 20th century, in the context of the investigation of the phenomenon of sustained
oscillations in vacuum-tube oscillators.
24 Chapter 2. Structurally Stable Equilibrium States of Dynamical Systems
In this and in the following two sections we will study the behavior of solutions
of the linearized system. Moreover, we will restrict ourselves to structurally
stable equilibrium states only.
Let us begin with low dimensional cases n = 2 and n = 3.
When n = 2 the system assumes this general form:
ẋ = a11 x + a12 y ,
(2.2.1)
ẏ = a21 x + a22 y .
1. Both λ1 and λ2 are real and negative: λ1 < 0 and λ2 < 0. Such an
equilibrium state O is called a stable node. When λ1 6= λ2 system (2.2.1)
can be reduced to
ξ˙ = λ1 ξ ,
(2.2.3)
η̇ = λ2 η
ξ = e λ 1 t ξ0 , η = e λ2 t η 0 . (2.2.4)
Since both λ1,2 are negative, all trajectories are attracted to the origin
as t → +∞. Furthermore, every trajectory approaches the origin O
tangentially either to the ξ-axis or to the η-axis. In order to verify this,
let us examine the following equation of the integral curves of the system
(2.2.3)
ηξ0ν = ξ ν η0 (2.2.5)
where ν = |λ2 |/|λ1 |. For definiteness, let |λ2 | be greater than |λ1 |. Then
ν > 1 and, by virtue of (2.2.5), all trajectories approach O tangentially
to the ξ-axis except for the two trajectories which lie on the η-axis, see
Fig. 2.2.1. The ξ- and η-axes are respectively called the leading and the
non-leading directions.
When λ1 = λ2 = −λ < 0 system (2.2.1) can be written in one of the
following forms below:
ξ˙ = −λξ + η ,
(2.2.6)
η̇ = −λη
ξ˙ = −λξ ,
(2.2.7)
η̇ = −λη .
26 Chapter 2. Structurally Stable Equilibrium States of Dynamical Systems
Fig. 2.2.1. A stable node. Double arrows label the strongly stable (non-leading)
direction which coincides with the η-axis.
Figure 2.2.2 shows the phase portrait in the first case. All trajectories
tend to O tangentially to the unique eigenvector, namely, the ξ-axis. In
the second case any trajectory approaches O along its own eigen-direction
as shown in Fig. 2.2.3. Such a node is called a dicritical node.
ξ˙ = −ρξ − ωη ,
(2.2.10)
η̇ = ωξ − ρη .
2.2. Qualitative investigation of linear systems 27
Fig. 2.2.2. Another stable node. Every trajectory enters the origin along the only
leading direction which is the ξ-axis.
Fig. 2.2.3. A dicritical node. Every trajectory tends to O along its own direction.
28 Chapter 2. Structurally Stable Equilibrium States of Dynamical Systems
ṙ = −ρr ,
(2.2.11)
ϕ̇ = ω .
r(t) = e−ρt r0 ,
(2.2.12)
ϕ(t) = ωt + ϕ0 ,
ξ˙ = γξ ,
(2.2.14)
η̇ = −λη .
ηξ ν = ξ0ν η0 , (2.2.16)
where ν = λ/γ. The portrait of the phase space (or just “the phase
space”) near the saddle is shown in Fig. 2.2.5. There are four exclusive
trajectories called the separatrices, two stable and two unstable, which
tend to the saddle O as t → +∞ and as t → −∞ respectively. All other
trajectories pass by the saddle. The pair of the stable separatrices to-
gether with the saddle O compose the stable invariant subspace of the
4. The case where the real parts of both characteristic exponents are positive
can be simply reduced to the cases (1) and (2) above by the change
of time t → −t, so that the directions of the arrows in the respective
phase portraits are reversed. When the characteristic exponents are real,
the associated equilibrium state is called an unstable node. In the case
of complex characteristic exponents it is called an unstable focus (see
Figs. 2.2.6 and 2.2.7).
ẋ = λ1 x , ẏ = λ2 y , ż = λ3 z . (2.2.17)
Fig. 2.2.6. An unstable node. The picture is obtained from Fig. 2.2.1 by reversing
the time.
2.2. Qualitative investigation of linear systems 31
Fig. 2.2.7. An unstable focus. A trajectory traces out a “clockwise” spiral on the
plane.
Fig. 2.2.8. A stable node in R3 . The fewer the arrows the weaker is the rate of
contraction. The leading subspace E L is one-dimensional, two-dimensional subspace
E ss is non-leading.
(see (2.2.13)). The phase portrait of this system is shown in Fig. 2.2.9.
It follows from (2.2.20) that
p q
y(t)2 + z(t)2 = e−ρt y02 + z02 .
Moreover, for any trajectory whose initial point does not lie in the
non-leading plane (y, z), we obtain
p
y(t)2 + z(t)2 = C|x(t)|ν (2.2.21)
p
where ν = ρ/|λ1 | and C = y02 + z02 /|x0 |ν . Since ν > 1, all such trajec-
tories approach O along the leading x-axis.
Fig. 2.2.9. Another possible stable node in R3 . Although the point O is a stable
focus on E ss , all trajectories outside E ss go to O along the one-dimensional leading
subspace E L .
Fig. 2.2.10. A stable focus in R3 . In contrast to Fig. 2.2.9, all trajectories outside
of the one-dimensional subspace E ss tend to O tangentially to the two-dimensional
leading subspace E L .
34 Chapter 2. Structurally Stable Equilibrium States of Dynamical Systems
7. When all characteristic exponents lie to the right of the imaginary axis
(i.e. in the open right-half plane (RHP)), by reversion of time t → −t, we
reduce the problem to the cases considered above. Here, all trajectories
tend to the equilibrium state as t → −∞. As before, there exist two kinds
of equilibrium states, namely: an unstable node, if the characteristic
exponent nearest to the imaginary axis is real, and an unstable focus,
when the characteristic exponents nearest to the imaginary axis comprise
a complex-conjugate pair. The corresponding phase portraits are similar
to those shown in Figs. 2.2.8–2.2.10 but with all the arrows pointing in
the opposite direction.
8. If there are characteristic exponents both to the left and to the right of
the imaginary axis, the equilibrium state is either a saddle or a saddle-
focus (this name was also given by Poincaré), see Figs. 2.2.11–2.2.14.
Let us assume that λ1 > 0 and λs < 0, (s = 2, 3) in (2.2.17). Then,
the equilibrium state of system (2.2.17) is a saddle, see Fig. 2.2.11.
The general solution is also given by (2.2.18). Because λ1 > 0, λ2 < 0,
λ3 < 0, the coordinates y and z decay exponentially to zero as t → +∞
Fig. 2.2.11. A saddle O with the two-dimensional stable subspace E s and the one-
dimensional unstable subspace E u .
2.2. Qualitative investigation of linear systems 35
and, therefore
eλ 1 t 0
eλ 2 t
eAt = .. .
.
0 e λn t
Thus, if we denote the components of the vector y ∈ Rn by (y1 , . . . , yn ) in the
given basis, then the solution of the system can be rewritten in the form:
If all eigenvalues (we will also call them the characteristic exponents) are
different as before, but some of them are complex, then there exists a basis in
which A attains a block-diagonal form:
A1 0
A2
.. , (2.3.5)
.
0 Am
where each block Aj corresponds to either a real eigenvalue, or to a pair of
complex-conjugate eigenvalues (recall that if A is the real matrix, then the
complex-conjugate λ∗i of any complex eigenvalue λi is also an eigenvalue). If
λj is real, its corresponding block is a (1 × 1)-matrix:
Aj = (λj ) . (2.3.6)
hence we obtain
eA 1 t 0
eA 2 t
eAt = .. ,
.
0 e Am t
where
λt
e for Aj = (λ)
Re eλt −Im eλt
cos(ωt) − sin(ωt)
ρt
= e
eA j t
= Im eλt Re eλt sin(ωt) cos(ωt)
ρ −ω
for Aj = .
ω ρ
or, equivalently
yi+1 (t) ··· yi+2k−1 (t) Λt yi+1,0 ··· yi+2k−1,0
=e eJk t ,
yi+2 (t) ··· yi+2k (t) yi+2,0 ··· yi+2k,0
(2.3.16)
and the exponential is separately found for each sub-block. Analogous calcu-
lations are carried out for each complex eigenvalue λ of multiplicity k with a
non-complete Jordan block.
We can now prove the following lemma which gives a standard estimate
for the norm of the exponential matrix. This estimate will be frequently used
throughout the book.
42 Chapter 2. Structurally Stable Equilibrium States of Dynamical Systems
Lemma 2.1. For given an arbitrarily small ε > 0, one can choose an appro-
priate basis in Rn such that the solution y(t) = eAt y0 of the linear system
ẏ = Ay
Proof. The proof is analogous for all four inequalities, so let us consider the
first one. To prove (2.3.35) in the case when all characteristic exponents are
simple, we choose the basis such that equalities (2.3.12), (2.3.13) and (2.3.10)
hold whence (2.3.35) follows immediately.
In the case of multiple characteristic exponents, after choosing the Jordan
basis the formulae for y(t) (see (2.3.29) and (2.3.31)) have power factors t k
which give the following estimate for the norm of y(t):
To make the constant C equal to 1 we note that the Jordan basis can be
chosen such that non-zero values δj ’s in (2.3.15) and (2.3.22) are equal to the
given arbitrarily small ε. To do this, instead of the coordinates
(yi+1 , . . . , yi+k ) ,
2.3. High-dimensional linear systems. Invariant subspaces 43
which correspond to the Jordan block (2.3.15) for real eigenvalues, we must
choose the coordinates
Similarly, for complex eigenvalues (see formula (2.3.22)), we must replace the
coordinates
(yi+1 , . . . , yi+2k )
by the coordinates
(yi+1 /εk−1 , yi+2 /εk−1 , yi+3 /εk−2 , yi+4 /εk−2 , . . . , yi+2k−1 , yi+2k ) .
In this basis, the factor εs−j appears in front of ts−j in formulae (2.3.29) and
(2.3.31) or, equivalently, the coefficient ε appears in front of Jk t in (2.3.30) and
(2.3.32). As a result we obtain the following estimate for y(t)
Thus, instead of (2.3.35), for example, we obtain the following inequality for
x(t):
kx(t)k ≤ Ce(max Re λi +ε)t kx0 k for t ≥ 0 , (2.3.22)
44 Chapter 2. Structurally Stable Equilibrium States of Dynamical Systems
where
C = kP k kP −1 k ≥ 1 . (2.3.23)
In the particular case where all of the characteristic exponents λi lie to the
left of the imaginary axis, inequality (2.3.35) becomes the following
where λ > 0 is such that Re λi < −λ at all i (and if the characteristic exponents
nearest to the imaginary axis are simple, one may choose λ = min |Re λi |).
Thus, in this case, every trajectory of the linear system (2.3.1) tends exponen-
tially to O as t → +∞. Such an equilibrium state is called an exponentially
asymptotically stable equilibrium state.
Let us reorder the characteristic exponents of the stable equilibrium state
so that Re λ1 ≥ Re λ2 ≥ · · · ≥ Re λn . We assume also that first m expo-
nents have the same real parts Re λi = Re λ1 (i = 1, . . . , m) and Re λi <
Re λ1 (i = m + 1, . . . , n). Let us denote by E L and E ss the m-dimensional and
the (n − m)-dimensional eigen-subspaces of the matrix A, which correspond
to the characteristic exponents (λ1 , . . . , λm ) and (λm+1 , . . . , λn ), respectively.
The subspace E L is called the leading invariant subspace and E ss is called the
non-leading or the strongly stable invariant subspace.
These names are derived from the fact that as t → +∞ all trajectories,
except for those lying in E ss , tend to the equilibrium state O tangentially to
the subspace E L . Moreover, the trajectories from E ss tend to O faster than
e(Re λm+1 +ε)t , whereas the convergence velocity of the other trajectories does
not exceed e(Re λ1 −ε)t , where the constant ε > 0 may be chosen arbitrarily
small.
To prove this statement we note that each vector y ∈ Rn is uniquely
represented in the form y = u + v, where u ∈ E L and v ∈ E ss . In the (u, v)
coordinates system (2.3.1) is written as
u̇ = A1 u ,
v̇ = A2 v ,
According to Lemma 2.1 (see (2.3.36), (2.3.35)), it follows from (2.3.43) that
where ν > 1. It is seen from (2.3.44) that if ku0 k 6= 0, then any trajectory
approaches O tangentially to the leading subspace v = 0.
In the case m = 1, i.e. where λ1 is real and Re λi < λ1 , (i = 2, . . . , n), the
leading subspace is a straight line. Such an equilibrium state is called a stable
node (see Figs. 2.2.8, 2.2.9).
If m = 2 and λ1,2 = −ρ ± iω, ρ > 0, ω 6= 0, then the corresponding
equilibrium state is called a stable focus. The leading subspace here is two-
dimensional and all trajectories which do not belong to E ss have the shape of
spirals winding around towards O, see Fig. 2.2.10.
The unstable case, where Re λi > 0, (i = 1, . . . , n), is reduced to the previ-
ous one by reversion of time t → −t. Therefore, the solution can be estimated
as:
ky(t)k ≤ e−λ|t| ky0 k , for t ≤ 0, (2.3.27)
u̇ = A− u ,
(2.3.28)
v̇ = A+ v
According to Lemma 2.1, for the variables u and v the estimates analogous,
respectively, to (2.3.42) and (2.3.45) are valid, i.e. any trajectory from the stable
invariant subspace E s : v = 0 tends exponentially to O as t → +∞, and any
trajectory from the unstable invariant subspace E u : u = 0 tends exponentially
to O as t → −∞; the neighboring trajectories pass nearby but away from the
saddle.
Thus, the saddle is the stable equilibrium for the system on E s and is
completely unstable on E u . Furthermore, stable and unstable leading and non-
leading subspaces, respectively, E sL , E uL , E ss and E uu can be defined in the
subspaces E s and E u . We will call a direct sum E sE = E s ⊕ E uL the extended
stable invariant subspace, and E uE = E u ⊕ E sL the extended unstable invariant
subspace. The invariant subspace E L = E sE ∩ E uE is called the leading saddle
subspace.
If the point O is a node in both E s and E u , such an equilibrium state is
called a saddle. Therefore, the dimensions of both E sL and E uL are equal
to 1.
When the point O is a focus on at least one of two subspaces E s and E u ,
then O is called a saddle-focus. Depending on the dimensions of the stable
and the unstable leading subspaces, we may define three types of saddle-foci;
namely:
The phase portraits for two types of three-dimensional saddles and saddle-
foci (2, 1) and (1, 2) are shown in Figs. 2.2.11–2.2.14; a four-dimensional saddle-
focus (2, 2) is schematically represented in Fig. 2.3.1.
2.4. Trajectories near a saddle 47
Fig. 2.3.1. The pseudo-projection of a saddle-focus (2,2) into R 3 . Both stable and unstable
invariant subspaces are of dimension two.
ẋ = −λ1 x ,
u̇ = −λ2 u ,
ẏ = γy ,
where 0 < λ1 < λ2 , γ > 0. The unstable subspace E u here coincide with
the y-axis, and the stable subspace E s is the (x, u)-plane. The u-axis is the
non-leading subspace E ss and the x-axis is the leading subspace E sL . The
extended unstable subspace E uE is the (x, y)-plane and the extended stable
subspace E sE is the entire space R3 .
The general solution of the system is
x(t) = e−λ1 t x0 ,
u(t) = e−λ2 t u0 ,
y(t) = eγt y0 .
δ = eγt y . (2.4.2)
2.4. Trajectories near a saddle 49
Fig. 2.4.1. Trajectories near a saddle. The image of the rectangle Π+ is a curvilinear triangle
on Π− which is tangential to the extended unstable subspace E uE . The intersection of the
stable subspace E s on Π+ is mapped into the point M − .
1 δ
t= ln .
γ y
Substituting the latter into (2.4.1) gives us the expression for the coordinates
of the points M in terms of those of M̄
y ν
x̄ = x+ ,
δ (2.4.3)
y αν
ū = u ,
δ
where ν = λ1 /γ, α = λ2 /λ1 > 1. It is clear from this formula that the map
M 7→ M̄ is contracting with respect to the non-leading coordinate u provided
50 Chapter 2. Structurally Stable Equilibrium States of Dynamical Systems
Fig. 2.4.2. The behavior of trajectories near this saddle is the reverse situation depicted in
Fig. 2.4.1.
2.4. Trajectories near a saddle 51
Fig. 2.4.3. The map near a saddle-focus (2, 1). The zebra pattern on Π+ is mapped along
the trajectories inside two spirals around the point M − which is the image of the intersection
of E s and Π+ .2
When O is a saddle-focus (2,1) (see Fig. 2.4.3) the system can be repre-
sented in the form
ẋ = −ρx − ωu ,
u̇ = ωx − ρu , (2.4.5)
ẏ = γy ,
2 Remark. We must notice that the formulae below are derived for the case where the
x̄ = xe−ρt cos(ωt) ,
ū = xe−ρt sin(ωt) ,
δ = yeγt ,
or
y ν
ω y
x̄ = x cosln ,
δ γ δ
y ν
ω y
ū = −x sin ln ,
δ γ δ
ẋ1 = −ρ1 x1 − ω1 x2 ,
ẋ2 = −ρ1 x2 + ω1 x1 ,
ẏ1 = ρ2 y1 − ω2 y2 ,
ẏ2 = ρ2 y2 + ω2 y1
The map T from Π+ = {x2 = 0, |x1 − x+ 1 | < ε, |y1 | < ε, |y2 | < ε} into
Π = {ȳ2 = 0, |ȳ1 − y1− | < ε, |x̄1 | < ε, |x̄2 | < ε} along the trajectories of the
−
give us the values of x̄1 , x̄2 , y1 , y2 such that the trajectory starting with the
point M = (x1 , 0, y1 , y2 ) of Π+ intersects Π− at the point M̄ = (x̄1 , x̄2 , ȳ1 , 0).
The domain D is composed of all points M whose x1 -coordinate lies in the
interval |x1 − x+
1 | < ε and whose y1,2 -coordinates are found from (2.4.9) for
some appropriate choice of ȳ1 and t.
By virtue of (2.4.9), the point (y1 , y2 ) traces out a logarithmic spiral as
t varies while x1 and ȳ1 are kept fixed. This means that the set D has the
shape of a roulette stretched along the x1 -direction, and twisted in the (y1 , y2 )-
coordinates. The image of D in Π− has a similar shape, see Fig. 2.4.4.
x
2
y y
1 2
y
1
x
1 x
1
Fig. 2.4.4. The map near a saddle-focus (2, 2). See comments in the text.
2.4. Trajectories near a saddle 55
We observe that (by fixing the size of both neighborhoods) the flight time t
grows to infinity as y → 0, proportionally to ln kyk. Moreover, by Lemma 2.1,
the following estimates
+
kūk ≤ keB t k kuk ≤ e−λ̂t kuk ,
−
kvk ≤ ke−B t k kv̄k ≤ e−γ̂t kv̄k ,
hold, which implies that the map is strongly contracting in the non-leading
stable directions and strongly expanding in the non-leading unstable directions,
provided kyk is sufficiently small. In fact, it follows from formulae (2.4.10) and
(2.4.11) that:
ẏ = Ay (2.5.1)
ẏ = Ay + g(y) (2.5.2)
Fig. 2.4.5. The map near a saddle in R4 . The point O has two eigenvalues with positive
real parts and two eigenvalues with negative real parts, i.e. the saddle has a two-dimensional
stable and a two-dimensional unstable subspace.
Fig. 2.4.6. The map near a saddle in R4 . The point O is a saddle with a three-dimensional
stable subspace E s and a one-dimensional unstable subspace E u .
58 Chapter 2. Structurally Stable Equilibrium States of Dynamical Systems
Fig. 2.4.7. The map along trajectories passing by a saddle-focus (2,1). The difference between
the maps is in the dimension of the stable and the unstable subspaces. See the location of
eigenvalues.
Fig. 2.4.8. The map along trajectories passing by a saddle-focus (2,1). The difference between
the maps is in the dimension of the stable and the unstable subspaces. See the location of
eigenvalues.
2.5. Topological classification of equilibrium states 59
η : U 1 → U2 ,
ẋ = −ωy + g1 (x, y) ,
(2.5.3)
ẏ = ωx + g2 (x, y) ,
where we suppose that the functions g1 and g2 vanish at the origin along with
their first derivatives. The general solution of the associated linearized system
is given by
x = x0 cos(ωt) − y0 sin(ωt) ,
y = y0 cos(ωt) + x0 sin(ωt) .
then the following general solution of Eq. (2.5.3) can be easily found in polar
coordinates:
1
r2 = , ϕ = ωt + ϕ0 .
2t + r0−2
In this case, all trajectories have the shape of spirals winding around the
origin as shown in Fig. 2.5.2. Evidently, in any small neighborhoods of both
equilibrium states there is no homeomorphism that maps trajectories of such
a system onto those of the linearized system (since a homeomorphism maps
closed curves onto closed curves). Thus, our system is not topologically equiv-
alent to its linearization.
For our second example, let one exponent λ1 be equal to zero and let the
other exponent be equal to λ2 = −λ < 0. This system can be written in the
form
ẋ = g1 (x, y) ,
(2.5.4)
ẏ = −λy + g2 (x, y) ,
2.5. Topological classification of equilibrium states 61
Fig. 2.5.2. Accounting for nonlinearities causes a change in the behavior of the trajectories
near the center. They behave like spirals winding around O.
where the functions g1 and g2 vanish at the origin along with their first deriva-
tives. The solution of the linearized system is
x = x0 , y = e−λt y0 .
The phase portrait is shown in Fig. 2.5.3. The entire x-axis consists of equi-
librium states of the linearized system, and, each equilibrium state attracts
only one pair of trajectories. It is obvious that the nonlinear system may
preserve a continuum of equilibrium states only for a very special choice of
functions g1 and g2 and therefore topological equivalence between the original
and linearized system is scarcely expected here.
Figure 2.5.4. demonstrates the phase portrait when g1 = x2 , g2 = 0. One
can see that the two local phase portraits have nothing in common. The
equilibrium state in Fig. 2.5.4 is called a saddle-node.
The problem of topological classification of structurally stable equilibrium
states finds its solution in the following theorem:
Theorem 2.1. (Grobman Hartman) Let O be a structurally stable equilib-
rium state. Then, there are neighborhoods U1 and U2 of O where the original
and the linearized systems are topologically equivalent.
62 Chapter 2. Structurally Stable Equilibrium States of Dynamical Systems
Fig. 2.5.3. Each point on the x-axis is a stable equilibrium state which attracts a pair of
trajectories.
Fig. 2.5.4. A structurally unstable point of saddle-node type. The point O is stable in the
“node” region (x < 0) but unstable in the “saddle” region (x > 0).
2.5. Topological classification of equilibrium states 63
We note that the equilibrium state of the nonlinear system (2.5.2) is then
said to be locally topologically equivalent to that of its linear part (2.5.1).
Let us go further and set up the question of topological equivalence among
linear systems. Assign to a structurally stable equilibrium state the topological
type (k, n − k) when k characteristic exponents lie to the left of the imaginary
axis, and (n − k) to the right of it.
Theorem 2.2. Linear systems with equilibrium states of the same type are
topologically equivalent.
The proof of this theorem is constructive in the sense that the homeomor-
phism η: Rn 7→ Rn may be found explicitly. For example, let us consider two
linear systems, the first has a focus at the origin
ẋ = −x + y ,
(2.5.5)
ẏ = −x − y
ẋ = Ak x (2.5.7)
u
where by Ii we denote the i-dimensional identity matrix. If we assume x = v
where u ∈ Rk , v ∈ Rn−k , then the system (2.5.7) may be represented in the
form
u̇ = −u ,
(2.5.8)
v̇ = v .
In the case k = n all trajectories of the system (2.5.9) tend to the equilib-
rium state at the origin as t → +∞. Hence, by virtue of Theorems 2.1 and
2.2, any trajectory from a sufficiently small neighborhood of an equilibrium
state of type (n, 0) of the nonlinear system also tends to the equilibrium state
as t → +∞. Such an equilibrium state is called a stable topological node or
sink. We remark that the n-dimensional stable foci and nodes considered in
the previous section are topologically equivalent by virtue of Theorem 2.2 and
therefore both are stable sinks.
For an equilibrium state O of type (0, n), any trajectory from a small
neighborhood of O tends to O as t → −∞. As t → +∞ any trajectory,
excluding O, leaves the neighborhood. Such an equilibrium state is called an
unstable topological node or source.
We call the remaining structurally stable equilibrium states topological sad-
dles. It follows from the Grobman–Hartman theorem that a topological saddle
of the original nonlinear system has local stable and local unstable manifolds
s u
Wloc and Wloc of dimension k and (n − k), respectively. Namely, if h is a local
homeomorphism which maps the trajectories of the linearized system onto tra-
jectories of the nonlinear system (such a homeomorphism exists here by virtue
of Theorem 2.1), then the images hE s and hE u of the stable and unstable in-
variant subspaces of the linearized system are, exactly, the stable and unstable
manifolds. As in the linear case, a positive semi-trajectory starting with any
s s
point in Wloc lies entirely in Wloc and tends to O as t → +∞. Similarly, a
u u
negative semi-trajectory starting with any point of Wloc lies entirely in Wloc
s u
and tends to O as t → −∞. The trajectories of points outside of Wloc ∪ Wloc
s
escape from any neighborhood of the saddle as t → ±∞ The manifolds Wloc
u
and Wloc are invariant manifolds, i.e. they consist of whole trajectories (until
they leave some neighborhood of the topological saddle).
2.6. Leading and non-leading manifolds 65
Theorem 2.3. Two structurally stable equilibrium states are locally topologi-
cally equivalent if and only if they are of the same topological type.
ẏ = Ay (2.6.3)
66 Chapter 2. Structurally Stable Equilibrium States of Dynamical Systems
estimate (2.3.22) holds (according to Lemma 2.1) which implies that any tra-
jectory tends to the origin exponentially. Does this property of exponential
convergence of trajectories to the equilibrium state persist for the original
nonlinear system? The following theorem answers this question in the affirma-
tive.
Theorem 2.4. For a sufficiently small δ > 0 and for any y0 such that ky0 k <
δ, the trajectory y(t) of the nonlinear system (2.6.1) starting with y0 satisfies
the following inequality for all t ≥ 0:
where the positive constant ε > 0 may be chosen infinitesimally small by means
of decreasing δ, and C > 0 is some factor depending upon the choice of the basis
in Rn .
for the second term in (2.6.5). In order to obtain an estimate for the first
term 2hy, Ayi in (2.6.5) we choose the Jordan basis similar to the case when
2.6. Leading and non-leading manifolds 67
zi = λ i y i (2.6.9)
when λi+1 = λi+3 = · · · = λi+2k−1 and λi+2 = λi+4 = · · · = λi+2k are the
complex-conjugate characteristic exponents of multiplicity k. Recall that the
quantities δj in (2.6.11) and (2.6.12) may be either 0 or ε/2.
It follows from formulae (2.6.9) and (2.6.10) that if the exponents are sim-
ple, then
n
X
hy, zi = yi2 Re λi . (2.6.13)
i=1
ε ≤ Kkyk .
2.6. Leading and non-leading manifolds 69
if y decays exponentially to zero, then the value ε(t) tends to zero asymptoti-
cally as ∼ |h(y(t))|/|y(t)| ∼ 1t .
The above theorem asserts that a topologically stable node is an exponen-
tially stable equilibrium state. As we have seen in Sec. 2.3, in the linear case
the velocity and the character of convergence of the most of trajectories to the
equilibrium are determined by the leading coordinates. This feature persists in
the nonlinear case as well. Here, the role of the non-leading subspace is played
by an invariant non-leading manifold, whose existence in high-dimensional
nonlinear systems was discovered by Petrovsky.
Let us reorder the characteristic exponents so that
0 ≥ Re λ1 ≥ Re λ2 ≥ · · · ≥ Re λn .
Let m > 0 be such that the first m exponents have the same real part
Re λi = Re λ1 , (i = 1, . . . , m)
and
Re λi < Re λ1 , (i = m + 1, . . . , n) .
70 Chapter 2. Structurally Stable Equilibrium States of Dynamical Systems
y = u+v,
where u = (u1 , . . . , um ) and v = (v1 , . . . , vn−m ) are the projections onto the
leading and the non-leading eigen-subspaces E L and E ss of the matrix A, re-
spectively. With the new variables the system takes the form
u̇ = A1 u + f (u, v) ,
(2.6.16)
v̇ = A2 v + g(u, v) ,
The proof of the existence and smoothness of the non-leading manifold will
be given in Chap. 5. For now, let us prove the second part of the theorem,
namely, inequalities (2.6.18) and (2.6.19).
ss
Since Wloc is tangential to the subspace u = 0, it is defined by an equation
of the form
u = ϕ(v) , (2.6.20)
2.6. Leading and non-leading manifolds 71
where ϕ ∈ Cr and
ϕ(0) = 0 , ϕ0 (0) = 0 . (2.6.21)
ss
Since Wloc is an invariant set it follows that if u0 = ϕ(v0 ), then u(t) =
ϕ(v(t)) for t ≥ 0, and therefore
w = u − ϕ(v)
ss
so that in the new coordinates the equation of Wloc is w = 0. Such a change
of variables is called a straightening of a manifold. The system (2.6.16) is now
recast as
v̇ = A2 v + g(w + ϕ(v), v) ,
where
Z 1 Z 1
g̃ ≡ fu0 (ϕ(v) + sw, v) ds − ϕ0 (v) gu0 (ϕ(v) + sw, v) ds ∈ Cr−1 .
0 0
g̃(0, 0) = 0 .
κ(t) → 0 as t → +∞ . (2.6.28)
d µt
By (2.6.27) we obtain dt e z(t) ≤ eµt κ(t) or
Z t
z(t) ≤ z0 e−µt + e−µ(t−s) κ(s) ds .
0
whence
Z ! Z
T t
−µt µs −µ(t−s)
I(t) ≤ e e ds max κ(s) + e ds max κ(s)
0 s≥0 0 s≥T
1 µT 1
≤ e−µt e max κ(s) + max κ(s) . (2.6.29)
µ s≥0 µ s≥T
By virtue of (2.6.28), the second summand in (2.6.29) can be made to
be infinitesimally small if we choose a sufficiently large T . By choosing a
sufficiently large t, the first term in (2.6.29) can be made infinitesimally small
too. Therefore, I(t) → 0 as t → +∞.
Thus, if w0 6= 0, then kv(t)k/kw(t)k → 0, i.e. any trajectory that does
ss
not lie in Wloc , touches tangentially the leading subspace as t → +∞. In
ss
particular, this implies that Wloc is a unique m-dimensional smooth invariant
manifold which is tangential to the non-leading subspace at the point O.
ss
In order to prove estimate (2.6.19) for trajectories from Wloc , we notice that
in the restriction of the system (2.6.23)–(2.6.24) to the non-leading manifold
v̇ = A2 v + g(ϕ(v), v) (2.6.30)
node and a stable focus. The point O is called a node if m = 1, i.e. the leading
characteristic exponent λ1 is simple and real:
It is clear that if the point O is either a node or a focus, then for any matrix
close to A, it remains a node or a focus, respectively. Conversely, in the case
where neither (2.6.31) nor (2.6.32) holds, a small perturbation of the matrix
A always ensures the validity of at least one of these relations.
In the case where the equilibrium state is a node, the non-leading manifold
is (n − 1)-dimensional and it partitions a neighborhood of the point O into
ss
two components. Trajectories outside of Wloc approach O tangentially to the
w-axis along two opposite directions, from side w > 0 for the first compo-
nent, and from side w < 0 for the second. Equation (2.6.24) for the leading
coordinate w has the form
or in polar coordinates as
ṙ = (−ρ + · · · )r ,
(2.6.35)
ϕ̇ = ω + · · · ,
Fig. 2.6.1. A stable node. There is a certain hierarchy of strongly stable local manifolds.
Any trajectory converges monotonically to the node O.
Fig. 2.6.2. In contrast to Fig. 2.6.1, the loci of convergence of trajectories to the node include
an oscillating character, namely, a focus.
76 Chapter 2. Structurally Stable Equilibrium States of Dynamical Systems
ẇ = −w ,
(2.6.36)
v̇ = −2v + w 2 ,
If we take two trajectories, one from the region w > 0 and another from
w < 0, their union with O forms an invariant manifold tangential to the
leading subspace E L at O, see Fig. 2.6.4. Any such manifold can be considered
as a leading one but each has only C1 -smoothness at the point O because, by
virtue of (2.6.37),
dv −t 2v0
=e − w0 + 2w0 t ,
dw w0
whence
d2 v 2v0
2
= 2 − 2 + 2t
dw w0
hence, d2 v/dw2 → +∞ as t → +∞. In the general case the following theorem
holds (see Chap. 5 for the proof).
Here, [x] denotes the largest integer which is strictly less than x.
We have already seen in Sec. 2.5 that such an equilibrium state has the
s u
stable and the unstable invariant sets Wloc and Wloc which are locally homeo-
morphic to a k-dimensional and an (n − k)-dimensional disk, respectively. The
s u
invariant sets Wloc and Wloc intersect at only one point, namely, the equilib-
s
rium state O. If we puncture O off, both sets consist of semi-trajectories: W loc
u
is composed of positive semi-trajectories and Wloc is composed of negative
semi-trajectories.
s u
Continuation of Wloc and Wloc along trajectories outside of a neighborhood
of the saddle yields us the global stable invariant manifold W s and the global
unstable invariant manifold W u of the equilibrium state O. In the case of a
linear system they are just a k-dimensional and an (n − k)-dimensional invari-
ant subspaces of the matrix A. In the nonlinear case the manifolds W s and
W u may be embedded in Rn in a very complicated way. We will see below
how the relative location of both W s and W u in the phase space greatly af-
fects the global dynamics of the system. This is one reason why calculation
(analytical, when possible, and numerical) of these manifolds is a key element
in the qualitative study of specific systems.
We must emphasize that the results which may be obtained from the
Grobman–Hartman theorem do not allow one to determine W s or W u , or
s
to estimate their smoothness. At the same time, the local manifolds W loc and
u
Wloc are well-defined smooth objects. The existence of analytical invariant
manifolds of a saddle in analytic systems was proven by Poincaré as well as by
Lyapunov who used a different method (in terms of the so-called conditional
stability). For smooth systems Perron and Hadamard obtained similar results.
A linear non-singular change of variables transforms a nonlinear system
near a saddle equilibrium state into the following form
u̇ = A− u + f (u, v) ,
(2.7.1)
v̇ = A+ v + g(u, v) ,
where u ∈ Rk , v ∈ Rn−k , spectr A− = {λ1 , . . . , λk }, spectr A+ = {λk+1 , . . . ,
λn } and f and g are some Cr -smooth (r ≥ 1) functions which vanish at the
origin along with their first derivatives.
Theorem 2.7. A structurally stable saddle O has Cr -smooth invariant man-
s u
ifolds Wloc and Wloc (see Figs. 2.7.1 and 2.7.2) whose equations are
s
Wloc : v = ψ(u) (2.7.2)
u
Wloc : u = ϕ(v) (2.7.3)
80 Chapter 2. Structurally Stable Equilibrium States of Dynamical Systems
where
Relations (2.7.6) and (2.7.7) yield an algorithm for computing the invariant
manifolds near the saddle. First of all we expand ϕ and ψ in the Taylor series
with symbolic coefficients and then substitute them into (2.7.6) and (2.7.7)
and collect similar terms. As a result the formulae obtained allow one to
sequentially determine any number of terms in the expansions of the functions
ϕ and ψ through the Taylor coefficients of the functions f and g.
For example, for the two-dimensional analytic system
X
u̇ = −λu + αij ui v j ,
i+j≥2
X
v̇ = γv + βij ui v j
i+j≥2
ψ(u) = ψ2 u2 + ψ3 u3 + · · ·
82 Chapter 2. Structurally Stable Equilibrium States of Dynamical Systems
with indefinite coefficients ψi into this equation and equate the coefficients of
u2 , we obtain
γψ2 + β20 = −2λψ2 ,
hence we find
ψ2 = −β20 /(2λ + γ) .
Having equated coefficients of u3 , we obtain
yielding
ψ3 = −(β30 + β11 ψ2 − 2ψ2 α20 )/(3λ + γ) .
Repeating this procedure, we find step by step all of the coefficients of the
Taylor expansion of the function ψ. The formula for calculating the m-th
coefficient is the following:
ξ = u − ψ(v) ,
η = v − ϕ(u) .
The functions h1,2 are small near the saddle and therefore, as long as the
trajectory remains in a neighborhood of the saddle, the inequalities
d
kξ(t)k ≤ (Re λ1 + ε)kξ(t)k for t ≥ 0 ,
dt
d
kη(t)k ≤ (Re γ1 − ε)kη(t)k for t ≤ 0
dt
hold in the Jordan basis (see the proof of Theorem 2.4). After integrating we
obtain
v̇ = A+ v + h2 (ϕ(v), v) . (2.7.13)
s u
• saddle : a node on Wloc and on Wloc ;
s u
• saddle-focus (2,1) : a focus on Wloc and a node on Wloc ;
s u
• saddle-focus (1,2) : a node on Wloc and a focus on Wloc ;
s u
• saddle-focus (2,2) : a focus on Wloc and a focus on Wloc .
Theorems 2.5 and 2.6 hold for the systems (2.7.12) and (2.7.13). This im-
plies that there exist a non-leading and a leading stable invariant sub-manifolds
ss sL s uu uL u
Wloc , Wloc in Wloc and unstable invariant sub-manifolds Wloc , Wloc in Wloc .
In addition, we will close this section by describing the existence of three more
smooth invariant manifolds of the saddle equilibrium state. Let us introduce
the notation
" #
λ̂
rsL = , (2.7.14)
Re λ1
γ̂
ruL = , (2.7.15)
Re γ1
where λ̂ and γ̂ are the real parts of the non-leading stable and unstable ex-
ponents nearest to the imaginary axis, respectively, and where [x], as before,
denotes the largest integer which is strictly less than x.
Theorem 2.8. In a small neighborhood of a structurally stable equilibrium
state of saddle type there exist the following smooth invariant manifolds:
• a Cmin (r,ruL ) -smooth extended stable manifold Wloc
sE s
which contains Wloc
and which is tangential at the point 0 to the direct sum of the stable and
sE
the leading unstable eigen-space of the linearization matrix (thus, W loc
uu
is transverse to Wloc );
• a Cmin (r,rsL ) -smooth extended unstable manifold Wloc
eE
which contains
u
Wloc and which is tangential at the point 0 to the direct sum of the
unstable and the leading stable eigen-space of the linearization matrix
ss
(it is transverse to Wloc );
• a Cmin (r,rsL ,ruL ) -smooth leading saddle manifold Wloc
L uE
= Wloc sE
∩ Wloc .
The proof of this theorem is given in Chap. 5. We notice merely that
sE
the manifold Wloc , generally speaking, is not unique, however any two such
s
manifolds have the same tangent at each point of Wloc . Similarly, any two
uE u
manifolds Wloc are tangential everywhere on Wloc .
2.8. Solution near a saddle. The boundary-value problem 85
Of course, the same problem also arises near a completely unstable equi-
librium point. However, in this case, a solution of the initial-value problem
becomes stable upon reversing the time variable. In the case of a saddle,
however, after the change t → −t a solution becomes stable with respect to
the variables v, but unstable with respect to the variables u.
The principal idea for overcoming these obstacles is to integrate the system
with respect to the variables u in forward time, and with respect to v in
backward time. More specifically, instead of solving an initial-value problem,
we must solve the following boundary value problem:
given any u0 and v1 , where ku0 k ≤ ε and kv1 k ≤ ε, and any τ > 0, find a
solution of the system (2.8.1) on the interval t ∈ [0, τ ] such that
For linear systems the solution of the boundary value problem has the form
− +
u(t) = eA t
u0 , v(t) = e−A (τ −t)
v1 . (2.8.3)
− +
As both keA t k and ke−A (τ −t) k are bounded for all t ∈ [0, τ ] (see inequality
(2.3.35)), this solution is stable with respect to perturbations of the initial
values (u0 , v1 , τ ). This technique can be applied in the nonlinear case as well.
As we will see, a solution of the boundary value problem of a nonlinear system
can be obtained by the method of successive approximations, starting with
solution (2.8.3) of the linear problem.
Theorem 2.9. For sufficiently small ε > 0 and any τ ≥ 0, and u0 and v1
such that ku0 k ≤ ε, kv1 k ≤ ε a solution of the boundary value problem (2.8.2)
exists, is unique and depends continuously on (u0 , v1 , τ ).
v1
Fig. 2.8.1. The boundary value problem near a saddle. The flight time τ → +∞ as a nearby
initial point on u = u0 tends to the stable manifold v = 0.
with respect to the functions u(t) and v(t) on the interval t ∈ [0, τ ]. By
differentiating the right-hand side of formula (2.8.4) with respect to t one can
easily verify that any continuous solution {u(t), v(t)} of this system is a solution
of the system (2.8.1). Moreover, as u(0) = u0 and v(τ ) = v1 , the solution of the
system (2.8.4) is the desired solution of the boundary value problem (2.8.2).
The converse is also true. Let {u(t), v(t)}t∈[0,τ ] be a solution of the bound-
ary value problem. It follows from (2.8.1) that
d −A− t −
(e u(t)) = e−A t f (u(t), v(t)) ,
dt
d −A+ t +
(e v(t)) = e−A t g(u(t), v(t)) ,
dt
whence
Z t
−A− t −
e u(t) = u0 + e−A s f (u(s, v(s)) ds ,
0
Z τ
−A+ t + +
e v(t) + e−A s g(u(s), v(s)) ds = e−A τ v1 ,
t
88 Chapter 2. Structurally Stable Equilibrium States of Dynamical Systems
and, consequently, (2.8.4) is valid. Thus, the solution of the boundary value
problem is identical to that of the system of the integral equations (2.8.4).
We will construct a solution of system (2.8.4) by the method of successive
approximations. As the first approximation we choose
− +
u(1) (t) = eA t u0 , v (1) (t) = e−A (τ −t)
v1 ,
We will show that this sequence converges uniformly to some limit function
{u∗ (t), v ∗ (t)}. First, let us prove that
for all n and t ∈ [0, τ ]. When n = 1 it follows directly from ku0 k ≤ ε and
kv1 k ≤ ε as well as from the inequalities
−
keA t k ≤ e−λt ,
+
(2.8.7)
ke−A (τ −t)
k ≤ e−γ(τ −t) ,
where λ > 0 and γ > 0 are such that the spectrum of the matrix A− lies
strictly to the left of the straight-line Re z = −λ in the complex plane, and
the spectrum of the matrix A+ lies strictly to the right of the straight-line
Re z = γ.
We will use mathematical induction to prove inequality (2.8.6) for all n.
Firstly, observe that since both f and g vanish at the point O along with their
first derivatives, it follows that
∂(f, g)
≤ δ, (2.8.8)
∂(u, v)
kf, gk ≤ δku, vk , (2.8.9)
where ku, vk denotes max{kuk, kvk}. In this equation, the constant δ can be
made arbitrarily small by decreasing the size of the neighborhood of the point
2.8. Solution near a saddle. The boundary-value problem 89
O. Choose ε small such that for all u and v from the 2ε-neighborhood of the
saddle, the inequality
2δ max (λ−1 , γ −1 ) ≤ 1 (2.8.10)
is satisfied. From (2.8.5) and (2.8.9) we obtain
Z t
ku(n+1) (t)k ≤ ku0 k + δ e−λ(t−s) ku(n) (s), v (n) (s)k ds ,
0
Z τ
kv (n+1) (t)k ≤ kv1 k + δ e−γ(s−t) ku(n) (s), v (n) (s)k ds ,
t
whence
ku(n+1) (t), v (n+1) (t)k ≤ ε + δ max (λ−1 , γ −1 ) max ku(n) (s), v (n) (s)k .
0≤s≤τ
then
ku(n+1) (t), v (n+1) (t)k ≤ 2ε ,
i.e. (2.8.6) holds for all n. Let us now prove that
1
≤ max ku(n) (s) − u(n−1) (s), v (n) (s) − v (n−1) (s)k . (2.8.11)
2 0≤s≤τ
Indeed, by (2.8.5)
Analogously,
Since the values u(n) and v (n) lie inside the 2ε-neighborhood of the saddle
for all n, the value δ in the last two inequalities satisfies (2.8.10), and hence
inequality (2.8.11) follows.
By virtue of (2.8.11) the series
∞
X
(u(n+1) (t) − u(n) (t), v (n+1) (t) − v (n) (t))
n=1
Proof. Let {u∗ (t), v ∗ (t)}t∈[0,τ ] be a solution of the boundary value prob-
lem corresponding to (u0 , v1 , τ ). Denote v0 = v ∗ (0). The trajectory {u∗ , v ∗ }
depends Cr -smoothly on (u0 , v0 , t, τ ) as it is a solution of the initial-value prob-
lem. Therefore to prove the theorem we must show that v0 depends smoothly
on (u0 , v1 , t, τ ). Since v1 = v ∗ (t = τ ) is a smooth function with respect to
(u0 , v0 , t, τ ), it is sufficient (according to the implicit function theorem) to
verify that the derivative ∂v1 /∂v0 = ∂v ∗ /∂v0 |t=τ is non-singular.
The derivatives ∂u∗ /∂v0 and ∂v ∗ /∂v0 can be found as solutions of the
system of variational equations
V (τ ) Q = Im . (2.8.14)
Equations (2.8.12) are linear with respect to the variables U and V , therefore,
if we postmultiply both the right and the left-hand sides of (2.8.12) by Q, it
is easily seen that Ũ ≡ U Q and Ṽ ≡ V Q also satisfy Eqs. (2.8.12). Moreover,
92 Chapter 2. Structurally Stable Equilibrium States of Dynamical Systems
in order that both satisfy (2.8.13) and (2.8.14) the following conditions must
hold
Ũ (0) ≡ 0 , Ṽ (τ ) ≡ Im . (2.8.16)
Thus, the derivative matrix ∂v1 /∂v0 is non-singular if and only if the boundary-
value problem (2.8.16) for the system of variational equations has a solution.
To complete the proof of the theorem we note that the existence and
uniqueness of a solution of boundary value-problem (2.8.16) for the system
of variational equations (2.8.12) follows from the remark to the previous theo-
rem. Moreover, since the derivatives of the right-hand side of system (2.8.12)
with respect to U and V are (without the constant matrices A+ and A− )
f 0 (u∗ (t), v ∗ (t)) and g 0 (u∗ (t), v ∗ (t)), respectively, it follows that they can be
estimated by the same constant δ as the derivatives of the nonlinear part of
the right-hand side of system (2.8.1) with respect to u and v. The boundary
value problem for the system of variational equations is therefore solvable for
ku0 , v1 k ≤ ε where ε may be taken the same as in the boundary value problem
for the original system.
As it is seen from the proof, the derivative of the solution (u∗ (t), v ∗ (t)) of
the boundary-value problem for system (2.8.1) with respect to v1 is given by
−1
∂(u∗ (t), v ∗ (t)) ∂(u∗ (t), v ∗ (t)) ∂v1
=
∂v1 ∂v0 ∂v0
= (U (t), V (t))V (τ )−1 = (Ũ (t), Ṽ (t)) ,
U (0) = In , V (τ ) = 0 . (2.8.17)
U (0) = 0 , V (τ ) = 0 (2.8.18)
2.8. Solution near a saddle. The boundary-value problem 93
Ż = A− Z + fu0 Z + fv0 W
00
+fuu U ∗ U ∗ + fuv
00
V ∗ U ∗ + fvu
00
U ∗ V ∗ + fvv
00
V ∗V ∗ ,
Ẇ = A+ W + gu0 Z + gv0 W
00
+guu U ∗ U ∗ + guv
00
V ∗ U ∗ + gvu
00
U ∗ V ∗ + gvv
00
V ∗V ∗ ,
Z(0) = 0 , W (τ ) = 0 ,
where we have introduced the notation
and
Z ≡ ∂ 2 (u∗ )/∂(v1 )2 , W ≡ ∂ 2 (v ∗ )/∂(v1 )2 .
Let us show how this theory can be applied to prove the existence and
smoothness of the stable and the unstable manifolds (Theorem 2.7 of the pre-
vious section). We will consider only the case of the stable manifold W s ; as
for the unstable manifold W u one repeats the method for the system obtained
from (2.8.1) by reversing time.
Note that our conclusions concerning the uniform convergence of succes-
sive approximations (2.8.5) and, consequently, successive approximations for
boundary value problems for variational equations (2.8.12) as well as for non-
homogeneous variational equations (2.8.19) and for the variational equations
for the variational equations etc., remain valid for all τ ≥ 0 including τ = +∞.
Thus, the system
Z t
− −
u(t) = eA t u0 + eA (t−s) f (u(s), v(s)) ds ,
0
Z (2.8.21)
∞
−A+ (s−t)
v(t) = − e g(u(s), v(s)) ds
t
obtained from (2.8.4) with τ = +∞, for all ku0 k ≤ ε has a unique solution
which lies in the 2ε-neighborhood of the point O for t ≥ 0. Moreover, this
solution depends Cr -smoothly on u0 .
It is easy to verify by direct differentiation that a solution of the system
(2.8.21) is also a solution of the system (2.8.1).
2.9. Problem of smooth linearization. Resonances 95
Conversely, it follows from the proof of Theorem 2.9 that for any solution
{u(t), v(t)} of system (2.8.1) which stays in a small neighborhood of O for all
t ≥ 0, relation (2.8.4) holds with u0 = u(0) and v1 = v(τ ) satisfying for any
τ ≥ 0. Therefore, taking the limit in relation (2.8.4) we see that any bounded
solution of system (2.8.1) satisfies the system of integral equations (2.8.21).
Thus, for any u0 satisfying ku0 k ≤ ε, there exists a unique v0 such that
the trajectory beginning with (u0 , v0 ) does not leave a neighborhood of the
saddle as t → +∞. Let us denote v0 = ψ(u0 ). By definition, the union of all
points (u0 , ψ(u0 )) is an invariant set of the system (2.8.1): It consists of all
trajectories which remain in a neighborhood of the equilibrium state O for all
t ≥ 0. In particular, this set contains the point O itself; i.e. ψ(0) = 0. Because
ψ ∈ Cr , this set is a smooth invariant manifold. The system on this manifold
is written as
u̇ = A− u + f (u, ψ(u)) .
Since the derivatives fu0 and fv0 vanish at the origin, the linearized equation is
u̇ = A− u .
ẋ = Ax + f (x) , (2.9.1)
where
f (0) = 0, f 0 (0) = 0 ,
96 Chapter 2. Structurally Stable Equilibrium States of Dynamical Systems
ẏ = Ay . (2.9.2)
Let us now ask a very natural question: Can system (2.9.1) be reduced to
system (2.9.2) by some smooth change of variables
y = x + ϕ(x) , (2.9.3)
where ϕ(0) = 0 and ϕ0 (0) = 0? Poincaré was the first who posed this question
and considered it in the analytic case. We remark that a smooth change of
variables preserves the eigenvalues (λ1 , . . . , λn ) of the matrix A and, moreover,
when it is locally close to identity like (2.9.3) it preserves the matrix A itself.
Observe that such a change of variables is local, i.e. the smooth equivalence is
assumed to be valid only in a small neighborhood of the equilibrium state O.
While reducing the original nonlinear system to linear form we run into
a number of difficulties, and the main one is caused by the presence of
resonances.
The set {λ1 , . . . , λn } of the eigenvalues (λ1 , . . . , λn ) of the matrix A is called
a resonant set if there exists a linear relationship
n
X
λk = (m, λ) = mj λ j , (2.9.4)
j=1
where the last summand denoted by the ellipsis contains terms of degree N + 1
and higher. The remaining terms (except for Ay) must cancel each other,
therefore one must assume that ϕ2 (x) satisfies the equation
∂ϕ2 (x)
−Aϕ2 (x) + f2 (x) + Ax = 0 ; (2.9.9)
∂x
ϕ3 (x) satisfies
∂ϕ3 (x) ∂ϕ2 (x)
−Aϕ3 (x) + f3 (x) + Ax + f2 (x) = 0 ; (2.9.10)
∂x ∂x
.. .. .. ..
. . . .
ϕN (x) satisfies
∂ϕN (x) X ∂ϕp (x)
−AϕN (x) + fN (x) + Ax + fq (x) = 0 . (2.9.11)
∂x ∂x
p+q=N +1
Let us now prove the lemma for the case where the matrix A is diagonal
λ1 0
A=
..
.
.
0 λn
98 Chapter 2. Structurally Stable Equilibrium States of Dynamical Systems
Represent the polynomials ϕlk (x) and flk (x) in the form
P
ϕlk (x) = cmk xm ,
m1 +···+mn =l
P
flk (x) = dmk xm .
m1 +···+mn =l
n n
X ∂ϕ3k (x) X ∂ϕ2k (x)
−λk ϕ3k (x) + λj xj + f3k (x) + f2j (x) = 0 , (2.9.13)
j=1
∂xj j=1
∂xj
.. .. .. ..
. . . .
n n
X ∂ϕN k (x) X X ∂ϕpk (x)
−λk ϕN k (x) + λj xj + fN k (x) + fqj (x) = 0 ,
j=1
∂xj j=1
∂xj
p+q=N +1
(2.9.14)
where k = 1, 2, . . . , n.
First solve (2.9.12). Equating the coefficients of the similar terms we obtain
the equation
[(m, λ) − λk ]cmk + dmk = 0 . (2.9.15)
It is clear that this equation may be resolved since there are no resonances,
i.e. we can find
dmk
cmk = , (2.9.16)
λk − (m, λ)
and, consequently, the expression for ϕ2 (x).
Substituting ϕ2 (x) into Eq. (2.9.13) we obtain the equation for unknown
coefficients of ϕ3 (x)
[(m, λ) − λk ]cmk + d˜mk = 0 (2.9.17)
2.9. Problem of smooth linearization. Resonances 99
with d˜mk = dmk + d0mk , where d0mk is the coefficient of xm in the second sum in
(2.9.13). Proceeding in the analogous manner we obtain all ϕl (l = 2, . . . , N )
satisfying Eqs. (2.9.9)–(2.9.11).
This proves the lemma when all eigenvalues of A are real and different
(because A is brought to the diagonal form by a linear change of variables in
this case). If there are simple complex eigenvalues, then A is brought to a
block-diagonal form:
Akk = λk if λk is real
Akk = Ak+1,k+1 = Re λk , Ak,k+1 = −Ak+1,k = Im λk if λk = λ∗k+1
where ∗ denotes complex conjugation. This matrix is made diagonal by a
complex coordinate transformation
x0k = xk if λk is real
xk = xk + ixk+1 , x0k+1 = xk − ixk+1 if λk = λ∗k+1 .
0
Ak,k+1 = δk .
may appear in (2.9.12)–(2.9.14). Obviously, this does not change the conclusion
of the lemma: relation (2.9.17), by which the sought coefficients cmk are defined
inductively, remains the same with the only difference that d0mk is expressed
now in terms of a wider range of coefficients cm0 k0 . Namely, d0mk is a function
of such cm0 k0 that:
(1) |m0 | < |m|, or (2) m0 = m and k 0 > k,
n n
X X (2.9.18)
or (3) |m0 | = |m| and j · m0j < j · mj
j=1 j=1
(only case 1) is possible when the eigenvalues are simple). Thus, one may
introduce the partial order among the vector-monomials xm ek (where ek =
(0, 0, . . . , 1, 0, . . . , 0)) in the following way:
| {z }
k
0
xm ek is of a higher order than xm ek0 if one
of the three options of (2.9.18) holds.
Formula (2.9.17) then allows for determining the coefficients cmk sequentially,
for monomials of higher and higher orders. End of the proof.
It is immediately seen from (2.9.15) that in the case of resonance λk =
(m, λ) it is impossible to eliminate monomials dmk xm ek . Thus, a sufficiently
smooth linearization is impossible in the resonant case. In the same way as
Lemma 2.2, the following result may be proved.
Lemma 2.3. Let f (x) ∈ CN . Then, there exists a polynomial change of vari-
ables which transforms system (2.9.1) into
with
2≤|m|≤N
X
R(y) = bmk y m ek , (2.9.20)
(m,λ)=λk
where ek is the k-th basis vector, and the coefficient bmk of the resonant mono-
mial y m ek is found in terms of the coefficients of the polynomials fl (x) with
l ≤ |m|.
Therefore, the two following theorems are only concerned with the formal
series.
Let us discuss next the question of the convergence of these series. Following
Arnold [10], we introduce a few useful preliminary notions.
Consider a complex n-dimensional space Cn . The set
n
X
λk = (m, λ) , mj ≥ 2 , mj ≥ 0 ,
j=1
Each point in the Poincaré region satisfies at most a finite number of res-
onances and lies in such a neighborhood which has no intersection with other
resonant planes. In contrast, the resonant planes are dense in the Siegel region.
Poincaré proved this theorem using majorant series. The principal require-
ment in his formulation is the condition of the existence of a straight-line
passing through the origin in the complex plane such that all n eigenvalues
(λ1 , . . . , λn ) lie to one side of this line.
In the case where there are resonances in the Poincaré region, the formal
series which defines the change of variables in Theorem 2.12 is convergent.
Therefore, the following result holds.
ẏ = Ay + R(y) ,
The situation outside the Poincaré region is much more complicated. Spe-
cific conditions on eigenvalues under which the linearizing series converges and
system (2.9.1) can still be reduced to linear form were found by Siegel. It is
important to note that eigenvalues satisfying these conditions compose a dense
set of positive measure.
Poincaré and Dulac considered complex analytical systems. We are inter-
ested in the real case. In the real case, the Poincaré region is determined by
the conditions Re λi < 0 or Re λi > 0 (i = 1, . . . , n), i.e. where the equilibrium
state is stable or completely unstable, respectively. If there are eigenvalues on
the imaginary axis, or both in the left and right half-planes, then the system
falls in the Siegel region. For example, suppose a two-dimensional system has
a saddle equilibrium state with the eigenvalues λ1 < 0 < λ2 . If the saddle
p
index ν = −λ1 λ−1 2 is rational (ν = q ), there is an infinite set of resonances of
the type
λ1 = (rq + 1)λ1 + prλ2 ,
λ2 = qrλ1 + (pr + 1)λ2 , r = 1, 2, . . . .
If the equilibrium state is of the saddle type, then even when the collection
{λ1 , . . . , λn } is not resonant, zero is a limit point of the set
{(m, λ) − λk }∞
|m|=2 , k = 1, . . . , n .
This is a reason why the linearizing series may not converge even provided that
there are no resonances (see Bruno [18]).
The situation becomes less difficult if we do not require that the system be
analytic, but C∞-smooth. In this case the following statement holds
Lemma 2.4. (Borel (see Hartman)) For any formal power series R(y)
there exists a C∞-smooth function whose formal Taylor series coincides with
R(y).
Theorem 2.15. (Sternberg) Let the system (2.9.1) be C∞-smooth and let it
have no resonances, then there exists a C∞-smooth change of variables which
brings (2.9.1) to linear form.
ẏ = Ay + R(y) ,
where R(y) ∈ C∞ and its formal Taylor series coincides with the formal series
(2.9.23).
ẏ = Ay
who proved the existence of a number K(k), depending on the spectrum of the
matrix A, such that any C∞-system
ẋ = Ax + oK (x)
a weak resonance. Observe that the notion of the weak resonance does not
employ analysis in the complex plane and, consequently, the reduction of the
linear part to the Jordan form. In contrast to the classical notion of reso-
nances being an obstacle in the linearization by polynomial transformations,
the notion of weak resonances arises in the problem of the reduction of systems
of a finite smoothness to linear form when we use wider classes of changes of
variables.
Of primary interest from the viewpoint of nonlinear dynamics are saddles.
The reason is because a saddle may have bi-asymptotical trajectories which
belong to both the stable and the unstable manifold. Such trajectories are
called homoclinic loops. In the case where an equilibrium state is a saddle-
focus, infinitely many periodic trajectories may arise from one homoclinic loop
2.9. Problem of smooth linearization. Resonances 105
under certain conditions. The study of such phenomena begins with the re-
duction of the system near a saddle to a simpler form. It is obvious that the
case where the system is reducible to linear form would be ideal. However,
the study of global bifurcations requires the consideration of finite-parameter
families of systems rather then an individual system. Reduction to linear form
is difficult for the family of systems because resonances are dense in the Siegel
regions. At the same time, most resonances, with a few exceptions, do not
play an essential role in the study of homoclinic bifurcations.
Let us discuss in detail an example of a planar system which was studied
by Andronov and Leontovich. Consider a one-parameter family of systems
ẋ = λ1 (µ)x + P (x, y, µ) ,
ẏ = λ2 (µ)y + Q(x, y, µ) ,
where P and Q are CN (N ≥ 1)-smooth functions vanishing at the origin along
with their first derivatives with respect to x and y, and λ1 (0) < 0 < λ2 (0).
Suppose that when µ = 0 this system has a separatrix loop, see Fig. 2.9.1.
Assume also that the saddle index ν = −λ1 (µ)/λ2 (µ) differs from 1 when
µ = 0, i.e. the so-called saddle value σ is non-zero:
σ(0) = λ1 (0) + λ2 (0) 6= 0 .
Andronov and Leontovich showed that under these conditions at most one
periodic trajectory can arise from the separatrix loop. The condition
λ1 (0) = (p + 1)λ1 (0) + pλ2 (0), λ2 (0) = qλ1 (0) + (q + 1)λ2 (0),
∞
X
ẏ = λ2 (0)y + bp xp y p+1 .
p=1
When µ 6= 0 it does not make sense to apply the theory of normal forms because
the dependence on the parameter is discontinuous on a dense set since the
saddle index ν(µ) may be rational or irrational. Nevertheless, Leontovich [40]
was able to transform the family into
" K−1
#
dy y X
p K
= 1+ cp (µ)(xy) + (xy) Φ(x, y, µ) . (2.9.25)
dx x p=1
The main method which she employed was that of sequentially eliminating
the non-resonant functions, i.e. such whose all terms in the formal Taylor
expansion are non-resonant. The procedure works for a CN-smooth family
provided that N ≥ 4K + 1. The function Φ(x, y, µ) in Eq. (2.9.25) is then of
smoothness CK .
2.9. Problem of smooth linearization. Resonances 107
ẋ = A1 (µ)x + f1 (x, y, u, v, µ) ,
u̇ = A2 (µ)u + f2 (x, y, u, v, µ) ,
(2.9.26)
ẏ = B1 (µ)y + g1 (x, y, u, v, µ) ,
v̇ = B2 (µ)v + g2 (x, y, u, v, µ) ,
lie to the left of the imaginary axis in the complex plane and those of the
matrix B(0)
B1 (0) 0
B(0) ≡
0 B2 (0)
lie to the right of the imaginary axis.
We assume also that the real parts of the eigenvalues (λ1 , . . . , λm1 ) of the
matrix A1 (0) are equal, i.e.
Re λ1 = · · · = Re λm1 = λ < 0 ,
and that the real parts of the eigenvalues (γ1 , . . . , γn1 ) of the matrix B1 (0) are
also equal i.e.
Re γ1 = · · · = Re γn1 = γ > 0 .
Regarding the eigenvalues of the matrices A2 (0) and B2 (0) we assume that the
real parts of the eigenvalues of A2 (0) are strictly smaller than λ, and those of
B2 (0) are strictly larger than γ. In this case the coordinates x and y are leading
stable and unstable, respectively, and the coordinates u and v are non-leading.
3 The same approach was applied near a saddle periodic orbit by Gonchenko and Shilnikov
[27]; for other applications see the papers by Afraimovich and by Lerman and Umanskii in
Methods of the Qualitative Theory of Differential Equations, edited by Leontovich, Gorky
State University, Gorky, 1984.
108 Chapter 2. Structurally Stable Equilibrium States of Dynamical Systems
where
fij |(x,u,y,v)=0 = 0 gij |(x,u,y,v)=0 = 0
f1j |(y,v)=0 = 0 g1j |(x,u)=0 = 0 (2.9.28)
fi1 |x=0 = 0 gi1 |y=0 = 0 (i, j = 1, 2) .
Since we will frequently use this theorem while studying homoclinic bifur-
cations, its complete proof is given in Appendix A. The smoothness of the coor-
dinate transformation and the functions fij , gij is defined as follows: It is Cr−1
with respect to (x, u, y, v) and the first derivatives with respect to (x, u, y, v)
are Cr−2 with respect to (x, u, y, v, µ). If r = ∞, then the transformation is
C∞ with respect to (x, y, u, v) (or even analytical in the real analytical case).
Nevertheless, even when r = ∞ the smoothness of the transformation with re-
spect to the parameters µ is, generically, only finite: It grows unboundedly as
µ0 → 0 (where kµk ≤ µ0 is the range of parameter values under consideration).
Let us consider next the case where some eigenvalues of the matrix A in
(2.9.1) lie on the imaginary axis. It is obvious that if there is one zero eigenvalue
λ1 = 0, then there exists an infinite set of resonances of the type:
λ1 = mλ1 , m ≥ 2. (2.9.29)
λ1 = (s1 + 1)λ1 + s1 λ2 ,
(2.9.30)
λ2 = s2 λ1 + (s2 + 1)λ2 ,
where R1 (µ) is a polynomial of degree not higher than (N − 1), R1 (0) = 0 and
|m|≤N
X
RN (y, µ) = bmk (µ)y m ek , (2.9.38)
(m,λ)=λk
where bmk (µ) are certain polynomials of degree not exceeding (N − |m|).
If the matrix A is non-degenerate, then R0 (µ) ≡ 0. Otherwise, when among
the eigenvalues (λ1 , . . . , λp ) there is λk = 0, R0 (µ) is a polynomial of degree
N such that R0 (0) = 0. The appearance of the term R0 (µ) in (2.9.37) is due
to the existence of resonances of the kind
0 = λk = (l, γ) .
The family
|m|≤N
X
ẏ = Ay + R0 (µ) + R1 (µ)y + bmk (µ)y m ek (2.9.39)
(m,λ)=λk
is called a truncated or shortened normal form. In many cases one may try
to restrict the consideration of the behavior of trajectories in a small fixed
neighborhood of an equilibrium state, as well as the study of the bifurcation
unfolding for small values of the control parameters by the investigation of the
truncated normal form for a suitable choice of N and q. Of course, the infor-
mation obtained from the analysis of the truncated system must be justified
before it is applied to the original family. Following this scheme we will carry
out the study of principal local bifurcations of equilibrium states in the second
part of this book.
Chapter 3
STRUCTURALLY STABLE
PERIODIC TRAJECTORIES OF
DYNAMICAL SYSTEMS
111
112 Chapter 3. Structurally Stable Periodic Trajectories
polynomial right-hand side can be estimated exactly, but the estimation of the
number of periodic trajectories of even just a planar system is the subject of
the famous Hilbert’s 16-th problem which is still unsolved. One exception is,
perhaps, nearly integrable two-dimensional systems for which the problem of
finding the periodic trajectories is reduced to that of finding the zeros of some
special integrals which can be explicitly calculated in certain specific cases.
In this chapter we will examine the behavior of trajectories near a struc-
turally stable periodic trajectory. The main idea of the study is based on
constructing the Poincaré return map.
ẋ = X(x) , (3.1.1)
Since the periodic trajectory L passes through the origin, this equation has a
solution t = τ for x0 = 0, where τ is the period of the periodic trajectory. By
virtue of (3.1.2) we may apply the implicit function theorem to Eq. (3.1.3),
whence the return time t(x0 ) is uniquely defined. Moreover, the function t(x0 )
has the same smoothness as the original system.
The map T may be written in the following form
x̄k = ϕk (t(x), x)
or as
x̄ = f (x) , (3.1.4)
where x is an n-dimensional vector of the coordinates on the cross-section S,
f (x) ∈ Cr .
If, at t = 0, we let a trajectory flow out from the point M̄ on S in backward
time, then it must return to S at the point M over the time interval t(M ).
Thus, the map T −1 , the inverse of the Poincaré map, is also defined on the
cross-section S. Because the property of smooth dependence of the return time
on the initial point persists in backward time, we can assert that the inverse
114 Chapter 3. Structurally Stable Periodic Trajectories
map T −1 also belongs to the class Cr . This implies that the Poincaré map is
a Cr -smooth diffeomorphism.
If we write down
x = f −1 (f (x)) ,
then after differentiating we obtain
[f −1 (f (x))]0 · f 0 (x) = I ,
x̄ = Ax . (3.1.6)
We will see below that just like the characteristic exponents of an equilib-
rium state, the key role in determining the dynamics of a Poincaré map near
3.2. Non-degenerate linear one- and two-dimensional maps 115
x = By + ψ(y) ,
where det B 6= 0, ψ(0) = 0 and ψ 0 (0) = 0, so that it does not move the origin,
then in the new coordinates the map (3.1.5) becomes
or
ȳ = B −1 ABy + · · · ,
where the ellipsis denotes nonlinear terms. Because the matrix B −1 AB is
similar to the matrix A, they have the same eigenvalues.
It is also obvious that the multipliers of the Poincaré map depend neither
on the choice of the point M ∗ on L nor on the particular choice of S with
respect to the periodic trajectory. Since the flight time from one transverse
cross-section to another depends smoothly on the initial point, the map of
one cross-section onto another cross-section along trajectories of the system is
a Cr -diffeomorphism, and, therefore, the change of the cross-section may be
simply considered as a change of coordinates.
In the following sections we will examine the behavior of the trajecto-
ries of dynamical systems near structurally stable (rough) periodic trajectories,
i.e. those which have no multipliers equal to 1 in absolute value. We shall be-
gin with the study of structurally stable fixed points of the Poincaré map. We
remark here that the theory of fixed points amounts to a partial, though not
absolute, repetition of the theory of equilibrium states. We therefore pursue
our study by following the same scheme as in Chap. 2: the linear case followed
by the nonlinear case and, by the correspondence between nonlinear and linear
maps.
In the present and the consequent sections we will study linear maps. We are
interested in the linearization of the Poincaré map near a periodic trajectory,
116 Chapter 3. Structurally Stable Periodic Trajectories
x̄ = Ax , det A 6= 0 .
Let us start with a linear map of dimension one. It is written in the form
x̄ = ρx , (3.2.1)
where ρ 6= 0, x ∈ R1 .
It is easy to see that the fixed point at the origin O is stable when |ρ| < 1.
The iterations xj of a point x0 are given by the formula
xj = ρ j x0 ,
where
lim xj = 0
j→+∞
if |ρ| < 1.
On the other hand, the fixed point is unstable when |ρ| > 1.
The behavior of the iterations of points in the one-dimensional case is con-
veniently interpreted by means of a Lamerey diagram which is constructed as
follows. For the map
x̄ = f (x)
the graph of the function f (x) and the bisectrix1 x̄ = x are drawn in the plane
(x, x̄). Trajectories are represented as polygonal lines: let {xj } be a trajectory;
each point with coordinates (xj , xj+1 ) lies on the graph f (x) while each point
(xj , xj ) lies on the bisectrix x̄ = x. Each point (xj , xj ) is connected vertically
with the subsequent point (xj , xj+1 ), which in turn is connected horizontally
with the subsequent point (xj+1 , xj+1 ) and so on. This process is iterated
repeatedly, as shown in the four typical Lamerey diagrams in Figs. 3.2.1 to
3.2.4.
When the function f (x) increases monotonically, then the construction ob-
tained is called a Lamerey stair (Figs. 3.2.1 and 3.2.2). When f (x) decreases
monotonically the construction is called a Lamerey spiral, see Figs. 3.2.3 and
3.2.4.
1 The 45◦ line.
3.2. Non-degenerate linear one- and two-dimensional maps 117
Fig. 3.2.1. A Lamerey stair. The origin is a stable fixed point: all points in its neighborhood
converge to O.
Fig. 3.2.2. A Lamerey stair where the origin is an unstable fixed point.
118 Chapter 3. Structurally Stable Periodic Trajectories
Fig. 3.2.3. An example of a Lamerey spiral; a trajectory starting from x 0 looks like a
clockwise-right-angled spiral.
2 1
Fig. 3.2.5. A stable node (+) with positive multipliers 0 < ρ2 < ρ1 < 1. A trajectory {T i M }
of the point M enters the origin tangentially to the leading direction x.
1 2 2 1
Fig. 3.2.6. A stable node “−” with the negative leading multiplier ρ 1 ; therefore, the x-
coordinate changes its sign after each iteration.
3.2. Non-degenerate linear one- and two-dimensional maps 121
(a)
Fig. 3.2.7(a). A saddle (+,+). The y-axis coincides with the unstable direction, the x-axis
with the stable one.
(b)
Fig. 3.2.7(b). A saddle (+, −). The sign of the y-coordinate of the trajectory {T i M } changes
after each iteration.
122 Chapter 3. Structurally Stable Periodic Trajectories
(c)
Fig. 3.2.7(c). A saddle (−, +). The “jumping” direction is here the x-axis because the
corresponding multiplier ρ1 is negative.
(d)
Fig. 3.2.7(d). A saddle (−, −). The trajectory of the initial point M runs away from the
origin along the hyperbolas located in 1st and 3rd quadrants.
3.2. Non-degenerate linear one- and two-dimensional maps 123
r̄ = ρ r ,
(3.2.5)
ϕ̄ = ϕ + ω .
rj = ρ j r0 ,
ϕj = ϕ 0 + ω j .
When ρ < 1, the point O is called a stable focus. In this case, all trajectories
lie on logarithmic spirals winding into the origin as shown in Fig. 3.2.8.
When ρ > 1, the fixed point O is called an unstable focus. In this case, all
trajectories diverge from any neighborhood of the point O as j → +∞.
In the degenerate case where ρ = 1, we note from (3.2.5) that r̄ = r, i.e. any
circle with center at the origin O is invariant with respect to the map. In its
restriction to an invariant circle, the map has the form
ϕ̄ = ϕ + ω (mod 2π) .
124 Chapter 3. Structurally Stable Periodic Trajectories
Fig. 3.2.9. An example of the behavior of the trajectories of a degenerate map. The entire
x-axis consists of fixed points.
Thus, all points are the fixed point of the N -th iterate of the map T . This
implies that T N is the identity map.
If ω is not commensurable with 2π, then the trajectory of any point ϕ0 is
non-periodic. Moreover, one can show that the set of the points
{ϕj }j=+∞
j=−∞
x̄ = Ax , A ∈ Rn . (3.3.1)
k∆Ak ≤ ε (3.3.6)
xj = A j x0 . (3.3.8)
126 Chapter 3. Structurally Stable Periodic Trajectories
When all eigenvalues of the matrix A lie strictly inside the unit circle, it follows
from (3.3.7) that
kxj k ≤ kAkj kx0 k ≤ (ρ0 + ε)j kx0 k for j ≥ 0, (3.3.9)
i.e. all trajectories converge exponentially to the fixed point at the origin as
j → +∞.
A transition to an arbitrary basis alters the estimate (3.3.9) so that some
constant, generally speaking greater than 1 (see formulas (2.3.17) and (2.3.18)
in Sec. 2.3), appears in the right-hand side.
As in the case of a stable equilibrium state, we may introduce the notions
of leading and non-leading multipliers of a stable fixed point.
Let us arrange the multipliers in order of decreasing absolute value, and let
the first m multipliers be of equal absolute value, i.e.
|ρ1 | = · · · = |ρm | = ρ0 , |ρi | < ρ0 < 1 for i ≥ m + 1.
Denote by E L the m-dimensional eigen-subspace of the matrix A which
corresponds to multipliers (ρ1 , . . . , ρm ), and by E ss the (n − m)-dimensional
eigen-subspace which corresponds to multipliers (ρm+1 , . . . , ρn ). The subspace
E L is called the leading invariant subspace and E ss is called the non-leading,
or strongly stable, invariant subspace.
Each vector x ∈ Rn is uniquely represented in the form
x = u+v,
where u ∈ E L and v ∈ E ss . In coordinate system (u, v) the map (3.3.1) may
be written as follows
ū = AL u ,
v̄ = Ass v ,
where spectr AL = {ρ1 , . . . , ρm } and spectr Ass = {ρm+1 , . . . , ρn }. A trajec-
tory of the map is given by the formula
uj = AjL u0 ,
(3.3.10)
vj = Ajss v0 .
As in (3.3.7) we have
kuj k ≥ (ρ0 − ε)j ku0 k ,
(3.3.11)
kvj k ≤ (|ρm+1 | + ε)j kv0 k ,
3.3. Fixed points of high-dimensional linear maps 127
(1) In the case m = 1, i.e. when ρ1 is real and |ρi | < ρ1 (i = 2, . . . , n), the
leading subspace is a straight line. When ρ1 > 0 all trajectories outside
of E ss converge to O along a certain direction, either from u > 0 or from
u < 0, as shown in Fig. 3.2.5. Such a fixed point is called a stable node
(+).
(3) When m = 2 and ρ1,2 = ρ0 e±iω , ω ∈ / {0, π}, the fixed point is called a
stable focus. The leading subspace of the fixed point is two-dimensional
and all trajectories outside of E ss approach O along spirals tangential to
the plane u.
The case of a completely unstable fixed point where the absolute values of
all of its multipliers ρi are greater than 1, is reduced to the previous case via
its inverse map (as the eigenvalues of the matrix A−1 are equal exactly to ρ−1
i ).
Therefore, an estimate
As in the case of stable fixed points, we may now define leading and non-
leading invariant subspaces and select three principal classes of completely
unstable fixed points: an unstable node (+), an unstable node (−) and an
unstable focus according to the signs of the multipliers.
When some multipliers of the fixed point lie strictly inside the unit cycle
|ρi | < 1 (i = 1, . . . , k) and all others lie outside of it: |ρj | > 1 (j = k +1, . . . , n),
the fixed point is called a saddle fixed point. A linear non-degenerate change
of coordinates transforms the map into the form
ū = A− u ,
(3.3.14)
v̄ = A+ v ,
Fig. 3.4.1. Graphical representation of the homeomorphism η(T1 (x)) = T2 (η(x)) realizing
the topological conjugacy between two maps T1 and T2 defined in subregions U1 and U2 ,
respectively.
130 Chapter 3. Structurally Stable Periodic Trajectories
has some multipliers equal to 1 in absolute value, then one can add a nonlinear
term g(x) such that the map
x̄ = Ax + g(x) (3.4.3)
is not topologically conjugate to its linear part (3.4.2). For example, the one-
dimensional map
x̄ = x + x2
has only one fixed point O (see Fig. 3.4.2), whereas all points on the bisectrix
are fixed points of the associated linearized map
x̄ = x .
x̄ = −x + x3 ,
Fig. 3.4.2. A Lamerey stair. The graph of the function f (x) = x + x2 is tangent to the
bisectrix at the fixed point of the saddle-node type.
3.4. Topological classification of fixed points 131
Fig. 3.4.3. A Lamerey spiral of the map x̄ = −x + x3 . Outside of the origin the derivative
of the map is less than 1 in absolute value: the fixed point is stable.
possessing the stable fixed point O (see Fig. 3.4.3) is not topologically conjugate
to its linear part
x̄ = −x ,
for which all points (apart from O) are periodic trajectories of period 2.
For our next example let us examine a linear map possessing a fixed point
with a pair of complex-conjugate multipliers e±iω
x̄ = x cos ω − y sin ω ,
(3.4.4)
ȳ = x sin ω + y cos ω .
All trajectories lie on invariant circles centered at the point O(0, 0) (see
Sec. 3.2). This map is not conjugate to the map
x̄ = x cos ω − y sin ω − x(x2 + y 2 ) cos ω ,
(3.4.5)
ȳ = x sin ω + y cos ω − y(x2 + y 2 ) cos ω
whose trajectories tend to O along spirals (the analogous example of an equi-
librium state is given in Sec. 2.5).
132 Chapter 3. Structurally Stable Periodic Trajectories
It follows from Theorem 3.1 that when all multipliers of the fixed point
O of the diffeomorphism T are less than 1 in absolute value, then all forward
trajectories tend to O. When all multipliers lie outside of the unit circle,
the backward trajectory of a point from a small neighborhood of the point
O tends to the fixed point. Forward iterations of the map T force all tra-
jectories (excluding the fixed point) to escape from a neighborhood of the
fixed point.
In the saddle case where there are multipliers both inside and outside of
s
the unit circle, the fixed point has (locally) a stable invariant manifold W loc
u
and an unstable invariant manifold Wloc which are the images of the invariant
subspaces E s and E u of the associated linearized system by the homeomorphism
η which establishes the topological conjugacy. Therefore, the forward semi-
s s
trajectory of any point in Wloc lies entirely in Wloc and tends to the saddle
u
point O. On the other hand, given any point in Wloc its backward semi-
u
trajectory lies entirely in Wloc and tends to O. The dimension of the stable
manifold is equal to the number of the multipliers inside the unit circle and the
dimension of the unstable manifold is equal to the number of multipliers outside
s u
of the unit circle. Trajectories of the points which do not lie in Wloc ∪ Wloc
diverge from any neighborhood of the saddle.
It is obvious that if one diffeomorphism near a saddle fixed point is topo-
logically conjugate to another diffeomorphism near another fixed point, then
the dimensions of the stable (unstable) manifolds of both saddle points must
be equal (for generality, in the case of a stable fixed point we can assume that
W u = {∅} and, therefore dim W u = 0; for a completely unstable point assume
that W s = {∅} and dim W s = 0). However, in contrast to the case of struc-
turally stable equilibrium states, the dimensions of the stable and the unstable
manifolds are not the only invariants of the topological conjugacy near the
fixed points.
In order to find new invariants we notice that the Grobman–Hartman the-
orem may be generalized as follows
In a neighborhood of the origin a linear non-singular map which has no
multipliers on the unit circle, is topologically conjugate to any, suffi-
ciently close map.
This implies in particular that any two close, linear maps are topologically
conjugate. It follows that for two arbitrary matrices A0 and A1 the maps
x̄ = A0 x and x̄ = A1 x
3.4. Topological classification of fixed points 133
x̄ = As x , ȳ = Au y , (3.4.6)
We should emphasize that maps of the kind (3.4.6) with different values
of δs cannot be topologically conjugate since the restriction x̄ = As x of the
map (3.4.6) to the stable invariant subspace y = 0 preserves the orientation in
Rk provided that δs = 1, but does not preserve it if δs = −1. This assertion
applies to δu as well. In conclusion we come to the following theorem.
Theorem 3.2. Two structurally stable fixed points are topologically conjugate
if and only if they are of the same topological type.
134 Chapter 3. Structurally Stable Periodic Trajectories
have the fixed points of node type but the first point is orientable whereas the
second is non-orientable.
(Note: a fixed point possessing a pair of complex-conjugate multipliers has
also the topological type of orientable node.)
Observe next that the maps
( (
x̄ = 2x x̄ = 2x
and
ȳ = 2y ȳ = −2y
This implies that the values δs and δu of the fixed point of the Poincaré map
must have the same sign. Nevertheless, restricted to an invariant (stable or
unstable) manifold, the Poincaré map may not continue to preserve orientation
(for example, when both δs and δu are negative). Thus, the study of fixed
points of non-orientable maps makes the sense.
We showed in Sec. 3.1 that the values of the multipliers of a fixed point of the
Poincaré map (and, consequently, its topological type) are independent of the
choice of the cross-section, i.e. we can always correctly define the topological
type (k, δs , n − k, δu ) of a periodic trajectory. In order to complete our
classification we refer to the following simple statement.
From this lemma and from Theorem 3.2 we have the following theorem.
Theorem 3.3. Two structurally stable periodic trajectories are locally topo-
logically equivalent if and only if they have the same topological type.
has a stable fixed point at the origin. This means that the absolute values of
all multipliers ρi (i = 1, . . . , n) of the matrix A are strictly less than 1. It is
not hard to verify that all trajectories beginning from a small neighborhood of
the origin tend exponentially to O. Indeed, if we choose the Jordan basis, we
can verify that the estimate (3.3.9) for the norm of the matrix A is fulfilled;
namely,
kAk ≤ ρ0 + ε ,
where ρ0 = max |ρi | < 1. From the relation
i=1,...,n
Z 1
h(x) = h(0) + x h0 (sx)ds
0
we have
kh(x)k ≤ kxk max kh0 (y)k .
kyk≤kxk
Hence
kh(x)k ≤ εkxk , (3.5.3)
where ε > 0 can be chosen arbitrarily small since x is assumed to be small.
Consequently, for map (3.5.1) we obtain
kx̄k ≤ kAk kxk + kh(x)k ≤ (ρ0 + 2ε)kxk ,
or, equivalently for the j-th iteration of an initial point x0
kxj k ≤ (ρ0 + 2ε)j kx0 k (3.5.4)
where xj → 0 as j → +∞ since ρ0 < 1 and since ε can be made arbitrarily
small.
Let us now reorder the eigenvalues of the matrix A so that
|ρ1 | = · · · = |ρm | = ρ0 , |ρi | < ρ0 for i = m + 1, . . . , n .
The matrix A can be then represented in the form
!
A1 0
A= ,
0 A2
where spectrA1 = {ρ1 , . . . , ρm } and spectrA2 = {ρm+1 , . . . , ρn }. The map
(3.5.1) now takes the form
ū = A1 u + f (u, v) ,
(3.5.5)
v̄ = A2 v + g(u, v) ,
3.5. Properties of nonlinear maps near a stable fixed point 137
u = ϕ(v)
where
ϕ(0) = ϕ0v (0) = 0 . (3.5.7)
where
f˜(0, 0) = 0 . (3.5.12)
138 Chapter 3. Structurally Stable Periodic Trajectories
In the Jordan basis (see Sec. 3.3), we have for the norm of the matrix A1
kA−1
1 k
−1
≥ ρ0 − ε/2 .
κj → 0 as j → +∞ . (3.5.15)
Observe that
J−1 j−1
!
X X
j−J j−(i+1)
Ij ≤ µ κi + µ max κi
i≤J
i=0 i=J
J−1
X
≤ µj−J κi + (1 − µ)−1 max κi . (3.5.16)
i≥J
i=0
tends to O with the exponential rate equal asymptotically |ρm+1 |. Those par-
ticular trajectories which tend to O faster, form a Cr -smooth manifold Wloc sss
or in polar coordinates
r̄ = (ρ + · · · ) r (3.5.23)
ϕ̄ = ϕ + ω + · · · , (3.5.24)
where the ellipsis denotes the terms of a higher order. It follows from (3.5.23)
ss
and (3.5.24) that all trajectories which do not belong to Wloc must spiral
towards to O (tangentially to the leading plane v = 0).
The case where |ρi | > 1, (i = 1, . . . , n) is reduced to that discussed above
by considering the inverse map. In this case a trajectory is estimated by
j
kxj k ≤ min ρi − 2ε kx0 k for j ≤ 0. (3.5.25)
i=1,...,n
Theorem 3.6. (On the leading invariant manifold) A stable fixed point O
has an m-dimensional Cmin(r,rL ) -smooth invariant manifold WlocL
(not unique
in general) which is tangent at the point O to the subspace v = 0; here
ln ρm+1
rL = ≥ 1, (3.5.26)
ln ρ1
where [x] denotes the largest integer strictly less than x, and m is the number
of the leading multipliers.
denote the multipliers inside the unit circle by (λ1 , . . . , λk ), and those outside
by (γ1 , . . . , γn−k ). We will also assume that the multipliers are ordered in the
following manner
ū = A− u + f (u, v) ,
(3.6.1)
v̄ = A+ v + g(u, v) ,
Theorem 3.7. (Hadamard’s theorem) The saddle fixed point O has two
s
invariant manifolds: a stable manifold Wloc : v = ψ ∗ (u) and an unstable man-
u ∗ ∗ ∗
ifold Wloc : u = ϕ (v), where ψ (u) and ϕ (v) satisfy the following Lipschitz
conditions:
kϕk ≤ δ (3.6.4)
kϕ0 k ≤ L (3.6.5)
for some L > 0. We will show below that when δ is sufficiently small then
the intersection T (W) ∩ (D1 ⊗ D2 ) is a surface of the same form ū = ϕ̃(v̄)
where ϕ̃ satisfies conditions (3.6.4)–(3.6.5) with the same constant L. This
allows us to consider a sequence of surfaces {Wj : u = ϕj (v)}j=∞ j=0 which are
the sequential images of the initial surface W under the map T : ϕj = T j ϕ.
We will show further that this sequence converges uniformly to some surface
u = ϕ∗ (v) satisfying a Lipschitz condition. Moreover ϕ∗ does not depend on
the initial function ϕ. By construction, ϕ∗ = ϕ̃∗ , i.e. its graph is invariant with
respect to the map T . Thus, the surface u = ϕ∗ (v) is the desired invariant
u u u
manifold Wloc : T (Wloc ) ∩ (D1 ⊗ D2 ) = Wloc .
Step 1. Let us choose an arbitrary surface W of the form u = ϕ(v) which
satisfies conditions (3.6.4) and (3.6.5) for some L. By substituting u = ϕ(v)
into (3.6.1) we obtain a parametric representation
v̄ = A+ v + g(ϕ(v), v) , (3.6.7)
for the image of the surface W under the map T , where v can take arbitrary
values in D2 .
Let us now show that for any v̄ whose norm does not exceed δ the value ū
is uniquely defined by (3.6.6) and (3.6.7). To do this we rewrite (3.6.7) in the
form
v = (A+ )−1 (v̄ − g(ϕ(v), v)) . (3.6.8)
When δ is sufficiently small the norm k∂(g, f )/∂(u, v)k is also small. It follows
that2
dg(ϕ(v), v)
≤ kgu0 k◦ · kϕ0 k◦ + kgv0 k◦
dv
2 Here k · k◦ = sup k · k.
144 Chapter 3. Structurally Stable Periodic Trajectories
whence
(kv̄k + kgu0 k◦ · kϕ(v)k◦ )
kvk ≤ k(A+ )−1 k .
1 − k(A+ )−1 k kgv0 k◦
Thus, if kv̄k ≤ δ, then kvk ≤ δ because k(A+ )−1 k < 1, kϕ(v)k◦ ≤ δ and
0
kg(u,v) k◦ is small.
So, by expressing v in terms of v̄ from (3.6.8) and by substituting the
resulting expression into (3.6.6), we determine that for each v̄ such that kv̄k ≤
δ, there exists a uniquely defined ū such that the point (ū, v̄) is the image of
some point (u, v) ∈ W. Let us denote this ū by ū = ϕ̃(v̄).
Step 2. Let us show that the surface T W: ū = ϕ̃(v̄) satisfies conditions
(3.6.4) and (3.6.5). In other words we will show that T W lies entirely in the
δ-neighborhood (D1 ⊗ D2 ) of the point O, and that the norm of the derivative
of the function ϕ̃ does not exceed L. It follows from (3.6.4) and (3.6.6) that
It follows that
kϕ̃(v̄)k ≡ kūk ≤ δ
− 0
as kA k < 1 and the norm kf(u,v) k◦
is small; i.e. condition (3.6.4) holds for ϕ̃
indeed.
Furthermore, from (3.6.6) and (3.6.7) we have
dū
= A− ϕ0 + fu0 (ϕ(v), v)ϕ0 + fv0 (ϕ(v), v) ,
dv
dv̄
= A+ + gu0 (ϕ(v), v)ϕ0 + gv0 (ϕ(v), v)
dv
whence
dū
ϕ̃0 (v̄) ≡ = (A− ϕ0 + fu0 (ϕ(v), v)ϕ0 + fv0 (ϕ(v), v))·
dv̄
· [A+ + gu0 (ϕ(v), v)ϕ0 + gv0 (ϕ(v), v)]−1 .
3.6. Saddle fixed points. Invariant manifolds 145
Finally
ϕ̃0 (v̄) = A− ϕ0 (v)(A+ )−1 + · · · ,
where the ellipsis denotes the terms of order k(f, g)0(u,v) k◦ which tend to zero
as δ → 0. It is easy to see that since kA− k < 1 and k(A+ )−1 k < 1, then
kϕ̃0 k◦ ≤ L provided kϕ0 k◦ ≤ L and δ is sufficiently small.
Step 3. We have shown that the map T takes any surface W satisfying
conditions (3.6.4) and (3.6.5), onto a surface satisfying the same conditions.
Hence, all iterations of the surface W are defined under the action of the map T .
Let us now show that the sequence of these iterations Wj : u = ϕj (v) converges
uniformly to some surface W ∗ : u = ϕ∗ (v). Since Wj+1 = T (Wj ) ∩ (D1 ⊗ D2 ),
it follows from continuity that W ∗ = T (W ∗ ) ∩ (D1 ⊗ D2 ), i.e. this surface is
invariant with respect to the map T .
In order to prove this, we will show that there exists K < 1 such that
sup kϕj+2 (v̄) − ϕj+1 (v̄)k ≤ K sup kϕj+1 (v) − ϕj (v)k . (3.6.9)
v̄∈D2 v∈D2
Let us choose any v̄ ∈ D2 and consider a pair of points M̄1 (ϕj+1 (v̄), v̄) and
M̄2 (ϕj+2 (v̄), v̄). By construction, each point M̄i has a pre-image Mi on the sur-
face u = ϕj+i−1 (v). Assume M1 (u1 = ϕj (v1 ), v1 ) and M2 (u2 = ϕj+1 (v2 ), v2 ).
Because M̄1 = T M1 and M̄2 = T M2 , we have from (3.6.1)
and
kv2 − v1 k ≤ k(A+ )−1 k(kgu0 k◦ ku2 − u1 k + kgv0 k◦ kv2 − v1 k)
whence
k(A+ )−1 k kgu0 k◦ ku2 − u1 k
kv2 − v1 k ≤ . (3.6.11)
1 − k(A+ )−1 k kgv0 k◦
For ku2 − u1 k = kϕj+1 (v2 ) − ϕj (v1 )k we have the following estimate
kϕj+1 (v2 ) − ϕj (v1 )k ≤ kϕj+1 (v2 ) − ϕj+1 (v1 )k + kϕj+1 (v1 ) − ϕj (v1 )k .
146 Chapter 3. Structurally Stable Periodic Trajectories
where the ellipsis denotes terms of the order k(f, g)0(u,v) k◦ . Since kA− k < 1,
we have
kū2 − ū1 k = kϕj+2 (v̄) − ϕj+1 (v̄)k ≤ K sup kϕj+1 (v) − ϕj (v)k
for some K < 1 which does not depend on v̄ (provided that kv̄k ≤ δ). If we
take the supremum in the left-hand side of this inequality with respect to all
v̄, then we obtain the desired inequality (3.6.9).
From (3.6.9) we obtain
kϕ0j (v)k ≤ L .
Thus, since kA− k < 1 and since the norm kfu0 k is small, the iterations of any
s
point on Wloc converge exponentially to the point O under the action of the
map T .
By the symmetry, an analogous result may be obtained for the map (3.6.1)
u
in its restriction to Wloc :
u
namely, for any point v̄ on Wloc there exists a uniquely defined image v = T −1 v̄
which iterations under the map T −1 tend uniformly and exponentially to the
point O.
s u
Theorem 3.8. The invariant manifolds Wloc and Wloc belong to the Cr -class
of smoothness and at the point O they are tangent respectively to the stable
eigen-subspace v = 0 and the unstable eigen-subspace u = 0, i.e.
Proof. As above we will only prove that part of this theorem which con-
u s
cerns Wloc . Smoothness of the manifold Wloc follows from the symmetry of
u
the problem. The invariance of the manifold Wloc implies that if some point
u ∗
M (u, v) belongs to Wloc , i.e. u = ϕ (v) and if its image M̄ (ū, v̄) remains in a
u
δ-neighborhood of the point O, then the point M̄ also belongs to Wloc , i.e. its
∗
coordinates satisfy ū = ϕ (v̄).
From (3.6.1) we have
for the convergence of successive approximations, the functions ηj are no longer required to
be smooth and to have the bounded derivatives. The reason is that the map T ∗ is a triangular
map, and the second equation in (3.6.18) does not depend on the variable η.
150 Chapter 3. Structurally Stable Periodic Trajectories
Let us determine the relation between the values of z(v) and z(v̄), where v̄
is given by (3.6.14). From (3.6.14) we have
∆v̄ = A+ + gv0 (ϕ∗ (v), v) + gu0 (ϕ∗ (v), v)η ∗ (v) ∆v
+gu0 (ϕ∗ (v), v)(∆ϕ − η ∗ (v)∆v) + o(∆v) , (3.6.25)
whence
From (3.6.24) and (3.6.25) we have the following estimate for k∆vk:
It was already noted that for any point v̄ which does not exceed δ in the
norm, there exists a pre-image v such that kvk ≤ δ. Therefore, for any point v0
the infinite sequence {vj } is defined such that vj = v̄j+1 . By virtue of (3.6.28)
j
z(v0 ) ≤ kA− k k(A+ )−1 k + · · · z(vj ) ,
and since z(vj ) is bounded and kA− k < 1, k(A+ )−1 k < 1, it follows that
z(v0 ) = 0. As v0 was chosen arbitrarily, it follows that z(v) ≡ 0, i.e. the
smoothness of the function ϕ∗ is established.
It should be noted that the statement concerning the existence of the invari-
ant manifold {η = η ∗ (v)} of the map (3.6.18) is, generally speaking, satisfied
only for sufficiently small v : kvk ≤ δ1 , δ1 > 0. We have disregarded the fact
that the value δ1 may be less than δ, which is the diameter of the neighbor-
hood of the origin in which the function ϕ∗ is defined. Nevertheless, one can
show that the function ϕ∗ is a smooth function for all v in a δ-neighborhood
of the origin. To do this we first note that since the backward iterations of any
u
point v̄ in Wloc in a δ-neighborhood of the origin converge uniformly to the
u
point O, the image of the δ1 -neighborhood of the origin on Wloc will cover the
δ-neighborhood after a number of forward iterations of the map T . Thus it is
implied that because the map T is smooth and because in the δ1 -neighborhood
u u
of the point O the manifold Wloc is also smooth, Wloc is smooth inside the orig-
inal δ-neighborhood.
Step 3. We have established that the map T has a smooth invariant mani-
fold of the form u = ϕ∗ (v). Moreover, the graph of the derivative η ∗ = dϕ∗ /dv
is itself an invariant manifold of the map T ∗ given by formulae (3.6.17) and
(3.6.18). If the smoothness of the right-hand side of the map T is greater than
one, then the right-hand side of the map T ∗ belongs to the C1 -class (because
it is expressed in terms of ϕ∗ and g). As the fixed point (v = 0, η = 0) of
the map T ∗ is a saddle point, all the arguments used for the map T can be
repeated, leading to the conclusion that the invariant manifold η = η ∗ (v) of
152 Chapter 3. Structurally Stable Periodic Trajectories
the map T ∗ is smooth and consequently that the function ϕ∗ belongs to the
C2 -class.4
Thus, when the smoothness of the right-hand side of the map T is greater
than two, the right-hand side of the map T ∗ is then already C2 -smooth. There-
fore, by virtue of the previous arguments the function η ∗ is of C2 -class, and
the function ϕ∗ is of C3 -class, respectively, and so on. By induction we arrive
at the existence of a Cr -smooth invariant manifold Wloc u
.
End of the proof.
s u
As in the case of equilibrium states, the invariant manifolds Wloc and Wloc
can be locally straightened by a change of variables:
ξ = u − ϕ∗ (v) ,
η = v − ψ ∗ (u) .
s u
Wloc : η = 0, and Wloc : ξ = 0.
In a small neighborhood of the saddle the functions h1,2 are small in norm
and as long as a trajectory remains in a neighborhood of the saddle, the
inequalities
¯ ≤ (|λ1 | + ε)kξk
kξk
and
kη̄k ≥ (|γ1 | − ε)kηk
4 With the only difference being that the linear part of the map T ∗ is not block-diagonal,
∗
whence dη
dv
(0) 6= 0, in general.
3.6. Saddle fixed points. Invariant manifolds 153
(see the proof of the analogous formulae (3.5.4) and (3.5.25) in the previous
s u
section). Thus, a trajectory that lies in neither Wloc nor in Wloc , leaves a
neighborhood of the saddle as j → ±∞. Moreover, the number of iterations
needed for a forward trajectory to escape from the neighborhood of the saddle
is of the order ln kη0 k, and that for a backward trajectory is of the order ln kξ0 k.
s
The map (3.6.1) on the stable manifold Wloc : v = ψ ∗ (u) is given by
Here, the point O is a completely unstable fixed point and in the generic case,
it is either a node or a focus.
We can now identify nine main types of saddle fixed points depending on
the behavior of trajectories in the leading coordinates:
s u
(1) a saddle (+, +): a node (+) on both Wloc and Wloc ;
s u
(2) a saddle (−, −): a node (−) on both Wloc and Wloc ;
s u
(3) a saddle (+, −): a node (+) on Wloc and a node (−) on Wloc ;
s u
(4) a saddle (−, +): a node (−) on Wloc and a node (+) on Wloc ;
s u
(5) a saddle-focus (2, 1+): a focus on Wloc and a node (+) on Wloc ;
s u
(6) a saddle-focus (2, 1−): a focus on Wloc and a node (−) on Wloc ;
s u
(7) a saddle-focus (1+, 2): a node (+) on Wloc and a focus on Wloc ;
s u
(8) a saddle-focus (1−, 2): a node (−) on Wloc and a focus on Wloc ;
s u
(9) a saddle-focus (2,2): a focus on both Wloc and Wloc .
154 Chapter 3. Structurally Stable Periodic Trajectories
Theorems 3.4 and 3.5 are valid for systems (3.6.33) and (3.6.34). This im-
s u
plies that in Wloc and Wloc there exists a non-leading stable invariant subman-
ss sL
ifold Wloc , a leading stable invariant submanifold Wloc , a non-leading unstable
uu uL
invariant submanifold Wloc , and a leading unstable invariant submanifold Wloc .
We further select an extra three smooth invariant manifolds of a saddle fixed
point. Introduce the notations:
" #
ln λ̂
rsL = , (3.6.35)
ln |λ1 |
ln γ̂
ruL = , (3.6.36)
ln |γ1 |
where λ̂ and γ̂ respectively are the absolute values of non-leading stable and
unstable multipliers nearest to the unit circle; [x] denotes the largest integer
strictly less than x.
Theorem 3.9. In a neighborhood of a structurally stable fixed point of the
saddle type of a Cr -smooth map there exists the following invariant manifolds:
1. a Cmin (r,ruL ) -smooth extended stable manifold WlocsE s
which contains Wloc ,
and which is tangent at the point O to the extended stable eigen-subspace
uu
of the linearized system and transverse to Wloc ;
2. a Cmin (r,rsL ) -smooth extended unstable manifold Wloc
uE
which contains
u
Wloc and which is tangent at O to the extended unstable eigen-subspace
ss
of the linearized system and transverse to Wloc ;
3. a Cmin (r,rsL ,ruL ) -smooth leading saddle manifold Wloc
L uE
= Wloc sE
∩ Wloc .
See the proof in the Chap. 5. We note that, generally speaking, the manifold
sE
Wloc is not unique but any two such manifolds are tangent to each other
s uE
everywhere on Wloc . Analogously, any two manifolds Wloc are tangent to each
u
other on Wloc .
where u ∈ Rm1 and v ∈ Rm2 . Let O(0, 0) be a saddle fixed point of the map T ,
i.e. spectrA− = {λ1 , . . . , λm1 } lies strictly inside and spectrA+ = {γ1 , . . . , γm2 }
lies outside of the unit circle. Assume that the functions f and g vanish at the
origin along with their first derivatives.
Just like in the case of a saddle equilibrium state which we have examined
in Sec. 2.8, exponential instability near a saddle fixed point is a typical feature
of the trajectories of the map (3.7.1). Therefore, in this case, instead of the
initial-value problem it is quite reasonable to solve the boundary-value problem
which can be formalized in the following way:
For any u0 and v 1 , and for an arbitrary k > 0 find a trajectory
u0 ≡ u 0 , vk ≡ v 1 , (3.7.2)
where we assume that ku0 k ≤ ε and kv 1 k ≤ ε for some sufficiently small ε > 0.
A trajectory {(uj , vj )}kj=0 of the map (3.7.1) is given by
uj+1 = A− uj + f (uj , vj ) ,
(3.7.3)
vj+1 = A+ vj + g(uj , vj ) .
Theorem 3.10. For sufficiently small ε > 0 and u0 , v 1 such that ku0 k ≤ ε
and kv 1 k ≤ ε, a solution of the boundary-value problem (3.7.2) for the map
(3.7.1) exists for any positive integer k. The solution is unique and depends
continuously on (u0 , v 1 ).
and for vj :
k−1
X (3.7.6)
(n+1)
vj = (A+ )j−k v 1 − (A+ )j−s−1 g u(n) (n)
s , vs ,
s=j
(j = 0, 1, . . . , k) .
3.7. The boundary-value problem near a saddle fixed point 157
Let us now show that the resulting sequence converges uniformly to some
limit vector i=k
z0∗ = (u∗i , vi∗ ) i=0 .
We first prove that
(n) (n)
kuj k ≤ 2ε , kvj k ≤ 2ε (3.7.7)
where 0 < λ < 1 and γ > 1 are such numbers that spectrA− lies strictly inside
the circle of diameter λ, and spectrA+ lies outside the circle of diameter γ.
We will prove inequality (3.7.7) for all n by induction. Since both functions
f and g vanish at the point O along with their first derivatives, the inequalities 5
∂(f, g)
≤ δ, kf, gk ≤ δku, vk (3.7.9)
∂(u, v)
are satisfied, where δ may be made arbitrarily small by decreasing the size of
the neighborhood of the point O. Choose ε sufficiently small so that for any u
and v in the 2ε-neighborhood of the saddle the inequality
1 1
2δ max , ≤1 (3.7.10)
1 − λ 1 − γ −1
holds. From (3.7.6), (3.7.8) and (3.7.9) we obtain
j−1
X
(n+1)
kuj k ≤ λj ku0 k + δ λj−s−1 ku(n) (n)
s , vs k
s=0
k−1
X
(n+1)
kvj k ≤ γ j−k kv 1 k + δ γ j−s−1 ku(n) (n)
s , vs k
s=j
1
max ku(n) − u(n−1)
≤ , vs(n) − vs(n−1) k . (3.7.11)
2 0≤s≤k s s
(n) (n)
Since the variables uj , vj lie in the 2ε-neighborhood of the saddle for
(n) (n)
all n, it follows that the estimates (3.7.9) are valid for values f uj , vj and
(n) (n)
g uj , vj . Now, from (3.7.6) and (3.7.10) we obtain
(n+1) (n)
kuj − uj k
j−1
X
≤ λj−s−1 kf u(n) (n)
s , vs − f u(n−1)
s , vs(n−1) k
s=0
δ
≤ max ku(n) − u(n−1) , vs(n) − vs(n−1) k
1 − λ 0≤s≤j s s
1
≤ max ku(n) − u(n−1) , vs(n) − vs(n−1) k .
2 0≤s≤j s s
(n+1) (n)
An analogous estimate applies to kvj − vj k.
(n+1) (n)
It follows from (3.7.11) that the norms of the differences kuj − uj k
(n+1) (n)
and kvj − vj k decay in a geometric progression. Therefore, the series
∞
X
(n+1) (n) (n+1) (n)
uj − u j , vj − vj (3.7.12)
n=1
1
ku∗∗ ∗ ∗∗ ∗
j − u j , vj − v j k ≤ max ku∗∗ − u∗s , vs∗∗ − vs∗ k
2 0≤s≤k s
(j = 0, . . . , k) ,
U 0 = I m1 , Vk = 0 , (3.7.14)
V k = I m2 , and U0 = 0 . (3.7.16)
Lemma 3.2. (λ-lemma) For all sufficiently large k the surface Hk is repre-
sented in the form u = hk (v), where the functions hk tend uniformly to zero,
along with all their derivatives, as k → +∞ (see Fig. 3.7.1).6
s
Proof. Transversality of the surface H0 with respect to the surface Wloc :
0
v = 0 at the point M (h0 (0), 0) implies that kh0 (0)k is bounded. Therefore,
the norm kh00 (v)k is bounded for all sufficiently small v. Let us consider the
surface Hk and choose an arbitrary point (uk , vk ) on it. By construction, there
always exists a point (u0 , v0 ) on H0 such that T k (u0 , v0 ) = (uk , vk ).
6 In u in the
other words, the sequence Hk converges to Wloc Cr -topology.
3.7. The boundary-value problem near a saddle fixed point 161
Fig. 3.7.1. A geometrical interpretation of the λ-lemma. With each successive iteration the
graph of the surface u = hk (v) becomes flatter and flatter while approaching the unstable
manifold W u along the stable manifold W s .
We have already noted that ηk along with all their derivatives up to order
r converges to zero as k → +∞, whereas kh00 k remains bounded for small v0 .
Hence, by virtue of the implicit function theorem, for sufficiently large k and
any vk whose norm does not exceed ε, the second equation in (3.7.19) can be
uniquely resolved with respect to v0 : v0 = ϕk (vk ), where the function ϕk tends
uniformly to zero along with all the derivatives up to order r as k → +∞.
162 Chapter 3. Structurally Stable Periodic Trajectories
The equation of the surface Hk can now be recast in the explicit form uk =
ξk (h0 (ϕk (vk )), vk ), what gives the lemma because the norms of the functions
ξk and ϕk converge uniformly to 0 as k → +∞.
Thus, the proof of the λ-lemma is reduced to verifying that the norms kξk k
and kηk k tend to 0 along with the norms of their derivatives. Let us prove
this.
Lemma 3.3. In the coordinate system where the stable and the unstable man-
ifold are straightened, the norms of the functions kξk k and kηk k tend uniformly
to zero as k → ∞.
Proof. Consider the system (3.7.5) which yields the solution of the
boundary-value problem for the map T . We have uk ≡ ξk (u0 , vk ) and v0 ≡
ηk (u0 , vk ). We will show that solutions of system (3.7.5) satisfy the inequalities
for some K and for some λ̄ < 1, γ̄ > 1. From the proof of Theorem 3.10
one can see that the solution
of system
(3.7.5) is found as the limit of the
(n) (n)
successive approximations uj , vj , which are calculated by formula (3.7.6).
It is therefore sufficient to check that inequalities (3.7.20) hold for all steps of
successive approximations with the same values of k, λ̄, γ̄.
For the first approximation
(1) (1)
uj = (A− )j u0 , vj = (A+ )−(k−j) v 1
the validity of (3.7.20) follows from (3.7.8) provided that we choose K > ε and
λ̄ > λ, γ̄ < γ. Now let us show that if (3.7.20) holds for the n-th approximation,
then it holds for the (n + 1)-th approximation as well. Observe first that it
follows from (3.7.17) and (3.7.9) that the functions f and g satisfy the following
estimates
!
kf (u, v)k ≤ kf (0, v)k + sup kfu0 k kuk ≡ δkuk (3.7.21)
ku,vk≤ε
and !
kg(u, v)k ≤ kg(u, 0)k + sup kgv0 k kvk ≡ δkvk . (3.7.22)
ku,vk≤ε
3.7. The boundary-value problem near a saddle fixed point 163
(n) (n) (n+1) (n+1)
Hence, if us , vs satisfies (3.7.20), then for uj , vj we have
j−1
(n+1)
X s j δK
kuj k ≤ λj ε + δ λj−s−1 Kλ ≤ λ ε+ ,
s=0
λ−λ
k−1
(n+1)
X δK
kvj k ≤ γ j−k ε + δ γ j−s−1 Kγ s−k ≤ γ j−k ε + .
s=j
γ−γ
(n+1) (n+1)
The inequalities (3.7.20) hold for uj , vj provided that
1 1
K> ε + δK max , .
λ−λ γ−γ
Since δ may be made arbitrarily small for sufficiently small ε, such a constant K
exists. Thus, one can select K, λ and γ such that the inequalities (3.7.20) hold
for all approximations, and, consequently, for the solution of the boundary-
value problem itself.
For the functions ξk and ηk we found
k −k
kξk k ≤ Kλ , kηk k ≤ Kγ ,
i.e. the norms of these functions tend uniformly and exponentially to zero as
k → +∞. End of the proof.
Lemma 3.4. The norms of the derivatives ∂(ξk , ηk )/∂(u0 , vk ) tend uniformly
to zero as k → ∞.
Proof. Let us consider the derivatives ∂ξk /∂u0 and ∂ηk /∂u0 . They are
found from the solutions of the boundary-value problem (3.7.13), (3.7.14):
∂ξk /∂u0 ≡ Uk , ∂ηk /∂u0 ≡ Vk . Before we show that both Uk (u0 , vk ) and
164 Chapter 3. Structurally Stable Periodic Trajectories
V0 (u0 , vk ) tend to zero as k → +∞, we will prove that all Uj and Vj are
bounded by a constant which depends neither on k nor on j.
The values Uj and Vj are found from the system (3.7.15) as a limit of the
successive approximations
j−1
X
(n+1)
Uj = (A− )j + (A− )j−s−1 fu0 u∗s , vs∗ Us(n) + fv0 u∗s , vs∗ Vs(n) ,
s=0
(3.7.23)
k−1
X
(n+1)
Vj =− (A+ )j−s−1 gu0 u∗s , vs∗ Us(n) + gv0 u∗s , vs∗ Vs(n) ,
s=j
where u∗s
and vs∗
are the solutions of the boundary-value problem (3.7.2) and
(3.7.3). We will prove that Uj and Vj are uniformly bounded if we show that all
(n) (n)
of the successive approximations Uj , Vj are bounded by a constant which
is independent of k, j and n. In order to verify this let us suppose that for
all j
(n) (n)
kUj , Vj k ≤ 2 . (3.7.24)
It follows from (3.7.23), (3.7.9) and (3.7.10) that
j−1
X
(n+1)
kUj k ≤ λj + λj−s−1 kfu0 (us , vs )k kUs(n) k + kfv0 (us , vs )k kVs(n) k
s=0
j−1
X
≤ 1 + 2δ λj−s−1 ≤ 1 + 2δ/(1 − λ) ≤ 2 ,
s=0
k−1
X
(n+1)
kVj k≤ γ j−s−1 kgu0 (us , vs )kUs(n) k + kgv0 (us , vs )k kVs(n) k
s=j
k−1
X
≤ 2δ γ j−s−1 ≤ 2δ/(γ − 1) ≤ 1 .
s=j
j−1
X
kUj k ≤ λj + λj−s−1 (δkUs k + ρs ) , (3.7.26)
s=0
j−1
X
j
Zj = λ + λj−s−1 (δZs + ρs ) . (3.7.27)
s=0
whence
Zj+1 = (λ + δ)Zj + ρj . (3.7.29)
Since δ may be chosen sufficiently small, we have λ + δ < 1. Now, the conver-
gence of Zj to zero is proven in the same way as it was done for the sequence
∂ξk
(3.5.14) (taking into account that ρj → 0). Thus, Uk = ∂u 0
tends to zero as
k → +∞. The remaining derivatives ∂ξk /∂vk , ∂ηk /∂u0 , and ∂ηk /∂vk may be
shown to tend to zero as k → +∞ in a similar fashion.
Lemma 3.5. The norms of the first r derivatives of the functions ξk and ηk
tend uniformly to zero as k → ∞.
i ∂ i uj i ∂ i vj
Uj(p,q) ≡ , Vj(p,q) ≡ ,
∂up0 ∂vkq ∂up0 ∂vkq
where p + q = i ≤ r. The values Uji and Vji may be found by the succes-
sive approximations method as solutions of the system obtained from (3.7.5)
166 Chapter 3. Structurally Stable Periodic Trajectories
+ Qi us , vs , . . . , Usi−1 , Vsi−1 ,
where Pi and Qi are certain polynomials of the variables Us1 , Vs1 , . . . , Usi−1 ,
Vsi−1 and of the derivatives of the functions f and g computed at u = us and
v = vs .
For example, for the derivatives
∂2u ∂ 2 vj
2 2 j
Uj(2,0) , Vj(2,0) ≡ ,
∂u20 ∂u20
we have
j−1
X
2
Uj(2,0) = (A− )j−s−1 fu0 (us , vs )Us(2,0) + fv0 (us , vs )Vs(2,0)
s=0
00
+ fuu (us , vs )(Us(1,0) )2 + 2fuv
00
(us , vs )Us(1,0) Vs(1,0)
00
+ fvv (us , vs )(Vs(1,0) )2
and
k−1
X
2
Vj(2,0) =− (A+ )j−s−1 gu0 (us , vs )Us(2,0) + gv0 (us , vs )Vs(2,0)
s=j
00
+ guu (us , vs )(Us(1,0) )2 + 2guv
00
(us , vs )Us(1,0) Vs(1,0)
00
+ gvv (us , vs )(Vs(1,0) )2 .
3.8. Behavior of linear maps near saddle fixed points 167
In the manner previously employed for the first derivatives, one can show
that the derivatives Uji and Vji of any higher order i are bounded by some
constant which depends neither on j nor on k (but it may depend on the order
i of the derivative).
In order to verify that the norm kUki k → 0 as k → +∞, we show that the
norms kUji k are bounded by a sequence which is independent of k and which
tends to zero as j → ∞. We have already proven this statement for i = 1. We
show by induction that it is valid for all i.
Let us assume that kUji k → 0 as j → ∞ for all i less than some i0 . Consider
Eq. (3.7.30) for Uji0 . Those terms in Pi0 which contain at least one of the values
Usi (i < i0 ) tend to zero as s → +∞ by virtue of our inductive hypothesis. The
remaining terms are products of certain values Vsi with certain derivatives of
f (us , vs ) with respect to the variable vs . Since all Vsi are bounded uniformly
and the derivatives of f (us , vs ) with respect to vs tends uniformly to zero
as us → 0 (because f (0, v) ≡ 0), it follows that all these terms, as well as
f (us , vs )Vsi , tend to zero as s → +∞.
Thus, in complete analogy with the discussion concerning the first deriva-
tive, we obtain the estimate
j−1
X
kUji k ≤ λj + λj−s−1 δkUsi k + ρis ,
s=0
k−1
X
kVji k ≤ γ j−s−1 δkVsi k + σk−s
i
,
s=j
i
where σk−s → 0 as (k − s) → +∞. Repeating the same arguments employed
i i
for Uj we can show that kVk−j k → 0 as (k − j) → +∞. Assuming j = k, we
found that all derivatives of ηk tend to zero as k → +∞.
End of the proof.
168 Chapter 3. Structurally Stable Periodic Trajectories
In this section we will study some geometrical properties of the linear saddle
maps. For suitable choice of coordinates a linear map T with a structurally
stable fixed point O of the saddle type can be written in the form
x̄ = AsL x , ū = Ass u ,
(3.8.1)
ȳ = AuL y v̄ = Auu v ,
where the absolute values of the eigenvalues of the matrix AsL are equal to λ,
0 < λ < 1, while those of the matrix Ass are less than λ. The eigenvalues of
the matrix AuL are equal in absolute value to γ, γ > 1 and those of matrix
Auu are greater in absolute value than γ. Then, the equation of the stable
invariant manifold W s is (y = 0, v = 0) and the equation of the non-leading
(strongly) stable manifold W ss is (x = 0, y = 0, v = 0). The equation of the
unstable manifold W u is (x = 0, u = 0), and the equation of thxe non-leading
(strongly) unstable manifold W uu is (x = 0, u = 0, y = 0).
Let us choose two points on the stable and the unstable manifolds:
M + (x+ , u+ , 0, 0) ∈ W s /O, and M − (0, 0, y − , v − ) ∈ W u /O and surround them
by some small rectangular neighborhoods
Π+ = kx − x+ k ≤ ε0 , ku − u+ k ≤ ε0 , kyk ≤ ε0 , kvk ≤ ε0
Π− = kxk ≤ ε1 , kuk ≤ ε1 , ky − y − k ≤ ε1 , kv − v − k ≤ ε1
such that T (Π+ ) ∩ Π+ = ∅ and T (Π− ) ∩ Π− = ∅. We assume also that
the leading eigenvalues of the saddle fixed point O are simple, i.e. there is
only one leading eigenvalue if it is real. Otherwise, there is a pair of leading
eigenvalues if they are complex-conjugate. This implies that in the first case
the vector x (or y) is one-dimensional, and AsL (or AuL ) is a scalar. In the
case where the leading eigenvalues are complex-conjugate, the vector x or y is
two-dimensional, and the matrix AsL or AuL has the form
cos ϕ − sin ϕ cos ψ − sin ψ
AsL = λ and AuL = γ ,
sin ϕ cos ϕ sin ψ cos ψ
where 0 < λ < 1, γ > 1, (ϕ, ψ) ∈
/ {0, π}.
We consider the following question: are there any points in Π+ whose
trajectories reach Π− ? What is the set of such points in Π− and what is the
set of their images in Π+ ?
3.8. Behavior of linear maps near saddle fixed points 169
We first consider the case where a saddle fixed point possesses leading
eigenvalues only. It was established in Sec. 3.6 that there are nine main types
of such saddle maps: four two-dimensional saddle points (all eigenvalues are
real); four three-dimensional: two saddle-foci (2,1) and two saddle-foci (1,2)
and one four-dimensional case: a saddle-focus (2,2).
Let us begin with two-dimensional maps. There may be four different (in
the sense of topological conjugacy) situations depending on the signs of the
eigenvalues of the saddle. The map may take one of the following forms:
(1) x̄ = λx , ȳ = γy ;
(2) x̄ = −λx , ȳ = γy ;
(3) x̄ = λx , ȳ = −γy ;
(4) x̄ = −λx , ȳ = −γy .
σk1 = (x, y): |x − λk x+ | ≤ λk ε0 , |y − y − | ≤ ε1 .
7 In the nonlinear case, the existence of such k̄ follows from the λ-lemma.
170 Chapter 3. Structurally Stable Periodic Trajectories
Fig. 3.8.1. The map T near a saddle fixed point. The initial rectangle Π + is expanded along
the unstable direction y and compressed along the stable direction x. The range of the map
T 0 : Π+ → Π− is composed of the strips σk1 lying in the intersection between the images
T k Π+ and the rectangle Π− .
Fig. 3.8.2. The inverse map T −1 near the saddle. The rectangle Π− is expanded along
the stable direction x and compressed along the unstable direction y under the action of
the inverse map T −1 . The strips σk0 form the domain of definition of the map
0 +
T :Π →Π . −
For case (4) the point O is a stable node (−) on W s and an unsta-
ble node (−) on W u . Therefore, the strips σk0 converge to W s ∩ Π+ from
both sides. The strips σk1 converge to W u ∩ Π− from both sides as well, see
Fig. 3.8.3(d).
Let us now consider the cases where the leading eigenvalues comprise a
complex-conjugate pair.
In the three-dimensional case where the point O is a saddle-focus (2,1), the
linear map can be written in the form
x̄1 = λ(cos ϕ · x1 − sin ϕ · x2 ) ,
x̄2 = λ(sin ϕ · x1 + cos ϕ · x2 ) , (3.8.2)
ȳ = γy ,
where λ1,2 = λe±iϕ and γ are the eigenvalues of the saddle O, ϕ ∈ / {0, π},
0 < λ < 1, |γ| > 1. To be specific, let us consider the case of positive γ. The
172 Chapter 3. Structurally Stable Periodic Trajectories
(a)
(b)
Fig. 3.8.3. The Poincaré map near saddle fixed points of different types. See captions to
Figs. 3.8.1 and 3.8.2. (a) Near a saddle (+, +), (b) near a saddle (−, +). The even and odd
iterations of Π+ lie on the opposite sides from the unstable manifold y(x), (c) near a saddle
(−, +), (d) near a saddle (−, −).
3.8. Behavior of linear maps near saddle fixed points 173
(c)
(d)
W u ∩ Π− = {x1 = x2 = 0, |y − y − | ≤ ε1 }
of the W u -axis, see Fig. 3.8.4. The strip σk1 ⊂ R− has a diameter of order ε0 λk
along the coordinates (x1 , x2 ), and σk1 is separated from σk+1
1
by an angle of
order ϕ in the angular coordinate θ.
In the case where the fixed point is a saddle-focus (1,2) the map T can be
written in the form
x̄ = λx ,
ȳ1 = γ(cos ψ · y1 − sin ψ · y2 ) ,
ȳ2 = γ(sin ψ · y1 + cos ψ · y2 ) ,
3.8. Behavior of linear maps near saddle fixed points 175
Fig. 3.8.4. The geometry of the Poincaré map near a saddle-focus (2,1+).
Fig. 3.8.5. The Poincaré map near a saddle-focus (1+,2). This is the inverse to the map in
Fig. 3.8.4.
of the W s -axis. The strip σk0 has a diameter of order ε1 · γ −k along the coor-
dinates (y1 , y2 ). Moreover, σk0 and σk+1
0
are separated by an angle of order ψ
in the angular coordinate θ.
Let us consider next the case where the fixed point O is a saddle-focus
(2,2). The corresponding linear map T can be written as
where ϕ, ψ ∈ / {0, π}, 0 < λ < 1 < γ. We choose two arbitrary points
M + (x+
1 , x +
2 , 0, 0) ∈ W s \O and M − (0, 0, y1− , y2− ) ∈ W u \O. Without affecting
formulae (3.8.4) we can always ensure that x+ −
1 = 0 and y1 = 0 by the or-
thogonal rotation of the coordinate frames on the planes (x1 , x2 ) and (y1 , y2 ).
Introducing the polar coordinates (r, θ) in the plane (x1 , x2 ) and (ρ, α) in the
plane (y1 , y2 ), map (3.8.4) recasts in the following simple form:
r̄ = λr , θ̄ = θ + ϕ , ρ̄ = γρ , ᾱ = α + ψ .
r̄ = λk r , θ̄ = θ + kϕ , ρ̄ = γ k ρ , ᾱ = α + kψ . (3.8.5)
Since 0 < λ < 1 < γ, it follows from (3.8.5) for sufficiently large k that
γ k ε0 > y2− + ε1 and λk (x+ k +
1 + ε0 ) < ε1 , and hence T (Π ) ∩ Π
−
6= ∅. The
1 k + −
four-dimensional strips σk ≡ T (Π ) ∩ Π converge to the two-dimensional
square
W u ∩ Π− = 0, 0, |y1 | ≤ ε1 , |y2 − y2− | ≤ ε1
as k → +∞. In the plane W s : (x1 , x2 , 0, 0) the points Mk+ ≡ T k (M + ) =
(λk x+
1 , kϕ) lie on the logarithmic spiral r̄ = x1 · λ
+ θ̄/ϕ
. Thus, the range σ1
of the map T : Π → Π is a union of a countable number of the strips σk1
0 + −
Fig. 3.8.6. The Poincaré map near a saddle-focus in R4 . The original three-dimensional
parallepiped Π+ is transformed into a “roulette” within the parallepiped Π− by the map T .
The image of the parallepiped Π+ under the inverse map T −1 has the same shape.
r = λ−k r̄ , θ = θ̄ − kϕ , ρ = γ −k ρ̄ , α = ᾱ − kψ . (3.8.6)
|x1 − x+
1 | ≤ ε0 , |x2 | ≤ ε0 ,
x̄ = λx , ū = λ2 u , ȳ = γy ,
where λ and γ are assumed for more definiteness to be positive, and where
0 < |λ2 | < λ. Since M + 6∈ W ss , it follows that x+ 6= 0 and therefore we can
let x+ > 0. The map T −k : Π− → Π+ is defined by
x = λ−k x̄ , u = λ−k
2 ū , y = γ −k ȳ ,
where (x, u, y) ∈ Π+ and (x̄, ū, ȳ) ∈ Π− . One observes that for sufficiently
large k such that λ−k ε1 > x+ + ε0 and |λ2 |−k ε1 > |u+ | + ε0 , the strips σk0 ≡
T −k (Π− ) ∩ Π+ are given by
σk0 = x, u, y : |x − x+ | ≤ ε0 , |u − u+ | ≤ ε0 , |y − γ −k y − | ≤ γ −k ε1 ,
x̄ = λk x , ū = λk2 u , ȳ = γ k y .
Fig. 3.8.7. The Poincaré map in a neighborhood of a saddle in R 3 . The images of the
points lying in the intersection of the two-dimensional stable manifold W s with the three-
dimensional area Π+ compose the edge of a wedge. The part of Π+ above W s is transformed
into the wedge itself. The closer the dashed area of Π+ is to W s the thinner and closer to
W u is its image inside Π− .
W u ∩ Π− = {x = 0, u = 0, |y − y − | ≤ ε1 }
and that they have the shape of vertical “bars” located inside a three-
dimensional wedge
x̄ > 0 , C2 x̄α ≤ ū ≤ C1 x̄α , |ȳ − y − | ≤ ε1 , α = ln |λ2 |/ ln λ ,
W u ∩ Π− = {x = 0, u = 0, |y − y − | ≤ ε1 } .
3.9. Geometrical properties of nonlinear saddle maps 181
Since α > 1 and C1,2 6= ∞, the wedge is tangent to the extended unstable
subspace E u ⊗ E sL : {u = 0} at the points of W u ∩ Π− as shown in Fig. 3.8.7.
In the case where W s is one-dimensional and W u is two-dimensional, the
map T can be written as
x̄ = λx , ȳ = γy , v̄ = γ2 v ,
where |γ2 | > |γ|. This case is reduced to the previous one if we consider the
inverse map T −1 . If we select the points M + ∈ W s and M − ∈ W u /W uu
and select their neighborhoods Π+ and Π− respectively, then the range of the
map T 0 : Π+ → Π− consists of a countable union of non-intersecting three-
dimensional plates σk1 converging to the square W u ∩ Π− . At the same time,
the domain of the map is a union of a countable number of three-dimensional
horizontal bars σk0 within the wedge
y > 0 , C̃2 y α < v < C̃1 y α , |x − x+ | ≤ ε0
W s ∩ Π+ = {y = 0, v = 0, |x − x+ | ≤ ε0 }
The results of the previous section have a primarily illustrative character. It is,
therefore, important that the geometrical structures considered in the linear
case persist for generic nonlinear maps.
182 Chapter 3. Structurally Stable Periodic Trajectories
Fig. 3.8.8. The map near saddle of other topological type, i.e. with a one-dimensional stable
manifold W s and a two-dimensional unstable manifold W u . This situation may be regarded
as inverse to the map in Fig. 3.8.7.
Near a saddle fixed point a nonlinear map T can be written in the form
x̄ = AsL x + f1 (x, u, y, v) ,
ū = Ass u + f2 (x, u, y, v) ,
(3.9.1)
ȳ = AuL y + g1 (x, u, y, v) ,
v̄ = Auu v + g2 (x, u, y, v) ,
where x and y are the leading coordinates, and u and v are the non-leading
coordinates. The absolute values of the eigenvalues of the matrix AsL are
3.9. Geometrical properties of nonlinear saddle maps 183
equal to λ (0 < λ < 1), those of the matrix Ass are less than λ, those of the
matrix AuL are equal to γ (γ > 1) and those of the matrix Auu are greater
than γ. The functions f and g along with their first derivatives vanish at
the origin. We suppose that in some sufficiently small neighborhood U of the
saddle point O the invariant stable and unstable manifolds are straightened,
s
i.e. f (0, 0, y, v) ≡ 0 and g(x, u, 0, 0) ≡ 0. The equation of the manifold Wloc is
u
then (y = 0, v = 0) and that of Wloc is (x = 0, u = 0).
We assume that the stable and unstable leading multipliers of the saddle
fixed point O are simple (namely, a real leading eigenvalue, or a pair of complex-
conjugate leading eigenvalues).
Let M + (x+ , u+ , 0, 0) and M − (0, 0, y − , v − ) be arbitrary points on the sta-
ble and unstable manifolds of the saddle such that neither point lies in the
corresponding non-leading manifolds. Let Π+ and Π− be sufficiently small
rectangular neighborhoods of M + and M − respectively:
Π− = {kxk ≤ ε1 , kuk ≤ ε1 , ky − y − k ≤ ε1 , kv − v − k ≤ ε1 }
Lemma 3.6. If identities (3.9.4) hold and if the leading eigenvalues are simple
(real or complex), then
are asymptotically much stronger than that in the leading coordinates). Esti-
mates (3.9.5) imply that in the leading coordinates the geometrical structure is
determined mainly by the linear terms of the map T : the strips belong to the
roulettes if the leading multipliers are complex; if the stable or unstable lead-
ing multiplier is real, then the corresponding strips accumulate, respectively,
u s
to Wloc or Wloc from one side if the multiplier is positive and from the both
sides if it is negative.
Note that we derive here this picture based on Lemma 3.6 which is valid
only for maps of class Cr with r ≥ 2. To prove that the same geometry persists
in the C1 -case, one may use the modified boundary-value problem introduced
in Sec. 5.2.
We must note that Lemma 3.6 may not hold unless one performs the pre-
liminary reduction of the map T to the special form where the functions f and
g satisfy condition (3.9.4). Let us show this on the following example.
Consider a three-dimensional map T0 of the following form
x̄ = λx, ū = λ2 u + xy, ȳ = γy ,
where 0 < λ2 < λ < 1 < γ. Here O(0, 0, 0) is a saddle fixed point. The
equation of the two-dimensional stable invariant manifold W s is y = 0, and
that of the one-dimensional unstable invariant manifold W u is x = u = 0. The
equation of the non-leading stable invariant manifold W ss ∈ W s is y = x = 0.
The boundary-value problem for the map T0 reads: Given the initial data
(x0 , u0 , yk ) and given k, find (xk , uk , y0 ) such that T0k (x0 , u0 , y0 ) = (xk , uk , yk ).
We can recast the system in the form
xj = λ j x0 ,
j−1
X
uj = λj2 u0 + λj−s−1
2 · λs x0 · γ s−k yk ,
s=0
(3.9.7)
j−k
yj = γ yk ,
(j = 0, 1, . . . , k) .
From (3.9.7) we see that xk = λk x0 and y0 = γ −k yk . Yet
k−1
X
uk = λk2 u0 + λk−s−1
2 · λs x0 · γ s−k yk
s=0
k−1
X λγ s
= λk2 u0 + λk−1
2 γ −k x0 yk .
s=0
λ2
186 Chapter 3. Structurally Stable Periodic Trajectories
Since
λγ
δ= > 1,
λ2
the coefficient
k−1
X λγ s δk − 1 λk−1
= ∼ k−1 γ k−1 .
s=0
λ2 δ−1 λ2
ẏ = A(θ)y + F (θ, y) ,
(3.10.2)
θ̇ = 1 ,
ẏ = A(θ)y + F̃ (θ, y) ,
(3.10.4)
θ̇ = 1 + b(θ, y) ,
where
F̃ (θ, 0) = 0, F̃y0 (θ, 0) = 0, b(θ, 0) = 0 .
Proof of the theorem. The original system may be reduced to the form
(3.10.2) in the following way. At each point Mθ (x = ϕ(θ)) we choose (n + 1)
linearly independent vectors (N0 (θ), N1 (θ), . . . , Nn (θ)), where N0 (θ) ≡ ϕ0 (θ) =
X(ϕ(θ)) is the velocity vector which is tangent to the periodic trajectory L at
Mθ . Assume that Ni (θ) (i = 1, . . . , n) are smooth functions of θ. Let Mθ be
the space spanned on (N1 (θ), . . . , Nn (θ)), i.e. the space Mθ is transverse to L.
Let us denote the coordinates in the space Mθ with the basis (N1 (θ), . . . ,
Nn (θ)) by (y1 , . . . , yn ). If a point M ∈ Mθ has coordinates the (y1 , . . . , yn ),
then the vector connecting the points Mθ and M (see Fig. 3.10.1) is given by
Mθ M = y1 N1 (θ) + · · · + yn Nn (θ) .
Thus, the original coordinates x of the point M are given by the formula
or by
where Nij is the j-th component of the vector Ni , and ϕj is the j-th coordinate
of the point M on the periodic trajectory L.
Formula (3.10.5) can be viewed as a smooth change of variables (θ, y 1 , . . . ,
yn ) ↔ (x1 , . . . , xn+1 ). In order to show that this is really a good change of
188 Chapter 3. Structurally Stable Periodic Trajectories
Fig. 3.10.1. The normal coordinates near a periodic trajectory. The vectors N i (θ) in the
cross-section S are orthogonal to the velocity vector N0 .
variables, one must check the non-singularity of the Jacobian matrix J. The
value of vector y for the points on L is equal to zero, i.e. (y1 , y2 , . . . , yn ) =
(0, 0, . . . , 0), and since we are concerned with a small neighborhood of the
periodic trajectory, it is sufficient to verify that J does not vanish at y = 0.
From (3.10.6) we obtain
Xn
0 0
The first column of J(θ, 0) is composed of the components of the vector N 0 (θ);
the remaining columns are the components of the vectors Ni (θ). By construc-
tion, these vectors must be linearly independent for all θ’s, i.e. the Jacobian
matrix J(θ, 0) is non-singular.
Let us write system (3.10.1) in the new variables. Substitution of (3.10.5)
into (3.10.1) gives
where η(θ) is a Cr+1 -smooth scalar function which is equal identically to zero
in a small neighborhood of θ = 0, and is equal identically to 1 in a small
neighborhood of θ = τ (this also implies that η 0 (θ) ≡ 0 for values of θ close
to 0, and to τ ). We assume, as well, that Ã0 (θ) = 0 for θ close to 0 and to
3.10. Normal coordinates 191
τ . Thus, Yθ is an identity map at all θ close to zero and it coincides with the
Poincaré map T at all θ close to τ .
Since F00 (0) = 0 and Ã(θ) is non-degenerate for all θ, the map (3.10.13) is
invertible, i.e.
y0 = Ã−1 (θ)yθ + F1 (θ, yθ ) , (3.10.14)
0 0
where F1 (0, 0) = 0, F1y (θ, 0) = 0 and F1θ ≡ 0 for all θ close to 0 and τ .
Let us make a coordinate transformation (θ, y1 , . . . , yn ) → (x1 , . . . , xn+1 )
by the following rule
x = X (θ, Yθ−1 (y)) . (3.10.15)
In other words, we identify the time θ shift of the point y0 ∈ S along the
trajectories of system (3.10.1) and the time θ shift of y0 which is given by
(3.10.13). When θ is close to 0, Eq. (3.10.15) reads as
x = X (θ, y) , (3.10.16)
x = X (θ − τ, y) .
Comparing the latter with (3.10.16) we get that the Cr -smooth transformation
of coordinates (3.10.15) is τ -periodic.
The evolution of the new y-coordinates is given by (3.10.13), where θ̇ = 1.
From (3.10.14) we have
Denoting
Lemma 3.7. All derivatives of the right-hand side of (3.10.2) with respect to
y are Cr -smooth functions of θ at y = 0.
Let Φ(t) be the fundamental matrix of solutions of the system (3.10.21), i.e. the
solution of system (3.10.21) has the form
It is seen from (3.10.23) that each successive approximation differs from the
previous one by terms of higher orders; namely
Z t
(m) (m−1)
y (t) − y (t) = Φ(t) Φ−1 (s)(F (s, y (m−1) (s)) − F (s, y (m−2) (s))) ds
0
whence
where the function Ψ(y) vanishes together with its first derivatives when y = 0.
The linear part of the Poincaré map has the form
ȳ = Φ(τ ) y .
194 Chapter 3. Structurally Stable Periodic Trajectories
The problem of the stability of a periodic trajectory does not differ essentially
from the corresponding problem for equilibrium states. In both cases the
stability conditions are determined by the equations of the first approximation,
i.e. by the associated linearized system for an equilibrium state, or the so-called
variational equation for a periodic trajectory .
Let x = ϕ(t) be a periodic solution of period τ of an (n + 1)-dimensional
autonomous system
ẋ = X(x) . (3.11.1)
Introduce a new variable ξ such that
x = ξ + ϕ(t) .
In terms of the new variable the system takes the following form
ξ˙ = D(t)ξ + · · · ,
where
∂X
D(t) =
∂x x=ϕ(t)
and the ellipsis denotes terms of a higher order with respect to ξ. Observe that
this change of variables reduced an (n + 1)-dimensional autonomous system to
an (n + 1)-dimensional non-autonomous system.
The linear periodic system
ξ˙ = D(t)ξ (3.11.2)
is called a variational equation. Obviously, if ξ(t) is a solution of (3.11.2), then
ξ(t + τ ) is also a solution. Indeed, after the shift of time t → t + τ we obtain
dξ(t + τ )
= D(t + τ )ξ(t + τ ) ,
d(t + τ )
3.11. The variational equations 195
and, consequently,
dξ(t + τ )
= D(t)ξ(t + τ ) .
dt
The general solution of (3.11.2) is
ξ(t) = Ψ(t)ξ(0) (3.11.3)
where Ψ(t) is the fundamental matrix, whose columns Ψ(i) (t) (i = 1, . . . , n + 1)
are the solutions of (3.11.2) which start at t = 0 with basis vectors. Since
Ψ(i) (t + τ ) is a solution as well, it follows from (3.11.3) that Ψ(i) (t + τ ) =
Ψ(t)Ψ(i) (τ ), or
Ψ(t + τ ) = Ψ(t)Ψ(τ ) . (3.11.4)
The equation
|Ψ(τ ) − ρI| = 0 (3.11.5)
is called a characteristic equation. The roots (ρ1 , . . . , ρn+1 ) of (3.11.5) are
called the characteristic roots or Floquet multipliers.
The characteristic equation is invariant with respect to any change of vari-
ables
η = Q(t)ξ , (3.11.6)
where the matrix Q(t) is non-singular for all t, depends smoothly on time and
is periodic of period τ . Indeed, after this change of variables system (3.11.2)
remains linear periodic system. Denote its fundamental matrix as Ψ̃(t); i.e. the
general solution is η(t) = Ψ̃(t)η(0). By (3.11.3), (3.11.6) we have Ψ̃(t) =
Q(t)Ψ(t)Q(0)−1 . Thus, by virtue of τ -periodicity of Q(t), the matrix Ψ̃(τ ) is
similar to Ψ(τ ):
Ψ̃(τ ) = Q(0)Ψ(τ )Q(0)−1 .
Thus,
|Ψ̃(τ ) − ρI| = |Q(0)Ψ(τ )Q(0)−1 − ρI| = |Q(0)(Ψ(τ ) − ρI)Q(0)−1 | = |Ψ(τ ) − ρI|
which proves the claim.
It follows that the Floquet multipliers of a periodic trajectory do not depend
on the specific choice of coordinates x. Indeed, let y = h(x) be a diffeomor-
phism transforming the system (3.11.1) into ẏ = Y (y) in some small neighbor-
hood of the periodic trajectory L. In the new variables the equation of L is
y = h(ϕ(t)) = ψ(t). The variational equation for ψ(t), which is now given by
∂Y
η̇ = η,
∂y y=ψ(t)
196 Chapter 3. Structurally Stable Periodic Trajectories
It is easy to see that the fundamental matrix of this system has the form
!
Φ(t) 0
,
β(t) 1
η̇ = A(t)η ,
ϕ̇(t) = X(ϕ(t)) ,
which are called the characteristic exponents. Observe from (3.11.7) that λ k is
defined modulo i2πmk /τ , where mk is an integer. However, (Reλ1 , . . . , Reλn+1 )
are uniquely defined. They are called the Lyapunov exponents of the periodic
trajectory x = ϕ(t).
These quantities have sense for any linear periodic system of type (3.11.2).
Recall that in case the variational equation is obtained from an autonomous
system there is always a trivial characteristic root, hence one Lyapunov expo-
nent is always zero in this case.
d
The columns of the fundamental matrix satisfy (3.11.2), i.e. dt Ψ(t) =
D(t)Ψ(t). Hence,
d
det Ψ(t) = tr D(t) · det Ψ(t)
dt
which gives the Wronsky formula
Rt
tr D(s)ds
det Ψ(t) = e 0 .
At t = τ we obtain Rτ
tr D(s)ds
ρ1 ρ2 · · · ρn+1 = e 0 , (3.11.8)
where (ρ1 , . . . , ρn+1 ) are the characteristic roots. It is clear that all ρ1 , . . . , ρn+1
are different from zero and that Ψ(τ ) is non-singular.
When the linear system (3.11.2) is obtained from the autonomous system
(3.11.1) this formula reads as
Rτ
div X|x=ϕ(s) ds
ρ1 ρ2 · · · ρ n = e 0 , (3.11.9)
or Z τ
1
λ1 + · · · + λ n = div X|x=ϕ(s) ds . (3.11.10)
τ 0
Note that in the general case, finding the fundamental matrix of the vari-
ational equation or its characteristic roots in the explicit form is not possi-
ble. The two-dimensional case is the only exception. In this case the formula
198 Chapter 3. Structurally Stable Periodic Trajectories
Therefore,
ξ (k) (t) = φk (t)eλk t , (3.11.11)
where φk (t) is a periodic function.
A more general statement also holds, known as Floquet theorem [24]:
the fundamental matrix Ψ of a linear time periodic system satisfies
where
A◦kk = ρ̃k if ρ̃k is real
! ! ! !−1
A◦kk A◦k,k+1 1 1 reiφ 0 1 1
=
A◦k+1,k A◦k+1,k+1 −i i 0 re−iφ −i i
!
r cos φ −r sin φ
= if (ρ̃k = reiφ , ρ̃k+1 = re−iφ ),
r sin φ r cos φ
(3.11.15)
all the other entries of A◦ are zero; the only non-zero entries in ∆A may be
s=1
s
where
(ln A◦ )kk = ln ρ̃k if ρ̃k is real,
! !
(ln A◦ )kk (ln A◦ )k,k+1 ln r −φ
= if ρ̃k = ρ̃∗k+1 = reiφ .
(ln A◦ )k+1,k (ln A◦ )k+1,k+1 φ ln r
(3.11.18)
◦
The formula (3.11.18) gives a real-valued matrix ln A because all real ρ̃k
are positive by construction; the matrix (A◦ )−1 exists because all ρk are non-
zero by (3.11.8); the series in (3.11.17) is convergent because (∆A)s ≡ 0 for
sufficiently large s. The expansion for ln(A◦ +∆A) in (3.11.17) is a calque of the
Taylor expansion for a scalar logarithm: the scalar arithmetic is applied here
because the matrices ln A◦ , A◦ and ∆A commute, i.e. A◦ · ln A◦ = ln A◦ · A◦ ,
A◦ · ∆A = ∆A · A◦ , ∆A · ln A◦ = ln A◦ · ∆A (see (3.11.15), (3.11.16),(3.11.18)).
Let us now take
1
Λ̃ = ln à ,
τ
so
Ψ(τ ) = σeΛ̃τ .
3.12. Stability of periodic trajectories. Saddle periodic trajectories 201
Definition 3.2. A solution x = ϕ(t) of system (3.12.1) is called stable (in the
sense of Lyapunov) if given arbitrary small ε > 0 there exists δ such that if
kx0 −ϕ(t0 )k ≤ δ, then kx(t)−ϕ(t)k ≤ ε for all t ≥ t0 , where x(t) is the solution
with the initial condition x0 .
Lyapunov proved this theorem for the case where the system (3.12.1) has
an analytic right-hand side, though it also holds when the function F is only
of C1 -smoothness with respect to x and continuous with respect to t.
In a small neighborhood of x = ϕ(t) system (3.12.1) can be brought to the
form (see Sec. 3.11)
ẏ = Λy + G(y, t) , (3.12.2)
where Λ is a constant matrix such that the real parts of its eigenvalues are
the Lyapunov exponents (Re λ1 , . . . , Re λn ). The function G(y, t) is periodic of
period τ or 2τ with respect to t. Moreover, G(0, t) ≡ 0, G0y (0, t) ≡ 0. It follows
that kG0y k is uniformly bounded by a small constant for all t and for all small
y. After reducing the system to the form (3.12.2) the proof of Theorem 3.13
repeats the proof of the theorem on the validity of the linearization near a
stable equilibrium state (Theorem 2.4).
Let us now consider an (n + 1)-dimensional autonomous system
ẋ = X(x) (3.12.3)
is always equal to zero. Therefore, from the point of view of Lyapunov stabil-
ity this situation corresponds to the critical case. Nevertheless, the following
theorem is valid.
This theorem justifies the linearization but only for a very weak form of
stability. The problem is as follows. Let L be the corresponding periodic tra-
jectory: L = {x : x = ϕ(θ), 0 ≤ θ ≤ τ )}. Then for any two neighboring points
on L, the associated solutions of (3.12.3) cannot approach asymptotically each
other as t → +∞. It is easily seen when the system is written in the normal
coordinates near L. Recall that in the normal coordinates (y, θ) near L where
kyk measures the distance to L and θ ∈ S1 is the angular variable, the system
recasts in the form (3.11.20) where kFy0 k is uniformly bounded by a small con-
stant for all t and for all small y. The real parts of the eigenvalues of Λ are the
non-trivial Lyapunov exponents of L. Like in the Lyapunov Theorem 3.13, if
all the eigenvalues have negative real parts, then
Theorem 3.15. (On the asymptotical phase) Let all non-trivial characteristic
exponents of a periodic trajectory L lie to the left of the imaginary axis. Then,
given sufficiently small ε there exists δ such that if kx0 − ϕ(θ0 )k < δ, then there
exists ψ, |ψ| < ε such that the solution x(t), x(0) = x0 , satisfies the inequality
where
dist(x, L) = inf kx − ϕ(θ)k .
0≤θ≤τ
Theorem 3.16. If all multipliers of the periodic trajectory L lie inside the
unit circle, then L is orbitally stable as t → +∞, and satisfies the following
estimate
dist(x(t), L) ≤ Ke−λt ,
where K and λ are positive constants.
This theorem follows directly from the theorem on the stability of the fixed
point of the Poincaré map and the theorem on the continuous dependence of
the solution on the initial conditions (or it immediately follows from (3.12.4))
because the multipliers of L lie inside the unit circle if and only if the non-
trivial characteristic exponents of L lie to the left of the imaginary axis (see
Sec. 3.11).
3.12. Stability of periodic trajectories. Saddle periodic trajectories 205
The case where all the multipliers of the periodic trajectory L are outside of
the unit circle is reduced to the above case by means of reversion t → −t. All
trajectories in a small neighborhood of such an unstable periodic trajectory
leave the neighborhood as t increases. The time, over which a trajectory
escapes from the neighborhood, depends on the position of the initial point
of the trajectory with respect to L, the closer the point is to the periodic
trajectory the larger the escaping time is. Such unstable periodic trajectories
are called completely unstable or repelling.
Let us consider next the case where some multipliers of L, (ρ1 , . . . , ρk ) lie
inside and the rest (ρk+1 , . . . , ρn ) lie outside of the unit circle.
In a neighborhood of the periodic trajectory L of period τ the system is
written in the form (3.10.2). Dividing the first equation of (3.10.2) by the
second we obtain the non-autonomous system with the periodic right-hand
side
dy
= A(θ)y + F (y, θ) , (3.12.6)
dθ
where F (0, θ) ≡ 0, Fy0 (0, θ) ≡ 0. Just like in Sec. 3.10 we can integrate (3.12.6)
with the initial data (y 0 , 0) and find a solution y = y(θ; y 0 ) which is Cr -smooth
with respect to both arguments. If we let θ = τ , we obtain the Poincaré map
where Ψ(0) = 0, and Ψ0 (0) = 0. The roots of the equation |Φ(τ ) − ρI| = 0 are
the multipliers (ρ1 , . . . , ρn ) of L.
It follows from Hadamard’s theorem (see Sec. 3.6) that two smooth invari-
s u
ant manifolds, stable Wloc (O) and unstable Wloc (O), pass through the point
O at the origin. These manifolds are tangent, respectively, to the stable k-
dimensional subspace and the (n − k)-dimensional unstable subspace of the
associated linearized map ȳ = Φy at the point O which we denote as E s and
as E u . Let y = (y1 , y2 ) where y1 ∈ Rk and y2 ∈ Rn−k . Let y2 = C s y1 be the
equation of E s , and let y1 = C u y2 be the equation of E u , where C s and C u
are some matrices.8 Thus, the equation of Wloc s
(O) is given by
∂ψ(0)
y2 = ψ(y1 ), where = Cs ,
∂y1
8 To obtain the equations of the stable and unstable manifolds of the saddle periodic
trajectory L there is no necessity to reduce the Poincaré map to the special form (3.6.1),
i.e. we do not assume that the linear part of the Poincaré map decouples into two equations.
206 Chapter 3. Structurally Stable Periodic Trajectories
u
and the equation of Wloc (O) is given by
∂φ(0)
y1 = φ(y2 ), where = C u.
∂y2
1. Let 0 < ρ1 < 1 and ρ2 > 1. The saddle fixed point O breaks the stable
and the unstable manifolds into two components each, such that
s
Wloc (O) = Γ1 ∪ O ∪ Γ2
and
u
Wloc (O) = Γ3 ∪ O ∪ Γ4 .
Moreover, each Γi , (i=1,. . . ,4) is invariant, i.e. it is taken into itself by the
s u
Poincaré map. Hence, Wloc (L) and Wloc (L) are smooth two-dimensional
surfaces which are homeomorphic to a cylinder, see Fig. 3.12.1. Observe
that the three-dimensional system (3.12.2) near the periodic trajectory
L may then be reduced to the following form (see Sec. 3.11)
3.12. Stability of periodic trajectories. Saddle periodic trajectories 207
Fig. 3.12.1. A saddle periodic trajectory with two-dimensional stable W s and unstable
W u manifolds which are homeomorphic to a cylinder.
ẏ1 = λ1 y1 + f1 (y1 , y2 , θ) ,
ẏ2 = λ2 y2 + f2 (y1 , y2 , θ) , (3.12.10)
θ̇ = 1 ,
ln ρ1,2
where λ1,2 = τ .
2. Let |ρ1 | < 1 and |ρ2 | > 1, moreover, ρ1 < 0 and ρ2 < 0. In this case the
s
Poincaré map takes Γ1 into Γ2 , and Γ2 into Γ1 . The manifold Wloc (L)
will be then diffeomorphic to a two-dimensional Möbius band. The same
u
is true for Wloc (L), see Fig. 3.12.2. In this case L is a middle line of the
Möbius band.
Thus, we can see that in the three-dimensional case saddle periodic tra-
jectories may be of two different topological types because there is no homeo-
morphism between cylinders and Möbius bands. An analogous situation holds
in the high-dimensional case. If sign (ρ1 × · · · × ρk ) = 1, which implies also
s
that sign (ρk+1 × · · · × ρn ) = 1, then Wloc (L) is homeomorphic to the multi-
dimensional cylinder D × S , and Wloc (L) is homeomorphic to Dn−k × S1 .
k 1 u
208 Chapter 3. Structurally Stable Periodic Trajectories
Fig. 3.12.2. A saddle periodic trajectory with two-dimensional stable W s and unstable W u
manifolds which are homeomorphic to a Möbius band.
s
If sign (ρ1 × · · · × ρk ) = −1 and sign (ρk+1 × · · · × ρn ) = −1, then Wloc (L) and
u
Wloc (L) are non-orientable manifolds of the type of multi-dimensional Möbius
bands (i.e. they are represented as a fiber bundle of Dk and, respectively,
Dn−k by S1 ). Similarly to the classification of structurally stable fixed points,
we can distinguish structurally stable periodic trajectories by introducing the
Qk Qn
invariants δs = sign ρi and δu = sign ρi .
i=1 i=k+1
Up to now we have been talking about the local manifolds of periodic
trajectories. However, we can define these manifolds globally. Let x = X(t; x 0 )
denote the trajectory with initial condition x0 .
Definition 3.4. The stable manifold of the periodic trajectory L is the set
The unstable manifold WLu is defined in a similar way with the difference that
t → −∞.
By that definition, for any point x ∈ W s (L) there is a moment of time at
which the trajectory of x enters a small neighborhood of L, so some time shift
of x belongs to the local stable manifold. Thus,
[
W s (L) = WLs (t∗ ) ,
t∗ ≤0
where
Since WLs (t∗ ) is a smooth image of WLs loc by X(t∗ ; ·), it follows that WLs is a
smooth image of either a cylinder Rk × S1 or a Möbius band.
3.13. Smooth equivalence and resonances 209
where
x̄ = Ax + f (x) , (3.13.1)
ρk = ρ m , (3.13.2)
where ρm = (ρm 1 m2 mn
1 ρ2 · · · ρn ), mk (k = 1, . . . , n) are some non-negative inte-
Pn
gers such that |m| = mk ≥ 2. The number |m| is called the order of the
k=1
resonance.
Lemma 3.8. Let the function f (x) ∈ CN and let there be no resonances of
the order |m| ≤ N . Then, the change of variables
ȳ = Ay + oN (y), (3.13.4)
where oN (y) vanishes at the origin along with its derivatives up to order N .
210 Chapter 3. Structurally Stable Periodic Trajectories
It is obvious that the changes of variables above and below, are local,
namely they are valid only in some small neighborhood of a fixed point of
diffeomorphism (3.13.1)
Lemma 3.8 is well known and it is valid even when A has multiple eigen-
values. Here, we will discuss only the case of simple eigenvalues (the extension
onto the general case is made in the same way as in Lemma 2.2). The matrix
A can then be represented in the form
ρ1 0
ρ2
A= .. .
.
0 ρn
Let us recast the function f (x) into the following form
where the ellipsis denotes the terms of degree higher than N (observe that other
summands above also contain the terms of degree (N + 1) and higher). The
process of eliminating the redundant terms begins with the quadratic terms.
In order to find ϕ2 (x) we write the following equation
k m1 +···+mn =2
where “{}” denotes the cubic terms. Since we have already found ϕ2 (x), we
have the following equation for the unknown coefficients of ϕ3 (x):
Here, d˜km = dkm + d0km , where d0km is the coefficient of xm in the k-th com-
ponent of the vector polynomial ϕ2 (Ax + f2 (x)). By repeating this procedure
we eliminate all terms up to degree N .
In the case where there are resonances of the form ρk = ρm one cannot
kill the monomials of the type d˜km xm ek . For this case we have the following
well-known lemma.
y = x + ϕ(x) ,
where
|m|≤N
X
RN (x) = bkm xm ek . (3.13.11)
ρm =ρ k
212 Chapter 3. Structurally Stable Periodic Trajectories
In the case where there are resonances the following theorem is valid.
ȳ = Ay + R(y) , (3.13.17)
It follows from these two theorems that the dependence of the normal forms
on the collection ρ = {ρ1 , . . . , ρn } has a discontinuous character.
Just like in the case of vector fields we may pose a question concerning the
reduction of diffeomorphisms to a linear form by changes of variables of only
finite smoothness.
where {i, j, k} ∈ (1, . . . , n). Then system (3.13.1) may be transformed into
linear form by a C1 -smooth change of variables.
x̄ = A1 (µ)x + f1 (x, y, u, v, µ) ,
ū = A2 (µ)u + f2 (x, y, u, v, µ) ,
(3.13.19)
ȳ = B1 (µ)y + g1 (x, y, u, v, µ) ,
v̄ = B2 (µ)v + g2 (x, y, u, v, µ) ,
9 Poincaré was the first to discover the existence of such trajectories in problems of Hamil-
tonian dynamics.
3.13. Smooth equivalence and resonances 215
lie strictly inside the unit circle, and that the eigenvalues of the matrix
!
B1 (0) 0
B(0) =
0 B2 (0)
lie outside of the unit circle. Assume also that the eigenvalues (ρ 1 , . . . , ρm1 ) of
the matrix A1 (0) satisfy the conditions |ρi | = ρ < 1 (i = 1, . . . , m1 ), and the
eigenvalues (γ1 , . . . , γp1 ) of the matrix B1 (0) satisfy |γi | = γ > 1 (i = 1, . . . , p1 ).
With regard to the eigenvalues (ρm1 +1 , . . . , ρm ) of the matrix A2 (0) and the
eigenvalues (γp1 +1 , . . . , γp ) of the matrix B2 (0) we will assume that
|ρi | < ρ, i = m1 + 1, . . . , m
(3.13.20)
|γi | > γ, i = p1 + 1, . . . , p .
Hence, the fixed point O is of the saddle type, the x and y coordinates are,
respectively, the leading stable and the leading unstable coordinates.
Theorem 3.22. Under the above assumptions there exists a Cr−1 -smooth
change of variables which transforms the family (3.13.19) into
where fij (x, y, u, v, µ) and gij (x, y, u, v, µ) (i, j = 1, 2) are Cr−1 -functions which
vanish at the origin and satisfy
ρk = ρ m
k , m≥2 (3.13.22)
when ρk = 1; and
ρk = ρ2m+1
k , m≥1 (3.13.23)
when ρk = −1; and
This diffeomorphism has a fixed point O(0, 0) with the Jacobian matrix
!
A h0ε (0, 0)
à = ,
0 I
3.13. Smooth equivalence and resonances 217
ρk = ρ m ,
where
p
X
ρm = ρ m mp
1 · · · ρp ,
1
mi ≥ 2 ,
i=1
ρk = ρ k γ l , (3.13.27)
ρk = ρ m γ l , (3.13.28)
γk = γ l , (3.13.29)
where
q
X
γ l = γ1l1 · · · γqlq , lj ≥ 2 .
j=1
The reduction of system (3.13.26) to normal form can be achieved via the
change of variables
y = x + ϕ(x, ε)
(3.13.30)
ε=ε
which leaves the second equation in (3.13.26) unchanged (the latter means that
we do not need to consider the resonances of the kind (3.13.29)). Similar to
the case in Lemma 3.9, the original family may be transformed into
where bkm (ε) are certain polynomials of degree not exceeding (N −|m|). More-
over, R0 (ε) ≡ 0 if, among the eigenvalues (ρ1 , . . . , ρp ), there is none equal to
one. Otherwise, R0 (ε) is a polynomial of degree not higher than N , and
218 Chapter 3. Structurally Stable Periodic Trajectories
R0 (0) = 0. The appearance of the term R0 (ε) in (3.13.31) is due to the exis-
tence of resonances of the kind
ρk = γ l . (3.13.33)
for some suitable choice of N and p. Just like the case of vector field, the in-
formation extracted from the analysis of the truncated normal form (3.13.34)
must be substantiated before it can be applied to the original family of dif-
feomorphisms. This is the method used in our study of the main cases of
bifurcations of periodic trajectories in the second part of this book.
reduced to
X
ẏk = λk yk + δk yk+1 + Fkm (θ)y m + o(kykr )
2≤|m|≤r (3.14.2)
(k = 1, . . . , n)
ρm mn
1 · · · · · ρ n = ρk ,
1
2πi
m1 λ 1 + · · · + m n λ n = λ k + mn+1 , (3.14.4)
τ
where mn+1 is also an integer, and may possibly be negative.
0
change the coefficients of the monomials y m ek0 of orders lower than the order
of y m ek (in the sense of Lemma 2.2; see (2.9.18)). Therefore, increasing (in
the above sense) multiindex (k, m) in (3.14.5) will finally give us the theorem.
Equating the coefficients of y m in the identity
d new d
y = [yk + fkm (θ)y m ]
dt k dt
we obtain
new old 0
Fkm (θ) = Fkm (θ) + (fkm (θ) + fkm (θ)[(m, λ) − λk ]) .
where
γkm = (m, λ) − λk .
One can see that if the monomial y m is non-resonant (i.e. γkm 6= 2πi τj ), then
the constant C in (3.14.6) may be taken such that fkm is a τ -periodic function
new
of θ with Fkm ≡ 0. Indeed, the condition of periodicity of fkm is
(here we used the τ -periodicity of Fkm (t)). If γkm 6= 2πi τj , then the coefficient
of C is non-zero, and the required C is immediately found.
For the resonant case we have two possibilities: γkm = 0 and γkm 6= 0. If
γkm = 0 (mn+1 = 0 in terms of (3.14.4)), Eq. (3.14.6) takes the form
Z θ
new old
fkm (θ) = C + {Fkm (t) − Fkm (t)}dt . (3.14.7)
0
3.14. Autonomous normal forms 221
R τ old
One can see immediately that Fkm new
= const. = τ1 0 Fkm (t)dt gives the τ -
periodic function fkm . Thus, we could reduce the system to the autonomous
normal form by a τ -periodic transformation, if the values γkm vanished for all
resonant monomials.
Although this is not the case in general, we will however prove that this can
be achieved if we consider the system as (Sτ )-periodic with some integer S ≥ 1.
The idea is that the characteristic exponents λk are not defined uniquely by
expression (3.14.3) because the logarithm is not a single-valued function. In
fact, we may write
1 jk
λk = ln ρk + 2πi
τ τ
where jk are arbitrary integers (if ρk = ρk+1 , chose jk+1 = jk to obtain
λk+1 = λk ). If we consider the system as (Sτ )-periodic, we have
1 jk jk
λnew
k = ln(ρSk ) + 2πi = λold
k + 2πi . (3.14.8)
Sτ Sτ Sτ
We will prove now that there exist integers S and j1 , . . . , jn such that in all
resonant relations the imaginary part11 will vanish simultaneously, when one
proceed to the new λk defined by formula (3.14.8) (the theorem then follow
immediately from the above discussion).
The resonant relation (3.14.4) is a particular case of the relation
m1 λ1 + · · · + mn λn + mn+1 iω = 0 (3.14.9)
11 This 2πi
is τ
mn+1 in (3.14.4).
222 Chapter 3. Structurally Stable Periodic Trajectories
We have
(1) (1)
m1 λ1 + · · · + m(1)
n λn + mn+1 iω = 0 ,
.. .. ..
. . .
(q) (q)
m1 λ1 + · · · + m(q)
n λn + mn+1 iω = 0 ,
This implies that in terms of these newly defined λ1 , . . . , λn , for the case
where the system is considered as (Sτ )-periodic with S calculated from
(3.14.10), the value γkm equals to zero for all resonant monomials. This com-
pletes the proof of the theorem.
The general meaning of this result is that the behavior of solutions which
remain in a small neighborhood of a periodic trajectory very much resembles
the behavior of the solutions in a small neighborhood of an equilibrium state
of an autonomous system. More precisely, if we reduce a system in the normal
coordinates to the autonomous form up to the terms of some order r, then the
trajectories in a sufficiently small ε-neighborhood of a periodic trajectory will
be close to the trajectories of the truncated autonomous system for a very long
period of time (of order, say, ε1r ). However, we must be cautious because as r
increases, the coordinate transformation which we have constructed may not
3.15. The principle of contraction mappings. Saddle maps 223
In this section we give a simple criterion for the existence of fixed points which
is based on the principle of the contraction mappings. This criterion, when
applied to the Poincaré map, gives conditions which guarantee the existence of
periodic trajectories. The principle of the contraction mappings is a rather gen-
eral mathematical result and its applicability is not restricted to the problem
of establishing the existence of periodic trajectories. In the following chapters
we will use an infinite-dimensional version of this principle (on the space of
continuous functions) while proving theorems on invariant manifolds.
1 ε(1 − K)
m> · ln .
| ln K| kT M − M k
kM ∗ − M ∗∗ k = kT M ∗ − T M ∗∗ k ≤ KkM ∗ − M ∗∗ k
and
∗ ∗
Tµ+∆µ Mµ+∆µ = Mµ+∆µ .
Hence
kMµ∗ − Mµ+∆µ
∗
k = kTµ Mµ∗ − Tµ+∆µ Mµ+∆µ
∗
k
≤ kTµ Mµ∗ − Tµ+∆µ Mµ∗ k + kTµ+∆µ Mµ∗ − Tµ+∆µ Mµ+∆µ
∗
k
≤ kTµ Mµ∗ − Tµ+∆µ Mµ∗ k + KkMµ∗ − Mµ+∆µ
∗
k,
whence
1
kMµ∗ − Mµ+∆µ
∗
k≤ kTµ Mµ∗ − Tµ+∆µ Mµ∗ k .
1−K
Since Tµ depends continuously on µ, the right-hand side of this last inequality
∗
tends to zero as ∆µ → 0, and, therefore, Mµ+∆µ → Mµ∗ as ∆µ → 0. End of
the proof.
The following criterion on the existence of fixed points of smooth mappings
follows immediately from the Banach principle.
kF 0 k ≤ K < 1 . (3.15.3)
Then, F (x) has a unique fixed point x∗ ∈ D such that all trajectories of F
converge to x∗ .
we have Z 1
x̄2 = x̄1 + F 0 (x1 + s(x2 − x1 ))(x2 − x1 )ds
0
226 Chapter 3. Structurally Stable Periodic Trajectories
and Z 1
kx̄2 − x̄1 k ≤ kF 0 kds · kx2 − x1 k .
0
Hence,
kx̄2 − x̄1 k ≤ Kkx2 − x1 k
i.e. F is a contraction mapping and it follows from Theorem 3.24 that it has
a unique fixed point in D.
Remark. Here, we have re-proved the well-known inequality
kF (x2 ) − F (x1 )k ≤ sup kF 0 k · kx2 − x1 k , (3.15.4)
D
Proof. Let us compute the first derivative dx∗ /dµ. Since x∗ is a fixed point,
x∗ = F (x∗ , µ) .
i.e.
(I − Fx0 )∆x∗ = Fµ0 ∆µ + o(k∆x∗ k) + o(k∆µk) ,
where I is the identity matrix. Since kFx0 k ≤ K < 1, it follows that (I − Fx0 )
is invertible. Therefore,
In order to overcome this difficulty we consider the map in the so-called cross
form.
x̄ = P (x, ȳ) ,
(3.15.7)
y = Q(x, ȳ) .
x̄ = F (x, y) ,
ȳ = G(x, y) .
228 Chapter 3. Structurally Stable Periodic Trajectories
whence
Here, the derivatives of the functions F and G are taken with respect to (x, y),
and those of P and Q are taken with respect to (x, ȳ). Equaling the coefficients
of dy and dx, we obtain
G0y = (Q0y )−1 ,
G0x = −(Q0y )−1 Q0x ,
(3.15.8)
Fy0 = Py0 (Q0y )−1 ,
Fx0 = Px0 − Py0 (Q0y )−1 Q0x
and
Q0y = (G0y )−1 ,
Q0x = −(G0y )−1 G0x ,
(3.15.9)
Py0 = Fy0 (G0y )−1 ,
Px0 = Fx0 − Fy0 (G0y )−1 G0x .
Observe that a smooth forward map does not always correspond to a
smooth cross-map. In the case where (Q0y )−1 is not defined, the smoothness of
the map T may be violated, or the map T may not even be a one-to-one map.
However, the results below remain valid for such map.
Definition 3.7. The map T defined in the cross form (3.15.7) by the smooth
functions P and Q on the direct product of the closed convex sets D1 and D2
(D1 ⊆ Rn , D2 ⊆ Rm ) is called a saddle map if:
kPx0 k◦ < 1, kQ0y k◦ < 1 ,
(3.15.10)
kPy0 k◦ kQ0x k◦ < (1 − kPx0 k◦ )(1 − kQ0y k◦ ) ,
where k · k◦ = sup k · k.
(x,y)∈D1 ×D2
3.15. The principle of contraction mappings. Saddle maps 229
x̄ = A− x, y = (A+ )−1 ȳ
such that the Spec A− lies strictly inside the unit circle and Spec A+ lies
strictly outside of it, is also of the saddle type. Here max{kA− k, k(A+ )−1 k} <
1, and D1 and D2 can be chosen to be certain balls in the x-space and the
y-space, respectively.
When the mapping T is written in the forward form, conditions (3.15.10)
are no longer symmetric.
and
kFy0 (G0y )−1 k◦ kG0x k◦ k(G0y )−1 k◦
≤ (1 − kFx0 k◦ − kFy0 (G0y )−1 k◦ kG0x k◦ ) · (1 − k(G0y )−1 k◦ ) .
Observe now that these inequalities follow from conditions (3.15.11).
The first two inequalities in conditions (3.15.11) mean that the mapping
T is expanding along the y-variables and contracting along the x-variables. If
the derivatives Fy0 and G0x were equal to zero as in the linear map considered
previously, then it would be sufficient for the map to be of the saddle type.
The last inequality in (3.15.11) simply means that the distortion induced by
Fy0 and G0x is not essential.
230 Chapter 3. Structurally Stable Periodic Trajectories
Proof. First of all observe that the fixed points of the forward map T and
those of the cross-map T × coincide. Therefore, we need only to show that T ×
is a contraction mapping and invoke Theorem 3.24.
Let us introduce in D1 × D2 the distance given by
and
kQ(x2 , ȳ2 ) − Q(x1 , ȳ1 )k ≤ kQ0x k◦ kx2 − x1 k + kQ0y k◦ kȳ2 − ȳ1 k
or
kx̄2 − x̄1 k ≤ kPx0 k◦ kx2 − x1 k + kPy0 k◦ kȳ2 − ȳ1 k ,
and
ky2 − y1 k ≤ kQ0x k◦ kx2 − x1 k + kQ0y k◦ kȳ2 − ȳ1 k ,
whence
kx̄2 − x̄1 k + Lky2 − y1 k
where
K = max{kPx0 k◦ + LkQ0x k◦ , L−1 kPy0 k◦ + kQ0y k◦ } .
By virtue of (3.15.13), K < 1, hence it follows that T × is a contraction
mapping. End of the proof.
One can show that the obtained fixed point is of the saddle type. In fact,
Theorem 4.2 from the next chapter can be applied here (both to the map T
3.15. The principle of contraction mappings. Saddle maps 231
and to its inverse T −1 ), so one can show that the fixed point of the saddle
map has a smooth stable and unstable manifolds in the form y = ψ(x) and
x = ϕ(y), where the functions ψ(x) and ϕ(y) are defined everywhere on D1
and D2 respectively.
Let us now discuss the abstract version of Banach principle. It is obvious,
that Theorem 3.24 remains valid if D is a closed subset of any Banach space X.
Recall, that a linear space X is called Banach space if it is complete; i.e. any
fundamental sequence {xi }∞ i=1 of elements of X converges: if for any there
exists m such that kxn+m − xm k ≤ for all n ≥ 0, then for some x∗ ∈ X
lim xi = x∗ .
i→∞
dist(x1 , x2 ) = kx1 − x2 k
The Hadamard theorem (Theorem 3.9) applies the Banach principle to the
operator ϕ 7→ ϕ̃ defined on the Banach space of continuous functions u = ϕ(v),
where v belongs to the δ-neighborhood of zero in Rn−k and u ∈ Rk , with the
norm
kϕk◦ = sup kϕ(v)k . (3.15.14)
kvk≤δ
X), and so on: the r-th derivative is an inductively defined linear operator
Y → (Y → (. . . (Y → X) . . .).
| {z }
r
Obviously, the r-th derivative f (r) can be considered to be a symmetric
polylinear operator Y r → X such that
1 (r)
f (y + ∆y) = f (y) + f 0 (y)∆y + · · · + f (y)(∆y)r + o(k∆ykr ) .
r!
The function f is Cr -smooth on D ⊆ Y if for each k ≤ r the k-th derivative
(k)
f (y) depends continuously on y and is uniformly bounded as an operator
Y k → X; i.e.
is finite.
For example, for any Cr -smooth function g defined on Rn , the operator
x(t) 7→ g(x(t)) acting on the space H of the continuous functions x(t)t∈[0,τ ] is
Cr -smooth. A bounded linear operator is Cr -smooth for any r. The superposi-
tion of smooth operators is an operator of the same smoothness. In particular,
the operator H → H which maps a continuous function x(t)t∈[0,τ ] into
Z t
x̄(t) = ψ(s)g(x(s), s)ds
0
INVARIANT TORI
ẋ = X(x) (4.0.2)
ẋ = X(x) + µp(x, θ) ,
(4.0.3)
θ̇ = 1 ,
235
236 Chapter 4. Invariant Tori
x̄ = f (x) , (4.1.2)
1 We remark that the study of systems with a piece-wise continuous right-hand side F (x, t)
having a finite number of discontinuity points on period can also be reduced to such a
diffeomorphism.
238 Chapter 4. Invariant Tori
ϕp (t) = ϕ0 (t + 2πp) ,
Brauer’s criterion is usually applied in the following situation. Let all in-
tegral curves of a system, which is defined in the region D × R1 enter this
region on the boundary D × R1 . Then the associated diffeomorphism satisfies
Theorem 4.1 and, consequently, the system itself has, at least, one periodic
trajectory.
Let us now return to the problem on periodically forced systems. In this
case, the study of (4.0.1) is reduced to that of a family of diffeomorphisms in
the form
x̄ = f (x, µ) , (4.1.3)
where f is represented as
ẋ = X(x) ,
(4.1.5)
θ̇ = 1 .
4.1. Non-autonomous systems 239
Fig. 4.1.2. An invariant torus T20 of the extended system (4.1.5) at µ = 0 is represented as
a direct product L0 × S1 .
Under the above assumption, the system (4.0.2) has a periodic solution L of
period τ , the equation of which is x = ϕ(t). Hence, system (4.0.3) will have “a
straight-edged” invariant torus T20 with a base defined by {L : x = ϕ(θ1 ), 0 ≤
θ1 ≤ τ }, as shown in Fig. 4.1.2. Therefore, diffeomorphism (4.1.3) has an
invariant smooth closed curve L0 at µ = 0. We will show below that if L0 is
a stable solution of (4.0.2), then for all µ sufficiently small, system (4.0.3) will
possess a smooth invariant torus T2µ close to T20 , see Fig. 4.1.3. This follows
from the fact that for all sufficiently small µ, diffeomorphism (4.1.3) will have
a smooth invariant closed curve Lµ .
Consider now the system
ẋ = X(x, t) , (4.1.6)
ẋ = X(x, θ) ,
(4.1.10)
θ̇ = Ω ,
4.2. Existence of an invariant torus. The annulus principle 241
x̄ = f (x, θ)
x̄ = F (x, θ̄) ,
(4.2.6)
θ = G(x, θ̄), (mod 2π) .
Observe that
It follows from this formula that the following estimates hold for the derivatives
of F and G:
−1
∂F ∂f ∂g ∂f ∂g
≤ + · ,
∂x ◦ ∂x ◦ ∂x ◦ ∂θ ∂θ
◦
−1
∂F ∂f ∂g
= ,
∂ θ̄ ◦ ∂θ ∂θ
◦
−1 (4.2.8)
∂G ∂g ∂g
≤ · ,
∂x ◦ ∂x ◦ ∂θ
◦
−1
∂G ∂g
= .
∂ θ̄ ◦ ∂θ
◦
One can check that the following inequality follows from these estimates,
and from (4.2.5):
s s
∂F ∂G ∂F ∂G
· + · < 1. (4.2.9)
∂x ◦ ∂ θ̄ ◦ ∂ θ̄ ◦ ∂x ◦
In particular
∂F ∂G
· < 1.
∂x ◦ ∂ θ̄ ◦
244 Chapter 4. Invariant Tori
According to Assumption 4.2, for each fixed x the map θ = G(x, θ̄) is a dif-
feomorphism of the torus Tm onto itself, and hence it cannot be a contraction
mapping. Therefore, the maximum of the norm of its Jacobian matrix is nec-
essarily greater than 1:
∂G
≥ 1.
∂ θ̄ ◦
This implies, in turn, that
∂F
< 1.
∂x ◦
Denote s −1
∂F ∂G
L= (4.2.11)
∂ θ̄ ◦ ∂x ◦
∂G
L< 1, (4.2.12)
∂x ◦
∂F ∂G ∂G 1 ∂F
sup · ≤ 1−L 1− , (4.2.13)
(x,θ̄) ∂x ∂ θ̄ ∂x ◦ L ∂ θ̄ ◦
2
∂F ∂G ∂G
sup · < 1−L , (4.2.14)
(x,θ̄) ∂x ∂ θ̄ ∂x ◦
∂F ∂G
<1−L . (4.2.15)
∂x ◦ ∂x ◦
The rest of the proof is based only upon these inequalities. Let us denote
by H(L) the space of vector-functions x = h(θ) with the graph in K: khk ≤ η 0 ,
where h satisfies a Lipschitz condition:
It is well known that H(L) is closed in the Banach space of bounded continuous
functions h(θ).
Lemma 4.10. Provided (4.2.12) and (4.2.13) are satisfied, the map T induces
the operator T : H(L) → H(L) (i.e. the image of the graph of a Lipschitz
function x = h(θ) by the map T is the graph of a function x̄ = h̃(θ̄) that
satisfies a Lipschitz condition with the same constant L).
Indeed, let h ∈ H(L). We must prove, first, that the image T {x = h(θ)}
is a surface of the kind x̄ = h̃(θ̄) for some single-valued function h̃. In other
words, we must show that for any θ̄ there exists a unique x̄ (which would give
h̃(θ̄)) such that (x̄, θ̄) = T (h(θ), θ) for some θ. This is equivalent (see 4.2.6) to
the existence, for any θ̄, of a unique solution of the following equation on θ:
and Z
1
∂G
G(x + ∆x, θ̄) − G(x, θ̄) = (x + s∆x, θ̄)ds ∆x ,
0 ∂x
whence
∂G
kG(h(θ + ∆θ, θ̄) − G(h(θ), θ̄)k ≤ L k∆θk .
∂x ◦
Thus, for any θ̄ there exists a unique θ for which the equality (4.2.17) holds.
Since the fixed point of a contracting map depends continuously on a parameter
(θ̄ in our case), it follows that the value of θ also depends continuously on θ̄.
By substituting the value of θ into the first equality in (4.2.6) we obtain a
function h̃ = T̃ h in the form
where we denote
Z 1
∂F
Fx = (x + s∆x, θ̄ + s∆θ̄)ds ,
0 ∂x
etc. We assume here that the points (x, θ) and (x + ∆x, θ + ∆θ) belong to
{x = h(θ)} (hence the points (x̄, θ̄) and (x̄+∆x̄, θ̄ +∆θ̄) belong to {x̄ = h̃(θ̄)}).
Therefore, k∆xk ≤ Lk∆θk. Substituting this into (4.2.21) gives
1 kFx k · kGθ̄ k
k∆x̄k ≤ L kF k + k∆θ̄k .
L θ̄ 1 − LkGx k
In the limit we have
∂F ∂G
·
k∆x̄k 1 ∂F ∂x ∂ θ̄
lim sup ≤L + . (4.2.22)
∆θ̄→0 k∆θ̄k L ∂ θ̄ ∂G
1−L
∂x
principle will guarantee the existence of a unique fixed point h∗ for the operator
T̃ on H(L). We would have h˜∗ = h∗ which means, by definition of T̃ , that
the image of the surface {x = h∗ θ} by the map T is the same surface; i.e. this
surface is the sought invariant manifold (to finish the proof we will also need
to establish the smoothness of h∗ ).
Let h1 and h2 be two elements of H(L) and h̃1 , h̃2 are their images by
T̃ . Fix any θ̄ and take the points (x̄1 , θ̄) and (x̄2 , θ̄) at which the surface of
constant θ̄ intersects the surfaces {x̄ = h̃1 (θ̄)} and {x̄ = h̃2 (θ̄)}, respectively.
Since these surfaces are, by definition, the images of the surfaces {x = h 1 (θ)}
and {x = h2 (θ)} by the map T , there exist points (x1 = h1 (θ1 ), θ1 ) and
(x2 = h2 (θ2 ), θ2 ) such that T (x1 , θ1 ) = (x̄1 , θ̄) and T (x2 , θ2 ) = (x̄2 , θ̄). By
(4.2.6)
( (
x̄1 = F (h1 (θ1 ), θ̄) , x̄2 = F (h2 (θ2 ), θ̄) ,
θ1 = G(h1 (θ1 ), θ̄) , θ2 = G(h2 (θ2 ), θ̄) ,
which gives
∂G
kθ1 − θ2 k ≤ kh1 (θ1 ) − h2 (θ2 )k
∂x ◦
(4.2.23)
∂F
kx̄1 − x̄2 k ≤ kh1 (θ1 ) − h2 (θ2 )k .
∂x ◦
∂G
kθ1 − θ2 k ≤ (L kθ1 − θ2 k + dist(h1 , h2 )) ,
∂x ◦
and
∂F
kx̄1 − x̄2 k ≤ (L kθ1 − θ2 k + dist(h1 , h2 )) ,
∂x ◦
or
−1
∂G ∂G
kθ1 − θ2 k ≤ 1− L · dist(h1 , h2 )
∂x ◦ ∂x ◦
248 Chapter 4. Invariant Tori
and, finally,
∂F
∂x ◦
h̃1 (θ̄) − h̃2 (θ̄) ≡ kx̄1 − x̄2 k ≤ · dist(h1 , h2 ) .
∂G
1− L
∂x ◦
The last equation defines the map T −1 on the invariant manifold. The same
arguments as in Lemma 4.10 shows that θ is a well-defined single-valued con-
tinuous function of θ̄.
It follows from a formal differentiation of (4.2.24) and (4.2.25) that the
∗
derivative η ∗ = dh
dθ (if it exists) must satisfy the equation
−1
∗ ∂F ∂F ∗ ∂G ∗ ∂G
η (θ̄) = + · η (θ) · I − · η (θ) · , (4.2.26)
∂ θ̄ ∂x ∂x ∂ θ̄
4.2. Existence of an invariant torus. The annulus principle 249
where all derivatives on the right-hand side are computed at (x = h∗ (θ), θ̄)
and θ is defined by (4.2.25) as a function of θ̄. Let us prove that a continuous
function η ∗ which satisfies this equality exists. Consider the space H 0 (L) of
bounded (kηk◦ ≤ L) continuous functions x = η(θ). It is a closed subset of a
Banach space of continuous functions with the norm
kη1 − η2 k = dist(η1 , η2 ) = sup kη1 (θ) − η2 (θ)k .
θ
0
Consider the map η 7→ η̃ defined on H (L):
−1
∂F ∂F ∂G ∂G
η̃(θ̄) = + · η(θ) · I − · η(θ) · . (4.2.27)
∂ θ̄ ∂x ∂x ∂ θ̄
This formula gives a rule for calculating η̃ when the function η is given: for
an arbitrary θ̄ find θ by formula (4.2.25) and substitute the result into the
right-hand side of (4.2.27).
We will prove that the map given by (4.2.27) takes H 0 (L) into itself and that
it is contracting — this implies the existence and uniqueness of the solution
η ∗ of (4.2.26). The continuity of η̃ is obvious so we only need to check that it
is bounded by L provided η is bounded by the same constant. Since
∂G ∂G
·η ≤ L<1
∂x ∂x ◦
whence
−1 +∞ k +∞ k
∂G X ∂G X ∂G
I− ·η ≤ ·η ≤ L
∂x ∂x ∂x ◦
k=0 k=0
1
= .
∂G
1−L
∂x ◦
kη̃k ≤ L
Theorem 4.3. Let X and Y be some convex closed subsets of some Banach
spaces. Suppose a map T is defined in the cross-form on X × Y :
x̄ = F (x, ȳ) ,
(4.2.35)
y = G(x, ȳ) ,
which means that two points (x, y) and (x̄, ȳ) from X × Y are related by the
map T if and only if (4.2.35) holds. Let F and G be smooth functions satisfying
the following two conditions
s s
∂F ∂G ∂F ∂G
sup · + <1 (4.2.36)
(x,ȳ)∈X×Y ∂x ∂ θ̄ ∂ θ̄ ◦ ∂x ◦
and s
∂F ∂F ∂G
+ < 1. (4.2.37)
∂x ◦ ∂ θ̄ ◦ ∂x ◦
Then the map T has an invariant C1 -smooth manifold M ∗ which contains the
ω-limit points of any forward orbit of T.
4.2. Existence of an invariant torus. The annulus principle 253
(simply rewrite formulas (4.2.24) and (4.2.26)). We see that the derivative η ∗
is a function whose graph is an invariant manifold of the map written in the
cross-form:
η̄ = F(η, ȳ) ,
(4.2.40)
y = G(ȳ) ,
∂F ∂G
·
∂F ∂x ∂ ȳ
≤ 2 .
∂η ∂G
1−L
∂x ◦
Here L is the Lipschitz constant — the upper bound for the norm of the
derivative η ∗ . By construction (see (4.2.11)),
s −1
∂F ∂G
L= .
∂ θ̄ ◦ ∂x ◦
Thus,
∂F ∂G
·
∂F ∂x ∂ ȳ
≤ 2 . (4.2.43)
∂η s
1 − ∂G ∂F
∂x ◦ ∂ θ̄ ◦
∂G
For the derivative ∂ ȳ we obtain the following estimate directly from (4.2.38):
∂G
∂G ∂ ȳ
≤ 2 . (4.2.44)
∂ ȳ s
1 − ∂G ∂F
∂x ◦ ∂ θ̄ ◦
Substituting these two inequalities into (4.2.41) we obtain the following addi-
tional sufficient condition for C2 -smoothness of M ∗ (condition (4.2.42) does not
introduce new restrictions in comparison with the conditions of Theorem 4.3):
( )
2
∂F ∂G
sup ·
∂x ∂ ȳ
s 3 < 1 .
1 − ∂G ∂F
∂x ◦ ∂ θ̄ ◦
4.2. Existence of an invariant torus. The annulus principle 255
or v
u ( ) s
2
u
3 ∂F ∂G ∂G ∂F
t sup · + < 1. (4.2.45)
∂x ∂ ȳ ∂x ◦ ∂ ȳ ◦
One may repeat the above procedure to derive sufficient conditions for
C3 -smoothness (by plugging (4.2.43) and (4.2.44) into (4.2.45)), etc. By in-
duction, we arrive at the following theorem.
Returning to the annulus principle, Theorem 4.4 gives the following result
(see estimates (4.2.7)–(4.2.8) relating the derivatives of the cross-map and of
the initial map).
256 Chapter 4. Invariant Tori
x̄ = f (x, θ) ,
(4.2.50)
θ̄ = Aθ + g0 (x, θ) = g(x, θ) (mod 1) ,
where f and g are periodic functions of period 1 with respect to θ, the annulus
principle is valid if in the ring K the following conditions are satisfied:
(3) The map (4.2.50) satisfies conditions (4.2.5) or, for more smoothness,
condition (4.2.47).
258 Chapter 4. Invariant Tori
The proof of this statement is an exact copy of the proof of Theorem 4.2
or 4.5.
The map (4.2.50) when restricted on the torus Tm : x = h(θ) can be written
in the form
θ̄ = Aθ + g0 (h(θ), θ) (mod 1) . (4.2.51)
ẋ = X(x) + p(x, θ, µ) ,
(4.3.1)
θ̇ = 1 ,
Theorem 4.6. Under the above assumptions system (4.3.1) has a Cr -smooth
two-dimensional invariant torus for all sufficiently small µ.
Proof. Let us introduce the normal coordinates (y, θ0 ) (see (3.11.20)) in-
stead of the x-coordinate in a small neighborhood of L. In the new variables
the family takes the form
ẏ = Λy + F0 (θ0 , y) + F1 (θ0 , y, θ, µ) ,
θ̇0 = Ω0 + b0 (θ0 , y) + b1 (θ0 , y, θ, µ) , (4.3.3)
θ̇ = 1 .
4.3. Theorem on persistence of an invariant torus 259
where
0
F0 (θ0 , 0) = 0, F0y (θ0 , 0) = 0, b0 (θ0 , 0) = 0 , (4.3.4)
and Ω0 = 2π/τ .
The functions in the right-hand side of (4.3.3) are periodic of period 2π
with respect to θ, and either of period 2π with respect to θ0 or (see (3.11.22))
they are antiperiodic:2
Let us now verify the conditions of the annulus principle (see the previous
section). For a moment, we will not consider θ0 as an angular variable but we
assume that θ0 ∈ (−∞, +∞); obviously, the conclusion of Theorem 4.2 on the
existence of an invariant curve y = h(θ0 ) will not change.
First, we must find δ such that the strip kyk ≤ δ is mapped into itself. Note
0
that by (4.3.6), kf0y k is small within such a strip for any sufficiently small δ.
Thus, at µ = 0, we have from (4.3.6), (4.3.7) that
and the fulfillment of the conditions of Theorems 4.2 and 4.5 follows
immediately.
Thus, we established the existence, for all sufficiently small µ, a unique
attractive invariant Cr -smooth curve y = h(θ0 , µ). By now θ0 ∈ (−∞, +∞).
Since the right-hand side of (4.3.6) is (anti)periodic, it follows that y = σh(θ 0 +
2π, µ) is also an invariant curve of this map. By uniqueness, we get
Recall, that by construction, the points (y, θ0 ) and (σy, θ0 + 2π) must be iden-
tified because they correspond to the same point in the original x-coordinates.
Thus, relation (4.3.8) shows that the invariant curve y = h(θ0 , µ) is homeo-
morphic to a circle.
We have found a stable invariant circle for the time 2π map of the cross-
section θ = 0 (mod 2π) of system (4.3.1). The union of the trajectories
starting on this circle is a two-dimensional stable invariant torus. End of the
proof.
Remark. It is easy to check that our proof is applied, without changes, to
the case where the function p(x, θ, µ) in (4.3.1) depends only continuously on
θ. In this case, the invariant torus is Cr -smooth in the intersection with any
cross-section θ = const.
In the same way one may consider the general case where the autonomous
system (4.3.2) has an arbitrary structurally stable periodic orbit, with m mul-
tipliers inside the unit circle and n multipliers outside the unit circle. The
4.3. Theorem on persistence of an invariant torus 261
where the spectrum of Λs lies strictly to the left of the imaginary axis and the
spectrum of Λu lies strictly to the right of it. The time 2π map {θ = 0} →
{θ = 2π} is written, at µ = 0, as
s
ū = e2πΛ u + o(u, v) ,
u
v̄ = e2πΛ v + o(u, v) ,
θ̄0 = θ0 + ω0 + O(u, v) .
At small u and v, it is easy to see that the conditions of Theorem 4.4 are
fulfilled for this map (one should consider x = u and y = (v, θ0 ) in (4.2.46)) and
for the inverse to this map (in this case one should put x = v and y = (u, θ 0 ) in
(4.2.46)). By continuity, this holds true for all small µ. Thus, we established
the existence of two smooth invariant manifolds at all small µ: a manifold
Mµu : u = hu (v, θ0 , µ) which attracts all forward iterations of the map, and a
repelling manifold Mµs : v = hs (u, θ0 , µ) which attracts all backward iterations.
The trajectories which stay in a small neighborhood of L for all forward and
backward iterations of the map, belong to the invariant circle Lµ = Mµu ∩ Mµs .
By construction, the ω-limit set of any point in Mµs and the α-limit set of any
point in Mµu belongs to Lµ . Returning from the map on the cross-section to
the original system we arrive at the following result.
ẋ = X(x) (4.3.10)
has a structurally stable equilibrium state O. Near the point O, the Poincaré
map of the cross-section θk = 0 mod 2π into itself is written as
x̄ = e2πA x + o(x) + . . . ,
θ̄j = θj + ωi (mod 2π) (j = 1, . . . , k − 1)
Ω
where ωj = 2π Ωkj ; the matrix A is the linearization matrix of (4.3.10) at O;
the ellipsis denotes the terms which vanish at µ = 0.
Applying the annulus principle we can prove that for all sufficiently small µ
the system (4.3.9) has a k-dimensional invariant torus Tk close to x = 0. Obvi-
ously, the stability of the torus is determined by the stability of the equilibrium
state with respect to the autonomous system (4.3.10).
The torus has the form x = h(θ, µ) (where h = 0 at µ = 0). Hence, the
motion on the torus is described by the second equation in (4.3.9) alone and
is represented as a quasi-periodic motion with the frequency basis Ω.
3 When the right-hand sides are smooth with respect to µ, the invariant curve depends
Let us now assume that the system (4.3.10) has a structurally stable peri-
2π
odic orbit L of period Ω 0
. In this case, at µ = 0 the system (4.3.9) possesses a
(k + 1)-dimensional invariant torus Tk+1 0 = L × Tk . This torus is a minimal set
if the frequencies Ω0 , Ω1 , . . . , Ωk form a basis. Otherwise, the torus is foliated
into a family of k-dimensional tori.
In a similar manner to Theorem 4.6, we can construct the Poincaré map
of the cross-section θk = 0 (mod 2π) into itself along the trajectories near
the torus Tk+1
0 . Applying the annulus principle we can prove that this map
possesses a Cr -smooth invariant torus Tkµ for all sufficiently small µ. Hence,
the system (4.3.9) possesses a (k + 1)-dimensional Cr -smooth invariant torus
Tk+1
µ .
The map on Tkµ has the form
ẋ = X(x) + p(x, Ω1 t, . . . , Ωk t, µ) ,
the same basis of frequencies, for all small µ. A stable periodic orbit L of
(4.3.10) corresponds to a stable quasi-periodic “tube” in Rn × Rk which is
homeomorphic4 to an infinite (k + 1)-dimensional cylinder.
If the function p is represented as
where g(θ) is a periodic function of θ with period 2π. Equation (4.4.1) may be
rewritten in the form
{θj }∞
j=0 be a positive semi-trajectory of an initial point θ0 . Poincaré showed
that there exists
θj
ω = lim ,
j→∞ 2πj
and that this limit ω does not depend on the choice of the initial point θ 0 . The
value ω is called the Poincaré rotation number.
Theorem 4.8. (Poincaré) If the rotation number ω is rational, then the set
of non-wandering points consists of periodic points, all having the same period.
If ω is irrational, then the non-wandering set contains no periodic points.
It follows from Denjoy’s theorem that in this case the entire circle is a
minimal set. When (4.4.1) is only C1 -smooth, Denjoy constructed examples
where the non-wandering set is a minimal set with a structure analogous to a
Cantor discontinuum. This is the reason why we have given a special attention
earlier to the necessity of proving, at least, the C2 -smoothness of the invariant
curves.
Let us consider next a one-parameter family of diffeomorphisms
Poincaré, and later, Krylov and Bogolyubov, certainly knew of this result,
the proof of which was given in an explicit form by Maier. In this connection
we must note the following result obtained by Hermann [34]: if the family
depends smoothly on µ and if
for all θ ∈ S1 and for µ from some interval ∆, then ω(µ) is a strictly monotonic
function of each µ ∈ ∆ at which ω(µ) is irrational.
Let us denote by B the space of all diffeomorphisms of the form (4.4.5) and
let us introduce the distance between any two such diffeomorphisms as follows.
Let T1 be
θ̄ = θ + g1 (θ, µ) (mod 2π) ,
and let T2 be
θ̄ = θ + g2 (θ, µ) (mod 2π) .
Then n o
dist(T1 , T2 ) = max |g1 (θ) − g2 (θ)| + |g10 (θ) − g20 (θ)| .
θ
where 0 < µ 1. A careful analysis reveals that only resonances of the kind
(1 : 1); (1 : 2); (1 : 3); (2 : 3) are easily observed. Using an averaging method,
it can be shown that the other synchronization intervals have a size of order
e−1/µ . The ratios of observable resonances in strongly nonlinear systems may
be different. Numerous experiments in the study of the problem on the onset
of turbulence have confirmed this fact.
So, only a finite number of visible synchronization regions may be detected
(we do not discuss the clearance of experimental observations). The rest of the
parameter regions where a two-dimensional invariant torus exists is usually
interpreted, or associated with regions of modulation and beatings. Using the
language of the theory of dynamical systems the modulation regimes can be
translated either as a stable torus with an aperiodic trajectory on it, or with
a stable periodic trajectory of a rather large period.6 It is not superfluous to
recall that systems with aperiodic trajectory behavior on a torus, in general,
do not form a region in the parameter space.
Here we run into the case where the mathematical interpretation of the
problem of synchronization differs in essence from that which is broadly used
in nonlinear dynamics. The reason is that our traditional qualitative analysis
employed the notion of irrational numbers which is a purely mathematical
abstraction. To summarize we remark that the above observation is not the
only example where the mathematical formalization of a problem based on the
notion of irrational numbers does not agree with that suggested by common
sense which lies beneath any empirical means or computer experiments. 7
multi-frequent regime.
7 A similar situation arises when studying numerically an autonomous system under the
action of a quasi-periodic external force, where the notion of a basis of independent frequen-
cies plays a primary role.
Chapter 5
269
270 Chapter 5. Center Manifold. Local Case
∂X(x(µ), µ)
A(µ) = .
∂x
Let us arbitrarily choose µ1 which satisfies the condition |µ1 − µ0 | < δ0 .
Repeating the above reasoning, we can find a new neighborhood |µ − µ1 | < δ1
where system (5.0.1) will have a stable equilibrium state Oµ , and so forth. As
a result we can construct a maximal open set G in the parameter space, which
is called the stability region of Oµ . This procedure for constructing the stabili-
ty region resembles the construction of a Riemannian surface of an analytical
function by means of Weierstrass’s method. It may turn out that the stability
region possesses a branched structure.
The boundary Γ of the stability region G corresponds to the case where
some characteristic exponents of the equilibrium state Oµ , say λ1 , . . . , λm , lie
on the imaginary axis, whereas the rest of the eigenvalues λm+1 , . . . , λn will
still reside in the open left-half plane. Thus, near the bifurcating equilibrium
state for some fixed parameter value on the boundary Γ, the system takes the
form
ẏ = Ay + f (x, y) ,
(5.0.2)
ẋ = Bx + g(x, y) ,
The modern approach for studying critical cases is restricted not only to
the problem of stability. It also includes finding out what causes an equilibrium
state to lose its stability and what happens beyond the stability boundary Γ.
To answer these questions, the system under consideration must depend on
some parameter µ:
ẏ = Ay + f (x, y, µ) ,
(5.0.3)
ẋ = Bx + g(x, y, µ) ,
where µ takes the values near some critical parameter value µ∗ (below we will
assume that µ∗ = 0). All of the above problems constitute the main issue of
the theory of local bifurcations. The basic results of this theory is the center
manifold theorem credited to Pliss [52] and Kelley [37].
The existence of a center manifold allows the problems related to the critical
cases to be reduced to the study of an m-dimensional system
ẏ = (A + F (x, y, µ)) y ,
(5.0.6)
ẋ = Bx + G(x, µ) ,
where G(x, µ) ≡ g(x, ψ(x, µ), µ) and F ∈ Cr−1 , F (0, 0, 0) = 0. This means
that the behavior of the “critical” variables x in a small neighborhood of the
structurally unstable equilibrium is independent of the other variables and
repeats the behavior on the center manifold. For the y-variables we have an
exponential contraction (because the spectrum of A lies strictly to the left of
the imaginary axis. Compare with Sec. 2.6).
An analogue of the center manifold theorem holds true in the general case,
i.e. when we consider an equilibrium state which has some characteristic ex-
ponents to the right of the imaginary axis as well. Therefore, in this case the
qualitative study of local bifurcations can also be reduced to a lower dimen-
sional system. Note, however, that the smooth reduction of the entire system
into a triangular form of type (5.0.6) is not always possible in this general case
(the corresponding coordinate transformation is only C0 ).
The same scheme works in studying the behavior of the solutions on the
boundary of the existence of periodic trajectories but with one significant re-
striction. For the periodic trajectories, in contrast to the case of equilibrium
states, the stability or existence boundaries may be of two different types,
namely:
The boundaries of the second type do not exist in the case of equilibrium
states, whereas it is well known that the periodic trajectories can disappear
upon approaching a bifurcation boundary: it is accomplished by collapsing into
an equilibrium state, or by merging into a homoclinic loop, or via a more com-
plicated structure — through “a blue sky catastrophe”. We will not consider
boundaries of the second type in Part I of this book.
Once we restrict ourself to the first case, we can construct a cross-section
through the critical periodic trajectory (which now exists by assumption) and
proceed with the study of the behavior of the trajectories of an associated
Poincaré map close to the bifurcating fixed point. After that the center mani-
fold theory can be applied just as in the case of equilibrium states.
The proof of the center manifold theorem which we present in this chapter
is based on the study of some boundary-value problems (Secs. 5.2 and 5.3)
as in Sec. 2.8 for the proof of the existence and smoothness of the stable and
unstable manifolds of a saddle equilibrium state. We will develop a unified
approach for both equilibrium states and periodic trajectories. Moreover, our
proof will include all other local invariant manifold theorems throughout this
book, and the theorems on invariant foliations as well.
We note that besides dynamical applications of the invariant manifold the-
orems, these results may also be used indirectly; for example in reducing a
system near a saddle point to a special form. To do this we must select the
strongly stable and unstable manifolds of the saddle. This topic is discussed
in detail in Appendix A.
ẏ = Ay + f (x, y) ,
(5.1.1)
ẋ = Bx + g(x, y) ,
274 Chapter 5. Center Manifold. Local Case
ẏ = Ay + f (x, y, µ) ,
(5.1.2)
ẋ = Bx + g(x, y, µ) .
The proof of this theorem is given in Sec. 5.4. Note that in the case where
the right-hand side of system (5.1.2) depends smoothly on µ, the center mani-
fold depends smoothly on µ as well. In particular, if the functions f and g are
Cr with respect to (x, y, µ), then the function ψ (whose graph y = ψ(x, µ) is
C
Wloc ) may be taken to be Cr with respect to (x, µ). This smoothness result
follows from Theorem 5.2 if we add, formally, an equation
µ̇ = 0
to system (5.1.2). If we now consider the pair (x, µ) as a new variable x, then
the form of the augmented system is analogous to the parameterless system
(5.1.1). This mean that one may apply the center manifold theorem which
gives now a center manifold depending Cr -smoothly on the new x; i.e. it is Cr
with respect to (x, µ). This trick often allows one to eliminate any dependence
of the system on µ. Consequently, we will omit the dependence on µ where it
is not essential.
What should also be mentioned concerning the smoothness of W C is that
even if the system is C∞ , the center manifold is not necessarily C∞ . Of course,
if the original system is C∞ , it is Cr for any finite r. Therefore, in this case
one may apply the center manifold theorem with any given r which implies
5.1. Reduction to the center manifold 275
that:
If the original system is C∞ , then for any finite r there exists a neigh-
C
borhood Ur of the origin where Wloc is Cr .
There is an ambiguity here, caused by the fact that the center manifold is
not uniquely defined by the system. Therefore our notion of the reduction of a
system onto a center manifold requires some logical analysis.
276 Chapter 5. Center Manifold. Local Case
In other words, if for some small x the trajectory of the point (x, ψ1 (x)) does
not leave a small neighborhood of O for all times, then ψ2 (x) = ψ1 (x); i.e. the
function ψ is uniquely defined for all x corresponding to the points of N . In
fact, the following, more general, statement holds.
Theorem 5.3. For any two center manifolds y = ψ1 (x) and y = ψ2 (x), at
each x0 such that (x0 , y0 ) ∈ N for some y0 , the function ψ1 coincides with ψ2
along with all of the derivatives:
dk ψ1 dk ψ2
= , k = 0, . . . , r .
dxk x=x0 dxk x=x0
Theorem 5.4. (On smooth conjugacy) For any two local center manifolds
W C1 and W C2 there exists a Cr−1 change of variables x which maps trajecto-
ries of the reduced system
ẋ = Bx + g(x, ψ1 (x))
ẋ = Bx + g(x, ψ2 (x)) .
2 Unlike the case of a stable equilibrium where N = {O}, the presence of a zero, or pure
imaginary characteristic exponents may make the structure of the set N quite non-trivial.
5.1. Reduction to the center manifold 277
ẏ = (A + F (x, y))y ,
(5.1.6)
ẋ = Bx + G(x) ,
where F ∈ Cr−1 , G ∈ Cr
to this particular form is Cr−1 , we have a Cr−1 -conjugacy when the system is
not reduced to this form.
Let us give a geometrical interpretation of Theorem 5.5. Obviously, when
the system is in the triangular form (5.1.6), the time-t shift of any surface
{x = const} lies again in a surface of the same kind, for any t (unless the
trajectories leave a small neighborhood of O). This means that the foliation of
a small neighborhood of O by surfaces of constant x is invariant with respect
to the system (5.1.6). The coordinate change which transforms system (5.1.6)
to the initial form (5.1.1) maps the surfaces {x = constant} into surfaces of
the kind
x = ξ + η(y, ξ) , (5.1.7)
where ξ is the x-coordinate of the intersection of the surface with a center man-
ifold; the Cr−1 -function η vanishes at the origin along with the first derivatives
(note that η ≡ 0 everywhere on W C ).
Since the transformation which maps the surfaces {x = constant} into the
surfaces given by (5.1.7) is a diffeomorphism, it follows that Eq. (5.1.7) defines
a foliation of a small neighborhood of the origin by surfaces corresponding to
fixed ξ. This implies that for each point (x, y) there is a unique ξ such that the
surface corresponding to the given ξ passes through the point (x, y). Such a
surface is called a leaf of the foliation: for each point from a small neighborhood
of O there is one and only one leaf which contains the point. Since the leaves
C
are parametrized by the points on Wloc , the center manifold is the base of the
foliation. Since the foliation {x = constant} is invariant with respect to system
(5.1.6), its image (i.e. the foliation given by (5.1.7)) is an invariant foliation
of system (5.1.1): for any t, the time-t shift of any leaf lies in a single leaf
of the same foliation unless the trajectories leave a small neighborhood of O.
After straightening the center manifold, the reduction to the triangular form
(5.1.6) is achieved by just transforming x 7→ ξ(x, y) (inverse of (5.1.7)): the
variable x is replaced by the x-coordinate of the projection of the point along
the leaves of the invariant foliation onto the center manifold. The invariance
of the foliation simply means that the evolution of the new coordinate x = ξ
is independent of y. Thus, we see that Theorem 5.5 basically establishes the
existence of a foliation of the kind (5.1.7), transverse to the center manifold
and invariant with respect to system (5.1.1). We will call it the strong stable
foliation and denote it by F ss . We will prove that the foliation is uniquely
defined at all points whose trajectories stay in a small neighborhood of O for
280 Chapter 5. Center Manifold. Local Case
purely imaginary, and the eigenvalues of the matrix C lie to the right of the
imaginary axis; the Cr -functions f , h and g vanish at the origin along with
their first derivatives. The right-hand sides of the system may depend on some
parameters µ, either continuously (in this case the smooth manifolds to be
discussed below depend continuously on µ), or smoothly. In the latter case we
will include µ among the “center” variables x so the manifolds and foliations,
which we discuss below, will have the same smoothness with respect to µ as
the smoothness with respect to x.
The center manifold theory is based here on the following theorem.
The proof will be given in Sec. 5.4. Note that if the system is C∞ -smooth,
the center stable manifold has, in general, only finite smoothness: for any finite
r there exists a neighborhood Ur of O where W sC is Cr -smooth. Just like the
reasoning above, we can conclude that
If the system is C∞ , and if every trajectory of Wloc
sC
tends to O as
sC ∞
t → +∞, then Wloc is C -smooth.
The reversion of time t → −t changes matrices A, B and C to −A, −B and
−C, respectively. Thus, the part of the spectrum of characteristic exponents
that corresponds to the z-variables is now to the left of the imaginary axis,
and the part of the spectrum that corresponds to the y-variables is now to the
right. We may apply the theorem on the center stable manifold to the system
obtained from (5.1.8) by a reversion of time and obtain the following theorem
on a center unstable manifold:
neighborhood of O for all negative times. For any two manifolds W1uC and W2uC
the functions ψ1uC and ψ2uC have the same Taylor expansion at O (and at each
point whose trajectory stays in a small neighborhood of O for all t ≤ 0). In
the case where the system is C∞ -smooth, the center unstable manifold has, in
uC
general, only finite smoothness, but if every trajectory of Wloc tends to O as
uC ∞
t → −∞, then Wloc is C -smooth.
Here, the local center unstable manifold is given by {y = 0}, the local
center stable manifold is given by {z = 0}, and the local center manifold
is given by {y = 0, z = 0}. The strong stable foliation is composed of the
surfaces {x = constant, z = 0} and the leaves of the strong unstable foliation
are {x = constant, y = 0}.
An analogous theory may be applied to the study of structurally unstable
periodic trajectories. The study of the dynamics in a small neighborhood of
a periodic trajectory is reduced to the study of the Poincaré map on a small
cross-section; the point O of intersection of the trajectory with the cross-section
is a fixed point of the Poincaré map.
Let the system be (n+1)-dimensional, so the cross-section is n-dimensional.
Let m multipliers of the periodic trajectory lie on the unit circle, k multipliers
lie strictly inside the unit circle and the other (n−m−k) multipliers are strictly
greater than 1 in absolute value. The Poincaré map near the fixed point O is
written in the following form:
ȳ = Ay + f (x, y, z) ,
z̄ = Cz + h(x, y, z) , (5.1.11)
x̄ = Bx + g(x, y, z) ,
set N + of all points whose forward iterations by the map (5.1.11) stay in a
uC
small neighborhood of O, and Wloc contains the set N − of all points whose
backward iterations never leave a small neighborhood of O. The intersection
sC uC
of Wloc and Wloc is a Cr -smooth invariant m-dimensional center manifold
C C
Wloc : (y, z) = ψ (x) which is tangent at O to the x-space and which contains
the set N = N + ∩ N − composed of all points whose iterations (both forward
and backward) never leave a small neighborhood of O. The Taylor expansions
of the functions ψ uC , ψ sC and ψ C at the origin (and at each point of the sets
N − , N + or N, respectively) are uniquely defined by the system. On the mani-
sC uC
folds Wloc and Wloc there exist, respectively, strong stable and strong unstable
C -smooth invariant foliations F ss and F uu with Cr -smooth k-dimensional
r−1
C
(resp. (n−m−k)-dimensional) leaves transverse to Wloc . The leaves of F ss are
uniquely defined at each point of the set N and the leaves of F uu are uniquely
+
defined at each point of N − . By projection along the leaves of the strong stable
and strong unstable invariant foliations, the restrictions of the same map on
different center manifolds are Cr−1 -conjugate.
The proof will be given in Sec. 5.4. We remark again that if even in the
case where the system under consideration is C∞ , the invariant manifolds are,
in general, of a finite smoothness only. Nevertheless,
ȳ = (A + F (x, y, z))y ,
z̄ = (C + H(x, y, z))z , (5.1.12)
x̄ = Bx + G0 (x) + G1 (x, y, z)y + G2 (x, y, z)z ,
286 Chapter 5. Center Manifold. Local Case
In this section we begin our proof of the center manifold theorems. The
method, which we will use, is based on a generalization of the boundary-value
problem which we have considered in Chap. 2 (see Shashkov and Turaev [52]).
Since the results will be applied to the proof of various invariant manifold
theorems beyond the center manifold theory, we will try to make the setting
sufficiently general.
Let us consider a system of differential equations
ż = Az + f (z, v, µ, t) ,
(5.2.1)
v̇ = Bv + g(z, v, µ, t) ,
where z ∈ Rn , v ∈ Rm , t is the time variable and µ is a vector of parameters.
We assume that f and g are Cr -smooth (r ≥ 1) with respect to the variables
(z, v) and that they depend continuously on (µ, t) along with all the derivatives
(a particular case of interest is when f and g are Cr -smooth with respect to all
of their arguments (z, v, µ, t)). Concerning the matrices A and B, we assume
that the following conditions hold
spectr A = {α1 , . . . , αn } , spectr B = {β1 , . . . , βm } ,
(5.2.2)
max Re αi < α < β < min Re βj ;
i=1,...,n j=1,...,m
5.2. A boundary-value problem 287
i.e. there is a strip in the complex plane (the strip α ≤ Re (·) ≤ β) that
separates the spectra of A and B. It follows from (5.2.2) that in the appropriate
(Jordan) basis the following estimates hold:
eAs ≤ eαs ,
(5.2.3)
e−Bs ≤ e−βs
∂(f, g)
<ξ (5.2.4)
∂(z, v)
for some sufficiently small constant ξ (the exact value of ξ can be obtained
from the proofs of the theorems below). We will also assume that all of the
derivatives of f and g are bounded uniformly for all z and v. The last conditions
mean that the nonlinear part essentially does not affect the behavior induced
by the specific structure of the linear part of the system (the separation of the
spectrum). To stress this property we will call the systems satisfying (5.2.1)–
(5.2.4) globally dichotomic.
Such systems appear naturally in the study of equilibrium states and peri-
odic trajectories. For example, if the spectrum of the characteristic exponents
of an equilibrium state is separated so that n characteristic exponents lie to the
left of the line Re (·) = α and the other m characteristic exponents lie to the
right of the line Re (·) = β in the complex plane, then near such an equilibrium
the system may be written locally in the form
ż = Az + f (z, v, µ) ,
(5.2.5)
v̇ = Bv + g(z, v, µ) ,
∂(f, g)
f (0, 0, 0) = 0 , g(0, 0, 0) = 0 , = 0. (5.2.6)
∂(z, v) (z,v,µ)=0
Of course, the last equality in (5.2.6) implies the fulfillment of (5.2.4) for an
arbitrarily small ξ, in a sufficiently small neighborhood of O.
288 Chapter 5. Center Manifold. Local Case
The difference between systems (5.2.1) and (5.2.5) is that the nonlinear part
of the latter remains small only near the origin, whereas for system (5.2.1) the
linear part prevails everywhere in Rn+m . A very useful trick which allows
one to proceed from the local system (5.2.5) to the global version (5.2.1) is as
follows. Consider a new system
ż = Az + f˜(z, v, µ)
(5.2.7)
v̇ = Bv + g̃(z, v, µ) ,
ż = Az + f (z, v, µ, t)
(5.2.10)
v̇ = Bv + g(z, v, µ, t) ,
4 Note that the solution of boundary-value problem (5.2.11) for system (5.2.1) is not
prohibited from leaving a small neighborhood of zero. Therefore, the theorem cannot be
directly applied to the local systems (5.2.5), or (5.2.10); namely, additional estimates are
necessary to guarantee that the solution of the boundary-value problem stays bounded by a
small constant.
290 Chapter 5. Center Manifold. Local Case
where
α<γ <β. (5.2.14)
Obviously, H is a complete metric space.
Let us consider the integral operator T : H → H, which maps a function
(z(t), v(t)) onto the function (z(t), v(t)) via the following formula:
Z t
At 0
z(t) = e z + eA(t−s) f (z(s), v(s), µ, s) ds ,
0
Z τ (5.2.15)
−B(τ −t) 1
v(t) = e v − e−B(s−t) g(z(s), v(s), µ, s) ds .
t
converges to a uniquely defined fixed point (z ∗ (t), v ∗ (t)) of T , which is, at the
same time, the solution of boundary-value problem (5.2.11) for system (5.2.1).
This completes the proof.
Note that the solution (z ∗ (t), v ∗ (t)) depends also on {z 0 , v 1 , τ, µ}. Since
the integral operator T given by (5.2.15) is continuous with respect to these
data, the solution depends continuously on {z 0 , v 1 , τ, µ} (as the fixed point of
a contraction operator, see Theorem 3.25). Moreover, by Theorem 3.27, since
the operator T is Cr -smooth and depends Cr -smoothly on {z 0 , v 1 }, it follows
that the solution of the boundary value problem is a Cr -smooth function of
{z 0 , v 1 }, and if the right-hand sides of the system are Cr with respect to all
variables including t and µ, then the solution depends smoothly on {t, τ, µ} as
well.
The derivatives of (z ∗ , v ∗ ) with respect to z 0 , v 1 and µ are found as fixed
points of an operator obtained by formal differentiation of (5.2.15); i.e. they are
found as the solutions of the boundary-value problems for the corresponding
variational equations with the boundary conditions obtained by formal differ-
entiation of the boundary conditions (5.2.11) (see Sec. 2.8 for more details).
For example, the solution of the boundary-value problem
Z(0) = I , V (τ ) = 0 , (5.2.19)
(I is the identity matrix) for the system of variational equations
Ż = AZ + fz0 (z ∗ , v ∗ , µ, t)Z + fv0 (z ∗ , v ∗ , µ, t)V ,
(5.2.20)
V̇ = BV + gz0 (z ∗ , v ∗ , µ, t)Z + gv0 (z ∗ , v ∗ , µ, t)V ,
gives the derivative of (z ∗ , v ∗ ) with respect to z 0 :
∂z ∗ ∂v ∗
Z∗ = , V∗ = .
∂z 0 ∂z 0
292 Chapter 5. Center Manifold. Local Case
Z(0) = 0 , V (τ ) = I (5.2.21)
for the same system (5.2.20) gives the derivative of (z ∗ , v ∗ ) with respect to v 1 .
If f and g are smooth with respect to the parameter µ, then the derivatives
(Z ∗ , V ∗ ) = (∂z ∗ /∂µ, ∂v ∗ /∂µ) are the solution of the boundary-value problem
Z(0) = 0 , V (τ ) = 0 (5.2.22)
One can immediately see that system (5.2.20) or (5.2.23) is globally dichotomic:
since the principal part of the right-hand sides is determined by the same
matrices A and B, it follows that the condition (5.2.2) on the separation of the
spectra still holds here; the residual part of the right-hand side is
In both cases
∂(F, G) ∂(f, g)
= ,
∂(Z, V ) ∂(z, v)
so condition (5.2.4) on the smallness of the derivatives is also fulfilled with the
same ξ. Thus, the existence (and uniqueness) of solutions of the boundary-
value problems (5.2.19), (5.2.21) and (5.2.22) simply follows from Theorem 5.11.
Now, by induction, one can see that the higher-order variational equations
also belong to our class of globally dichotomic systems. Therefore, for the
corresponding boundary-value problems, the existence and uniqueness of the
solutions (which are the higher-order derivatives of (z ∗ , v ∗ )) is also given by
Theorem 5.11.
5.2. A boundary-value problem 293
∂z ∗
≡ ż ∗ = Az ∗ + f (z ∗ , v ∗ , µ, t)
∂t
and
∂v ∗
≡ v̇ ∗ = Bz ∗ + g(z ∗ , v ∗ , µ, t) .
∂t
The higher-order derivatives involving time are obtained by a repeated use of
these identities. The derivatives with respect to τ may be calculated using the
following lemma.
∂(z ∗ , v ∗ ) ∂v ∗ ∂(z ∗ , v ∗ )
+ ≡ 0, (5.2.24)
∂v 1 ∂t t=τ ∂τ
implies (5.2.25).
294 Chapter 5. Center Manifold. Local Case
Our next theorem gives estimates on the derivatives of the solution of the
boundary-value problem. We use the following notation for the derivatives of
a vector-function φ = (φ1 , . . . , φq ) ∈ Rq with respect to a vector-argument
x = (x1 , . . . , xp ) ∈ Rp :
s1 +···+sp
∂ |s| φ ∂ φ1 ∂ s1 +···+sp φq
≡ s , . . . , s
∂xs ∂xs11 · · · ∂xpp ∂xs11 · · · ∂xpp
where the multi-index s = (s1 , . . . , sp ) consists of non-negative integers and |s|
denotes s1 + · · · + sp .
Theorem 5.12. The following estimates hold for the solution (z ∗ , v ∗ ) of the
boundary-value problem (5.2.11) for system (5.2.1) (here C is a positive con-
stant independent of (z 0 , v 1 , µ, τ ), but depending on the order of differentiation
k = |k1 | + |k2 | + |k3 |).
1. If 0 < α < β, then
if |k1 | = |k2 | = 0
|k1 |+|k2 |+|k3 | ∗ ∗
C
∂ (z , v )
(a) k k
∂ (z 0 , µ) 1 ∂ (v 1 , τ ) 2 ∂tk3 if |k2 | = 0
≤ C e|k1 |αt
∂ |k1 |+|k2 |+|k3 | (z ∗ , v ∗ )
and |k1 |α < β ,
(b) k1 k2
∂ (z 0 , τ, µ) ∂ (v 1 ) ∂(τ − t)k3
if |k2 | 6= 0
C eβ(t−τ )+|k1 |ατ
or |k1 |α > β .
(5.2.26)
(5.2.27)
5.2. A boundary-value problem 295
(5.2.28)
t → τ − t, α → −β , β → −α , z → v, v → z, z 0 → v1 , v1 → z0 .
(5.2.29)
Therefore, estimates 1(b), 2(b) and 3(b) follow from estimates 3(a), 2(a) and
1(a), respectively, according to the rule above and the change k1 ↔ k2 .
Note also that in the cases β > α > 0 and α < β < 0 the parameter µ can
be included among the variables z or v respectively, by adding the equation
µ̇ = 0 to system (5.2.1) and the requirement µ(0) = µ (in the case β > α > 0)
or µ(τ ) = µ (in the case 0 > β > α) to the boundary conditions (5.2.11).
Therefore, the derivatives involving µ have to be estimated separately only in
the case α < 0 < β.
To obtain the estimates for the derivatives with respect to time t we note
that it follows directly from (5.2.1) that
∂ |k1 |+|k2 |+|k3 |+|k4 |+|k5 | z ∗ ∂ |k1 |+|k2 |+|k3 |+|k4 |+|k5 | (Az ∗ +f (z ∗ , v ∗ , µ, t))
= ,
∂(z 0 )k1 ∂(v 1 )k2 ∂µk3 ∂τ k4 ∂tk5 +1 ∂(z 0 )k1 ∂(v 1 )k2 ∂µk3 ∂τ k4 ∂tk5
(5.2.30)
∂ |k1 |+|k2 |+|k3 |+|k4 |+|k5 | v ∗ ∂ |k1 |+|k2 |+|k3 |+|k4 |+|k5 | (Bv ∗ +g(z ∗ , v ∗ , µ, t))
= .
∂(z 0 )k1 ∂(v 1 )k2 ∂µk3 ∂τ k4 ∂tk5 +1 ∂(z 0 )k1 ∂(v 1 )k2 ∂µk3 ∂τ k4 ∂tk5
One can see from these formulae that if the estimates of the theorem hold for
the derivatives with respect to (z 0 , v 1 , µ, τ ), then an additional differentiation
with respect to t does not affect these estimates (except possibly changing the
value of the constant C).
296 Chapter 5. Center Manifold. Local Case
The derivatives with respect to τ are expressed in terms of the other deriva-
tives via relation (5.2.24). It can be seen then that a differentiation with respect
to τ at fixed t must give essentially the same estimates as a differentiation with
respect to v 1 .
Thus, in the case α < β < 0, or 0 < α < β, it is sufficient to prove estimates
|k1 |+|k2 |
(z ∗ ,v ∗ )
(5.2.26) or, respectively, (5.2.28) for the derivatives ∂∂(z 0 )k1 ∂(v 1 )k2 , and in the
case α < 0 < β < it is sufficient to prove estimates (5.2.27) for the derivatives
∂ |k1 |+|k2 |+|k3 | (z ∗ ,v ∗ )
.
In fact, the calculation of these derivatives in the case
∂(z 0 )k1 ∂(v 1 )k2 ∂µk3
α < β < 0 is not necessary because it can be reduced to the case 0 < α < β
by applying the time reversion by rule (5.2.29). In the two remaining cases
0 < α < β and α < 0 < β the calculations are quite similar so we will present
the proof only for the more difficult case 0 < α < β (estimates for the first
derivatives for α < 0 < β can be found in Shilnikov [67]). It remains for us to
prove that
if |k2 | = 0
C(k) e|k1 |αt
∂ |k1 |+|k2 | ∗
(z , v ) ∗
and |k1 |α < β ,
k k2
≤ (5.2.31)
∂ (z 0 ) 1 ∂ (v 1 )
if |k2 | 6= 0
β(t−τ )+|k1 |ατ
C(k) e
or |k1 |α > β ,
∂(z ∗ , v ∗ )
≤ C eαt
∂z 0
(5.2.32)
∂(z ∗ , v ∗ )
≤ C e−β(τ −t) .
∂v 1
Since the first estimate is symmetric to the second with respect to the
time reversion (5.2.29), it is sufficient to prove only the first inequality in
(5.2.32).
∗ ∗
As mentioned above, the derivative (Z ∗ , V ∗ ) ≡ ∂(z∂z,v
0
)
can be found as the
unique solution of the boundary-value problem Z(0) = I, V (τ ) = 0 associated
with the system of variational equations (5.2.20). The existence of this solution
is guaranteed by Theorem 5.11 (see remarks after the theorem). Moreover, it
5.2. A boundary-value problem 297
Z τ
V̄ (t) =−
e−B(s−t) gz0 (z ∗ (s), v ∗ (s), µ, s)Z(s) + gv0 (z ∗ (s), v ∗ (s), µ, s)V (s) ds
t
(5.2.33)
Since k(f, g)0z,v k is bounded by some small ξ (see (5.2.4)), it follows from
(5.2.33) that
Z t
kZ̄(t)k ≤ eαt + ξ eα̃(t−s) k(Z(s), V (s))k ds ,
0
Z τ
kV̄ (t)k ≤ ξ e−β(s−t) k(Z(s), V (s))k ds .
t
then
ξ ξ
kZ̄(t)k ≤ 1+C eαt , kV̄ (t)k ≤ C eαt .
α − α̃ β−α
For k ≥ 2, the derivatives (Zk∗1 ,k2 , Vk∗1 ,k2 ) of the solution (z ∗ , v ∗ ) of the
boundary-value problem (5.2.11) satisfy the equation
Z t k ∗ ∗
A(t−s) ∂ f (z , v , µ, s)
∗
Z (t) = e ds ,
k1 ,k2
k k
0 ∂ (z 0 ) 1 ∂ (v 1 ) 2
Z (5.2.35)
τ
∗ −B(s−t) ∂ k g(z ∗ , v ∗ , µ, s)
V
k1 ,k2 (t) = − e k1 k2
ds .
t ∂ (z 0 ) ∂ (v 1 )
|p|
∂ |p| φ(ψ(x)) X ∂ i φ X ∂ |j1 | ψ ∂ |ji | ψ
= Cj1 ,...,ji · ··· ·
∂xp i=1
∂ψ i ∂xj1 ∂xji
j1 + · · · + j i = p
|j1 | ≥ 1, . . . , |ji | ≥ 1
for the derivatives of the superposition of functions. Here φ and ψ are some
∂iφ
vector functions, p and j1 , . . . , ji are multi-indices, ∂ψ i denotes the vector
k Z t
X ∂if
+ eA(t−s)
i=2 0 ∂(z, v)i (z ∗ (s),v ∗ (s))
X
× Cj1 ,...,ji (Zj∗11 ,j12 (s), Vj∗11 ,j12 (s)) · · · (Zj∗i1 ,ji2 (s), Vj∗i1 ,ji2 (s)) ds
j
Z τ
V̄k1 ,k2 (t) = − e−B(s−t) gz,v
0
(z ∗ (s), v ∗ (s), µ, s)(Zk1 ,k2 (s), Vk1 ,k2 (s)) ds
t
k Z τ
X ∂ig
− e−B(s−t)
i=2 t ∂(z, v)i (z ∗ (s),v ∗ (s))
X
× Cj1 ,...,ji (Zj∗11 ,j12 (s), Vj∗11 ,j12 (s)) · · · (Zj∗i1 ,ji2 (s), Vj∗i1 ,ji2 (s)) ds
j
(5.2.36)
where the inner summation is taken over all multi-indices j such that j 11 +
· · · + ji1 = k1 , j12 + · · · + ji2 = k2 and |jp1 | + |jp2 | ≥ 1 for all p = 1, . . . , i.
To derive the estimates (5.2.31) for (Zk∗1 ,k2 , Vk∗1 ,k2 ) we follow the same pro-
cedure as in the case of the first derivatives. It is sufficient to check that if
(Zk1 ,k2 (s), Vk1 ,k2 (s)) satisfies these estimates, then (Z̄k1 ,k2 (t), V̄k1 ,k2 (t)) satis-
fies them as well, with the same constant C(q + 1).
Note that the second integrals in formula (5.2.36) include only the deriva-
tives of orders less than or equal to q = k − 1: since |j1 | + · · · + |ji | = k, it
follows that if |jp | = k for some p = 1, . . . , i, then all the other j’s must be zero
which is not the case (the summation is taken over non-zero multi-indices).
Therefore, according to the induction hypothesis, estimates (5.2.31) hold for
(Zj∗p1 ,jp2 (s), Vj∗p1 ,jp2 (s)) in (5.2.36). In particular, if |k2 | = 0 (no differentia-
tion with respect to v 1 ) and |k1 |α < β, then jp2 ≡ 0 and |jp1 |α < β for all
p = 1, . . . , i. Thus, in this case,
and
i
Y
k(Zj∗p1 ,jp2 (s), Vj∗p1 ,jp2 (s))k ≤ C(q)i e(|j11 |+···+|ji1 |)αs = C(q)i e|k1 |αs .
p=1
(5.2.38)
If |k2 | 6= 0, then at least one of jp2 is non-zero and the corresponding term in
the product can be estimated as follows:
k(Zj∗p0 1 ,jp0 2 (s), Vj∗p0 1 ,jp0 2 (s))k ≤ C(q)e−β(τ −s)+|jp0 1 |ατ . (5.2.39)
(compare with (5.2.31): we used α > 0 and t ≤ τ , so e|jp1 |αt ≤ e|jp1 |ατ ; we also
used β > 0 so e−β(τ −t) ≤ 1). It follows from these estimates that
i
Y
k(Zj∗p1 ,jp2 (s), Vj∗p1 ,jp2 (s))k ≤ C(q)i e−β(τ −s) e(|j11 |+···+|ji1 |)ατ
p=1
Finally, if |k2 | = 0 but |k1 |α > β, then for some multi-indices j these products
may be estimated by (5.2.38), and for the others by (5.2.41). Note that if
|k1 |α > β, then e−β(τ −s)+|k1 |ατ > e|k1 |αs at s ≤ τ ; i.e. in this case the estimate
(5.2.41) majorizes (5.2.38). Therefore, if |k2 | = 0 but |k1 |α > β, then all
products in the second integrals of (5.2.36) satisfy (5.2.41), as if |k2 | 6= 0.
Recall that all derivatives of f and g are uniformly bounded. Thus, it follows
from the above considerations that
Z t
kZ̄k1 ,k2 (t)k ≤ ξeα(t−s) kZk1 ,k2 (s)k ds
0
Z t
∗
eα(t−s) e|k1 |αs ds
C (q) if |k2 | = 0 and |k1 |α < β ,
0
+
Z t
∗
C (q)
eα(t−s) e−β(τ −s)+|k1 |ατ ds if |k2 | 6= 0 or |k1 |α > β ,
0
5.2. A boundary-value problem 301
Z τ
kV̄k1 ,k2 (t)k ≤ ξe−β̃(s−t) kVk1 ,k2 (s)k ds
t
Z τ
∗
C (q)
e−β̃(s−t) e|k1 |αs ds if |k2 | = 0 and |k1 |α < β ,
t
+ Z τ
C ∗ (q) e−β̃(s−t) e−β(τ −s)+|k1 |ατ ds
if |k2 | 6= 0 or |k1 |α > β ,
t
(5.2.42)
∗
where C (q) is some constant and β̃ > β is chosen close to β such that the
spectrum of the matrix B still lies strictly to the right of the line Re (·) = β̃.
This means that the following modification of the estimate (5.2.3) for the
matrix exponent holds:
e−Bs ≤ e−β̃s for s ≥ 0 .
According to (5.2.31), if |k2 | = 0 and |k1 |α < β, we have k(Zk1 ,k2 (s),
Vk1 ,k2 (s))k ≤ C(q + 1)e|k1 |αs . Substituting this into (5.2.42) gives
Z t
kZ̄k1 ,k2 (t)k ≤ eαt (ξC(q + 1) + C ∗ (q)) e(|k1 |−1)αs ds
0
∗
ξC(q + 1) + C (q) |k1 |αt
≤ e ,
(|k1 | − 1)α
Z τ
β̃t ∗
kV̄k1 ,k2 (t)k ≤ e (ξC(q + 1) + C (q)) e−(β̃−|k1 |α)s ds
t
k(Zk1 ,k2 (s), Vk1 ,k2 (s))k ≤ C(q + 1)e−β(τ −s) e|k1 |ατ .
then k(Z̄k1 ,k2 (t), V̄k1 ,k2 (t))k satisfies estimates (5.2.31) with the same constant
C(q + 1).
This completes the proof of the theorem.
For our purposes, the most important property of the globally dichotomic
systems introduced in the previous section is the existence of some invariant
foliation. We will prove the existence of this foliation by considering the limit
case of the above boundary-value problem which corresponds to τ = +∞ (we
have already used such method in Sec. 2.8). Recall that we call a system of
differential equations globally dichotomic if it has the form
ż = Az + f (z, v, µ, t) ,
(5.3.1)
v̇ = Bv + g(z, v, µ, t) ,
5.3. Invariant foliation 303
eA s ≤ eαs ,
(5.3.3)
e−B s ≤ e−βs .
Definition 5.1. Take any point (z0 , v0 ). Let (z0 (t), v0 (t)) be the trajectory
which starts with (z0 , v0 ) at some t = t0 . We denote as Wγs (z0 , v0 , t0 ) the set
of points (z1 , v1 ) such that the trajectory (z1 (t), v1 (t)) of (z1 , v1 ) starting with
the same t = t0 satisfies
for all t ≥ t0 . We call Wγs (z0 , v0 , t0 ) the conventionally stable or γ-stable set
of (z0 , v0 ) at t = t0 .5
Theorem 5.13. For any (z0 , v0 , t0 ), for any γ ∈ (α, β), the conventionally
stable set Wγs is a C q -smooth manifold (where q is a maximal integer such that
qα < β and q ≤ r) of the type
v = ϕ(z; z0 , v0 , t0 , µ) ,
where the function ϕ does not depend on γ; it is defined at all z and depends
continuously on (z0 , v0 , µ, t0 ).
5 Here we are concerned about the starting moment t = t because we consider a non-
0
autonomous system, so different starting moments correspond to different trajectories. Of
course, in the autonomous case where f and g do not depend on time the value of t 0 does
not matter.
304 Chapter 5. Center Manifold. Local Case
Proof. As in the previous section, a solution (z(t), v(t)) satisfies the follow-
ing integral relation
Z t
z(t) = eA(t−t0 ) z(t0 ) + eA(t−s) f (z(s), v(s), µ, s)d ,
t0
Z (5.3.5)
τ
−B(τ −t) −B(s−t)
v(t) = e v(τ ) − e g(z(s), v(s), µ, s)ds
t
for any τ . Thus, if a point (z1 , v1 ) belongs to the γ-stable set of a point (z0 , v0 ),
then
z1 (t) − z0 (t) = eA(t−t0 ) (z1 (t0 ) − z0 (t0 ))
Z t
+ eA(t−s) [f (z1 (s), v1 (s), µ, s) − f (z0 (s), v0 (s), µ, s)]ds ,
t0
Z +∞
v1 (t) − v0 (t) = − e−B(s−t) [g(z1 (s), v1 (s), µ, s) − g(z0 (s), v0 (s), µ, s)]ds
t
(5.3.6)
−B(τ −t) γτ
(we took into account that e e → 0 as τ → +∞ for any fixed t, and
that v1 (τ ) − v0 (τ ) = O(eγτ ) by the definition of the γ-stable set).
Denote ζ(t) = z1 (t) − z0 (t), η(t) = v1 (t) − v0 (t). The solution of (5.3.6) is
a fixed point of the integral operator
Z t
ζ̄(t) = eA(t−t0 ) ζ 0 + eA(t−s) [f (z0 (s) + ζ(s), v0 (s) + η(s), µ, s)
t0
for all s ≥ t0 , then the operator (5.3.7) maps such a function into a func-
tion (ζ̄(t), η̄(t)) which satisfies the same condition. Moreover, exactly as in
Theorem 5.11 (compare with (5.2.17)), one can prove that the operator under
consideration is contracting in the γ-norm on the Banach space H[t0 ,+∞) of
functions satisfying (5.3.9).
Thus, according to the Banach contraction mapping principle, for any given
z1 (t0 ), the system (5.3.6) has a unique solution (z1 (t), v1 (t)) which satisfies
(5.3.4). Due to uniqueness, this solution is independent of the choice of γ from
the interval (α, β).
By Theorem 3.25 the solution depends continuously on (z0 , v0 , t0 , µ) and
on initial z = z1 (t0 ) = z0 (t0 ) + ζ 0 . In particular, we have that v = v1 (t0 ) is
a continuous function of z = z1 (t0 ). Thus, we have proved that the conven-
tionally stable manifold of any point z0 is a graph of some continuous function
v = ϕ(z).
Let us now prove the Cq -smoothness of the conventionally stable manifold.
It is equivalent to the Cq -smoothness of the solution (z1 (t), v1 (t)) of (5.3.6)
with respect to the initial condition z1 (t0 ). By the formal differentiation of
(5.3.6), we have that the first derivative
∗ ∗ ∂z1 (t) ∂v1 (t)
(Z (t), V (t)) ≡ , ,
∂z1 (t0 ) ∂z1 (t0 )
when it exists, satisfies the equation
Z t
Z ∗ (t) = eA(t−t0 ) + eA(t−s) fz,v
0
(z1 (s), v1 (s), µ, s)(Z ∗ (s), V ∗ (s))ds ,
t0
Z +∞
V ∗ (t) = − e−B(s−t) gz,v
0
(z1 (s), v1 (s), µ, s)(Z ∗ (s), V ∗ (s))ds .
t
(5.3.10)
The further derivatives
∂ k z1 (t) ∂ k v1 (t)
(Zk∗ (t), Vk∗ (t)) ≡ ,
∂z1 (t0 )k ∂z1 (t0 )k
must satisfy
Z t
Zk∗ (t) = eA(t−s) fz,v
0
(z1 (s), v1 (s), µ, s)(Zk∗ (s), Vk∗ (s))ds + Pk (t) ,
t0
Z +∞
Vk∗ (t) =− e−B(s−t) gz,v
0
(z1 (s), v1 (s), µ, s)(Zk∗ (s), Vk∗ (s))ds − Qk (t)
t
(5.3.11)
306 Chapter 5. Center Manifold. Local Case
where
Z t k
X ∂if
Pk (t) = eA(t−s)
t0 i=2
∂(z, v)i (z1 (s),v1 (s))
X
× Cj1 ,...,ji (Zj∗1 (s), Vj∗1 (s)) · · · (Zj∗i (s), Vj∗i (s))ds
j1 +···+ji =k
(5.3.12)
Z +∞ k
X ∂ig
Qk (t) = e−B(s−t)
t i=2
∂(z, v)i (z1 (s),v1 (s))
X
× Cj1 ,...,ji (Zj∗1 (s), Vj∗1 (s)) · · · (Zj∗i (s), Vj∗i (s))ds.
j1 +···+ji =k
Thus, when (Zj∗ , Vj∗ ) are known for j < k, the k-th derivative (Zk∗ , Vk∗ ) is the
fixed point of the operator
Z t
Z̄(t) = eA(t−s) fz,v
0
(z1 (s), v1 (s), µ, s)(Z(s), V (s))ds + Pk (t)
t0
Z +∞
V̄ (t) = − e−B(s−t) gz,v
0
(z1 (s), v1 (s), µ, s)(Z(s), V (s))ds − Qk (t) .
t
(5.3.13)
These equations are similar to Eqs. (5.2.35), (5.2.36) for the derivatives
of the solution of the boundary-value problem (5.2.11). Absolutely in the
same way as it was done there (Theorem 5.12), one can show that when
(Zj∗ (s), Vj∗ (s)) in (5.3.12) satisfy at j < k
for a small ε, then at kα < β, the integral which defines Qk (t) is convergent
and
kPk (t), Qk (t)k ≤ const ekαt .
Moreover, the operator (5.3.13) maps the space of functions (Z(t), V (t))
bounded in the γ-norm into itself, provided γ ∈ (max(α, kα), β), and it is
contracting in that norm.
Thus, once (Zj∗ (s), Vj∗ (s)) satisfying (5.3.14) are known at j < k, the formal
solution (Zk∗ (s), Vk∗ (s)) of (5.3.11) exists and satisfies (5.3.14) with j = k.
Therefore, by induction we get the existence of bounded in the γ-norm (γ ∈
(max(α, kα), β)) formal derivatives (Zk∗ (s), Vk∗ (s)) up to the order q.
5.3. Invariant foliation 307
To prove that the formal derivatives are the derivatives indeed, we will show
that the solution (z1 (t), v1 (t)) of system (5.3.6) is the limit, as τ → +∞, of a
solution (zτ∗ (t), vτ∗ (t)) of the boundary-value problem discussed in the previous
section, with the boundary data (z 0 = z1 (t0 ) = z0 (t0 ) + ζ 0 , v 1 = v0 (τ )), and
∗ ∗
that for each k = 1, . . . , q the k-th derivative (Zkτ (t), Vkτ (t)) of (zτ∗ (t), vτ∗ (t))
0 ∗ ∗
with respect to z converges to the solution (Zk (t), Vk (t)) of (5.3.12). Precisely,
we will prove that on any fixed finite interval of time
sup k(zτ∗ (t), vτ∗ (t)) − (z1 (t), v1 (t))k → 0 as τ → +∞ (5.3.15)
and
∂ k (zτ∗ (t), vτ∗ (t))
sup − (Zk∗ (t), Vk∗ (t)) → 0 as τ → +∞
∂(z 0 )k
k = 1, . . . , q (5.3.16)
from which the Cq -smoothness of (z1 (t), v1 (t)) with respect to z 0 follows im-
mediately.
To prove (5.3.15) note that the operator given by (5.3.7) is the limit of the
operator
Zt h
A(t−t0 ) 0
ζ̄(t) = e ζ + eA(t−s) f (z0 (s) + ζ(s), v0 (s) + η(s), µ, s)
t0
i
− f (z0 (s), v0 (s), µ, s) ds ,
Z τ h (5.3.17)
− e−B(s−t) g(z0 (s) + ζ(s), v0 (s) + η(s), µ, s)
t
i
η̄(t) =
− g(z 0 (s), v 0 (s), µ, s) ds for t ≤ τ
0 for t ≥ τ
More precisely, as τ → +∞, the operator (5.3.17) defined on the space of
functions (ζ, η) which are bounded in the γ-norm for some γ ∈ (α, β) has the
operator (5.3.7) as a limit in the γ 0 -norm for any γ 0 ∈ (γ, β). To prove this
statement it is sufficient to check that
Z +∞ h
0
sup e−γ t e−B(s−t) g(z0 (s) + ζ(s), v0 (s) + η(s), µ, s)
t≥t0 max(t,τ )
i
− g(z0 (s), v0 (s), µ, s) ds → 0 as τ → +∞
308 Chapter 5. Center Manifold. Local Case
where
Z t k
X ∂if
Pkτ (t) = eA(t−s)
t0 i=2
∂(z, v)i (zτ∗ (s),vτ∗ (s))
X
× Cj1 ,...,ji (Zj∗1 τ (s), Vj∗1 τ (s)) · · · (Zj∗i τ (s), Vj∗i τ (s))ds
j1 +···+ji =k
Z +∞ k
X ∂ig
Qkτ (t) = e−B(s−t)
t i=2
∂(z, v)i (zτ∗ (s),vτ∗ (s))
X
× Cj1 ,...,ji (Zj∗1 τ (s), Vj∗1 τ (s)) · · · (Zj∗i τ (s), Vj∗i τ (s))ds .
j1 +···+ji =k
(5.3.19)
The operator (5.3.18) takes functions bounded in the γ-norm (with γ ∈
(kα, β)) into functions bounded in the same norm and it is contracting in that
norm, uniformly for all τ (see Theorem 5.12). Thus, the solution satisfies
∗ ∗
kZkτ (t), Vkτ (t)k ≤ Ce(max(α,kα)+ε)t . (5.3.20)
Now, take k0 ≤ q and assume that (5.3.16) holds at all k < k0 . Let us
extend the operator (5.3.18) onto the space of functions defined at all t ≥
t0 , by assuming that the right hand side of the second equation in (5.3.18)
vanishes identically at t ≥ τ . As above, one can see that on the space of
functions bounded in the γ-norm (with γ ∈ (kα, β)) the integral operator
depends continuously on τ in the γ 0 -norm (γ 0 ∈ (γ, β)) and its limit as τ → +∞
is given by the operator (5.3.11).
Indeed, by (5.3.15) and by assumed validity of (5.3.16) for all j < k0 , we
have Z t
0
kPk (t) − Pkτ (t)kγ 0 ≤ sup e−γ t
eα(t−s) ϕ(s, τ )ds
t≥t0 t0
where ϕ(s, τ ) → 0 (as τ → +∞) uniformly on any fixed bounded interval of s.
Thus,
Z t
0
lim kPk (t) − Pkτ (t)kγ 0 ≤ lim sup e−γ t eα(t−s) ϕ(s, τ )ds (5.3.21)
τ →+∞ τ →+∞ t≥t(τ ) t0
for some t(τ ) which tends to infinity as τ → +∞. Note that ϕ is the norm of
difference between the sums entering the integrands in (5.3.12) and (5.3.19).
Therefore, by (5.3.14) and (5.3.20),
ϕ ≤ const eγs .
310 Chapter 5. Center Manifold. Local Case
whence
lim kQk (t) − Qkτ (t)kγ 0 = 0 .
τ →+∞
Note that the manifold Wγs (z0 , v0 , t0 , µ) is not, in general, invariant with
respect to system (5.3.1), with the only exception when the system is au-
tonomous and (z0 , v0 ) is an equilibrium state. In this case Wγs is the set of
points whose forward trajectories tend to the equilibrium in the γ-norm:
The function Φ defines the field of tangents to the leaves of the invariant
foliation: {(v − v0 ) = Φ(z0 , v0 , t0 , µ)(z − z0 ), t = t0 }. This field must be
invariant with respect to the linearized system. The leaves of the invariant
foliation are recovered by integrating the field of tangents; i.e. each leaf satisfies
the equation (for each fixed t0 )
∂v
= Φ(z, v, t0 , µ) . (5.3.23)
∂z
Therefore, being a solution of the differential equation above, the function
v = ϕ(z; z0 , v0 , t0 , µ) must have at least the same smoothness with respect to
the initial conditions (z0 , v0 , t0 ) and parameter µ, as the smoothness of Φ.
In general, the function Φ (and ϕ as well) is not smooth with respect to
(z0 , v0 , t0 , µ). Let us study the question of smoothness of the foliation in more
detail. Let β̃ ≥ 0 be a constant such that for the trajectory (z(t), v(t)), the
derivatives with respect to the initial conditions (z0 , v0 ) = (z(t0 ), v(t0 )) and µ
satisfy the following estimates
∂ k (z(t), v(t))
≤ const ekβ̃t . (5.3.24)
∂(z0 , v0 , µ)k
It can be proved that when the spectrum of the matrix
A 0
0 B
312 Chapter 5. Center Manifold. Local Case
lies strictly to the left of the imaginary axis, then the constant ξ in (5.3.2)
which bounds the derivatives of f and g may be taken to be so small that all
k
the derivatives ∂∂(z(z(t),v(t))
0 ,v0 ,µ)
k are bounded. Hence, β̃ = 0 in this case.
k Z t
X ∂i
+ eA(t−s) f 0 (z(s), v(s), µ, s)
i=1 t0 ∂(z0 , v0 , µ)i z,v
∗0 ∗0
· (Zk−i (s), Vk−i (s))ds ,
Z +∞ (5.3.25)
V̄ (t) = − e−B(s−t) gz,v
0
(z(s), v(s), µ, s)(Z(s), V (s))ds
t
k Z +∞
X ∂i
+ e−B(s−t) g 0 (z(s), v(s), µ, s)
i=1 t ∂(z0 , v0 , µ)i z,v
∗0 ∗0
· (Zk−i (s), Vk−i (s))ds .
To assure that the fixed point of this operator does give the k-th derivative
of (Z ∗ , V ∗ ) we may consider the family of operators, depending on τ , where
5.3. Invariant foliation 313
the infinite upper limit of the integral in the second equation is replaced by τ ,
and then take the limit as τ → +∞.6
By (5.3.24), the derivatives
∂i
(f, g)0z,v (z(s), v(s), µ, s) ,
∂(z0 , v0 , µ)i
entering (5.3.25) are estimated from above as const eiβ̃s . Based on this es-
timate, one can see, as in the proof of Theorem 5.12, that the integrals in
(5.3.25) converge, provided
α + k β̃ < β . (5.3.26)
Moreover, for any τ , the operators in the family under consideration take the
space of the functions bounded in the γ-norm (with γ ∈ (α + k β̃, β)) into itself,
and are contracting on this space uniformly with respect to τ .
Thus, we obtain that the function Φ is Ck where k is the maximal possible
integer such that (5.3.26) holds. Of course, the order of differentiation may
not be higher than (r − 1) because the right-hand side of (5.3.10) contains
Cr−1 -smooth functions (f, g)0z,v . We arrive at the following result.
Lemma 5.2. If for some β̃ ≥ 0 all the eigenvalues of the matrices B and A lie
strictly to the left of the line Re(·) = β̃, then the foliation by the conventionally
stable manifolds is C k -smooth (provided the constant ξ in (5.3.2) is sufficiently
small), where k is the maximal integer such that k < (β − α)/β̃ and k ≤ r − 1.
the invariant manifold theorems, but is also used in the analysis of non-local
bifurcations. Bearing in mind a future application of this sort, we stress the
following observation which was, in fact, already mentioned in the proof of
Theorem 5.12.
Lemma 5.3. Let (z ∗ (t; z 0 , v 1 , τ, µ), v ∗ (t; z 0 , v 1 , τ, µ)) be the solution of the
boundary-value problem z ∗ (0) = z 0 , v ∗ (τ ) = v 1 for system (5.3.1), and let
z 0 and v 1 depend on τ so that (z ∗ (0), v ∗ (0)) have some finite limit (z0 , v0 ) as
τ → +∞. Then, the derivative of v ∗ (0) with respect to z 0 tends to a value of
the function Φ, which defines the tangent to the conventionally stable manifold,
at the point (z0 , v0 ).
To prove this lemma, observe that by hypothesis the solution (z ∗ (t), v ∗ (t))
of the boundary-value problem tends to the trajectory of (z0 , v0 ) uniformly on
any fixed finite interval of t. Therefore, the statement of the lemma is nothing
more but a repetition of the claim in the proof of Theorem 5.13 that at k = 1
the fixed point of the integral operator (5.3.18) (finite τ ) does have the solution
of (5.3.10) (τ = +∞) as a limit. In the same way it follows from (5.3.16) that
under the assumption of Lemma 5.3, all the derivatives of v ∗ (0) with respect
to z 0 up to the order q have a finite limit as τ → +∞ (where q is the maximal
integer such that qα < β and q ≤ r).
In this section we complete our proof of the center manifold theorem. In fact,
we prove a more general result which embraces all local invariant manifold
theorems of this book.
Consider a local system of differential equations
ż = Az + f (z, v, µ)
(5.4.1)
v̇ = Bv + g(z, v, µ)
lie to the right of the line Re (·) = β in the complex plane. As shown in Sec. 5.2,
this system may be extended into the whole phase space such that the resulting
system is globally dichotomic. Theorem 5.13 then implies the existence of an
invariant manifold; namely, the γ-stable set of the point O. In fact, there is
a variety of invariant manifolds, depending on how we sub-divide the phase
variables into “z” and “v” parts: different choices of α and β would lead to
different separations of the spectrum of characteristic exponents and, therefore,
to different invariant manifolds.
Theorem 5.15. Under the hypotheses of the previous theorem, if α > 0, then
for all small µ the system has an extended stable invariant Cq -manifold W sE
(here q is the largest integer such that qα < β and q ≤ r) which is tangent to
{v = 0} at O at µ = 0 and which contains the set N + of all trajectories which
stay in a small neighborhood of O for all positive times. Although W sE is not
unique, any two of them have the same tangent at each point of N + . Moreover,
when W sE is written as by v = ϕsE (z), all derivatives of the function ϕsE are
uniquely defined at all points of N + , up to order q. The manifold W sE depends
continuously on µ and if the system is Cr -smooth with respect to all variables
including µ, then the manifold W sE is Cq with respect to µ.
from the definition: W ss is the set of all trajectories which tend to O faster
than the decrease in the exponent eγt . If α > 0, then γ > 0 and, therefore, the
manifold Wγs becomes a set of trajectories of system (5.3.1) which diverge from
the origin sufficiently slowly. Hence, which points in a small neighborhood of
O are included in W sE depends on how we extend the local system (5.4.1) onto
the whole phase space. This implies that W sE is not uniquely defined by the
local system. Nevertheless, regardless of the method of extension of the local
system, all points of the set N + , which is composed of forward trajectories
which never leave a small neighborhood of O, belong, by definition, to the
γ-stable set of O for any γ > 0. Therefore, every manifold W sE contains N + .
The uniqueness of the tangent to W sE at any point of N + does not follow
directly from Theorem 5.13 but this can nevertheless be extracted from its
proof. Indeed, we have shown that
∂ϕsE
= V ∗ |t=0 ,
∂z z=z0
Here, (z0 (s), v0 (s)) is the trajectory of the point (z0 , v0 ) = ϕsE (z0 ). It follows
from the proof of Theorem 5.13 that this solution is defined uniquely along
with all derivatives with respect to z0 up to order (q − 1). Consequently, since
for (z0 , v0 ) ∈ N + the trajectory of this point is defined by the local system
only, it follows that the derivatives of ϕsE at all points of N + are uniquely
defined.
Concerning the smoothness of W sE with respect to the parameters µ, we
note that in the case α > 0 we can include µ amongst the variables z upon
adding the equation µ̇ = 0 to system (5.4.1). Therefore, in this case the
smoothness with respect to µ is the same as with respect to z. If α < 0,
this no longer works, and the smoothness of the non-leading manifold with
5.4. Proof of theorems on center manifolds 317
respect to the parameters does not follow from Theorem 5.13. We will study
this question below in a more general framework on the smoothness of an
associated invariant foliation.
Theorems 5.14 and 5.15 allow us to reconstruct the following hierarchy of
local invariant manifolds. Let us choose a coordinate frame near an equilibrium
state O such that the linear part of the system assumes the Jordan form. We
have, in general,
ẏi = Ai yi + fi (x, y, z, µ)
żj = Cj zj + hj (x, y, z, µ) (5.4.3)
ẋ = Bx + g(x, y, z, µ)
where the spectrum of the matrix Ai lies on the straight line Re (·) = αi in the
complex plane, the spectrum of B lies on the imaginary axis,7 and the spectrum
of a matrix Cj lies on the straight line Re (·) = βj (here the indices i and j
assume a finite range of values); the function f , g and h are nonlinearities. Let
of the kind 8
s
W−i : (x, z, y1 , . . . , yi−1 ) = ϕss
i (yi , yi+1 , . . . )
s
W−1 : (x, z) = ϕss
1 (y)
where the functions ϕ vanish at zero along with the first derivatives.
7 In the structurally stable case this part of the spectrum is missing.
8 Here Wjs are Cq -smooth if qβj < βj+1 and q ≤ r, and W−i s are Cr -smooth (including
W0 ).
318 Chapter 5. Center Manifold. Local Case
Here, the manifolds with negative indices are given by Theorem 5.14 and
the others are from Theorem 5.15. They are embedded into each other by
construction: they are the local pieces of the corresponding conventionally
stable manifolds of O for some globally defined system and the latter are
embedded into each other by definition — the trajectories which converge
to O in the γ-norm, converge to O in the γ 0 -norm as well, for any γ 0 > γ.
The manifold W0s is the center stable manifold of Sec. 5.1 and the manifold
s
W−1 is the strongly stable manifold in this case. If the equilibrium is struc-
turally stable, then there is no characteristic exponents on the imaginary axis
and the manifold W0s is the stable manifold of O; it coincides with W−1 s
at
s
µ = 0 and the manifold W−2 is now the non-leading manifold of Sec. 2.6 —
s
the other manifolds W−i are, consequently the manifolds W sss , W ssss , etc.,
defined in that section.
For the case of structurally stable saddles, the manifold W1s gives the ex-
tended stable manifold defined in Sec. 2.7. In the case where all characteristic
exponents of O have positive real parts and where O is completely unstable,
the manifold W1s of O is the leading unstable manifold introduced in Sec. 2.6.
By applying Theorems 5.14 and 5.15 to the system which is derived from
(5.4.1) by a reversion of time, we obtain the following sequence of convention-
ally unstable invariant manifolds
u u
· · · ⊂ W−2 ⊂ W−1 ⊂ W0u ⊂ W1u ⊂ · · ·
where
u
W−i : (yi+1 , . . . ) = ψiuE (x, z, y1 , . . . , yi , µ)
W0u : y = ψ uC (x, z, µ)
where all functions ψ vanish at the origin along with their first derivatives.
This sequence includes all other invariant manifolds discussed in Chap. 2 and
in this chapter. In particular, W0u ∩W0s is the center manifold in the structurally
u
unstable case, and W−1 ∩ W0s is the saddle leading manifold (see Chap. 2) in
the structurally stable case.
For the system on the invariant manifold W0s , the equilibrium state does
not have positive characteristic exponents. Therefore, we can use Lemma 5.2
5.4. Proof of theorems on center manifolds 319
to assert the existence of the smooth invariant foliations on W0s (one should
extend first the system to the whole phase space, establish the existence of
the globally defined smooth invariant foliations and, then return to the local
system). This results in the following theorem.
Theorem 5.17. On W0s there is a family of strongly stable invariant Cr−1 -
ss
foliations F−i with Cr smooth leaves l−i
ss
of the kind
(i)
(x, y1 , . . . , yi−1 ) = ηξ0 ,µ (yi , yi+1 , . . . ) ,
where ξ 0 denotes the point (x0 , y10 , . . . , yi−1
0
) of intersection of a corresponding
leaf with the invariant manifold W−(i−1) ∩ W0s . For any point M ∈ W0s , the
u
respect to the parameter (when the point O does not disappear as µ varies,
and when it depends smoothly on µ).
It follows from the remark to Lemma 5.2 that for C∞ -smooth systems the
non-leading manifold is C∞ -smooth with respect to parameters, provided the
equilibrium state is structurally stable (no characteristic exponents are on the
s
imaginary axes). Otherwise the smoothness of W−i with respect to µ is finite
only.
In the same manner where Theorem 5.13 is used to establish the existence
of different kinds of invariant manifolds near an equilibrium state, we can
also use this theorem to study periodic trajectories. A system of differential
equations near a periodic trajectory L of period τ may be written in the form
(see Chap. 3)
ż = Az + f (z, v, µ, t) ,
(5.4.4)
v̇ = Bv + g(z, v, µ, t) ,
320 Chapter 5. Center Manifold. Local Case
where
0
f (0, 0, 0, t) ≡ 0 , fz,v (0, 0, 0, t) ≡ 0 ,
0
g(0, 0, 0, t) ≡ 0 , gz,v (0, 0, 0, t) ≡ 0 .
i.e. the manifold Wγs (0, 0, t0 ) is invariant with respect to the map σ ◦ Xτ . This
map is nothing but the Poincaré map of the cross-section t = t0 (see Chap. 3 for
more details). Thus, we have established the existence of an invariant manifold
for the fixed point (0, 0) for the Poincaré map of the extended system. The
set of orbits starting from points on this manifold at the cross-section {t = t 0 }
gives the corresponding invariant manifold for the system itself. Similarly,
Lemma 5.2 can be used to assert the existence of certain smooth invariant
foliations. Now we can return to the local system, in exactly the same way as
we did in the case of equilibrium states.
Thus, we obtain a hierarchy of local invariant manifolds and foliations in
a small neighborhood of the periodic trajectory. The corresponding theorems
are the analogue of the above theorems which deal with equilibrium states.
Theorem 5.19. Under the hypotheses of the previous theorem, if α > 0, then
for all small µ the system has an extended stable (n + 1)-dimensional invariant
Cq -manifold W sE (here q is the largest integer such that qα < β and q ≤ r)
which is tangent to the eigensubspace corresponding to the first n multipliers
at each point of L at µ = 0, and which contains the set N + of all trajectories
which stay in a small neighborhood of L for all positive times. Though W sE
is not unique, any two of them have the same tangent at each point of N + .
Moreover, all derivatives up to order q are uniquely defined at all points of N + .
The manifold W sE depends continuously on µ, and if the system is Cr -smooth
with respect to all variables including µ, then the manifold W sE is Cq with
respect to µ.
ẏi = Ai yi + fi (x, y, z, µ, t)
żj = Cj zj + hj (x, y, z, µ, t) (5.4.5)
ẋ = Bx + g(x, y, z, µ, t)
where the spectrum of the matrix Ai lies on the straight line Re (·) = αi , the
spectrum of the matrix B lies on the imaginary axis and the spectrum of the
matrix Cj lies on a straight line Re (·) = βj (here the indices i and j assume a
322 Chapter 5. Center Manifold. Local Case
Theorem 5.20. There exist sequences of the conventionally stable and con-
ventionally unstable smooth local invariant manifolds
s s
· · · ⊂ W−2 ⊂ W−1 ⊂ W0s ⊂ W1s ⊂ · · ·
and
u u
· · · ⊂ W−2 ⊂ W−1 ⊂ W0u ⊂ W1u ⊂ · · ·
of the kind 10
s
W−i : (x, z, y1 , . . . , yi−1 ) = ϕss
i (yi , yi+1 , . . . ; t)
s
W−1 : (x, z) = ϕss
1 (y; t)
and
u
W−i : (yi+1 , . . . ) = ψiuE (x, z, y1 , . . . , yi , µ, t)
W0u : y = ψ uC (x, z, µ, t)
where the functions ϕ and ψ vanish at (x, y, z, µ) = 0 along with their first
derivatives.
On the invariant manifolds W0s and W0u there exist, respectively, a family
of strongly stable and a family of strongly unstable invariant Cr−1 -smooth
ss
foliations F−i and Fjuu with Cr -smooth leaves l−i ss
and, respectively, ljuu of the
kind
s(i)
(x, y1 , . . . , yi−1 ) = ηξ0 ,µ (yi , yi+1 , . . . ; t0 )
and
u(j)
(x, z1 , . . . , zj−1 ) = ηξ0 ,µ (zj , zj+1 , . . . ; t0 )
10 The dependence on t is τ -periodic.
5.4. Proof of theorems on center manifolds 323
· · · ⊂ l2uu ⊂ l1uu .
Finally, we note that these theorems are easily reformulated in terms of the
Poincaré map: the intersection of the invariant manifolds with the cross-section
{t = t0 } gives the invariant manifolds for the fixed point at the origin.
Chapter 6
The local center manifold theorem is a well-known standard tool for the study
of bifurcations in a small neighborhood of equilibrium states and periodic tra-
jectories. However, as mentioned in the previous chapter, the local bifurcations
do not exhaust all important bifurcations. It has been known since the work
of Andronov and Leontovich [40] that among the four principal types of sta-
bility boundaries of a periodic trajectory of a two-dimensional system there
are two which correspond to the disappearance of a periodic trajectory via a
homoclinic loop — the union of an equilibrium state and a trajectory which
tends to the equilibrium state both as t → +∞ and as t → −∞. Though
they are at least equally important and no separation between different types
of two-dimensional bifurcations was made in the classical work by Andronov
and Leontovich, such objects are not considered in the theory of local bifur-
cations. A global bifurcation theory which deals with homoclinic loops, and
more complicated homoclinic and heteroclinic cycles, as well as other non-
local structures of multi-dimensional systems, had emerged after the works of
Shilnikov [60–62] in the mid sixties. This theory proved to be a good source of
different models of complex dynamical behaviors, as well as various scenarios
of transitions between different types of non-local dynamics. In this book (in
the second part) we will separate that part of the global theory which deals
especially with dynamical systems with simple behavior (non-chaotic). In this
part of the book we touch only the general question of the existence of an
analogue of a center manifold in the non-local case.
We started the study of this particular problem at the beginning of the
eighties. Since then, it has attracted the attention of many researchers. The
existence of the non-local center manifold near a homoclinic loop has now
325
326 Chapter 6. Center Manifold. Non-Local Case
been established by Turaev [73], Homburg [36] and Sandstede [56] (the latter
also embraces infinite-dimensional cases). Results on the existence of such
center manifolds for heteroclinic cycles have been derived by Shashkov [57] near
certain heteroclinic cycles. Here we give a detailed proof only for the simplest
case (when at least one leading exponent is real). We finish this chapter by
discussing necessary and sufficient conditions for the existence of the non-local
center manifold near arbitrarily complicated homoclinic and heteroclinic cycles
obtained recently by Turaev [75].
It is important to note a number of significant differences between the lo-
cal and the non-local center manifold theories. First, the dimension of the
non-local center manifold has no relation to the level of degeneracy of the as-
sociated bifurcation problem. In the local theory the dimension of the center
manifold is equal to the number of characteristic exponents on the imaginary
axis, which implies that a high dimension of the center manifold corresponds to
a large number of degeneracies in the linearized system. In contrast, even sim-
ple (codimension one) global bifurcation problems may not necessarily give rise
to a low-dimensional center manifold. Another notable distinction of global bi-
furcations from local bifurcations is that in the non-local case the smoothness
of the center manifold is not high. In fact, its smoothness does not corre-
late with the smoothness of the system and, in general, the non-local center
manifold is only C 1 .
Therefore, when studying specific non-local bifurcation problems, one pos-
sibly cannot apply the reduction to the center manifold directly: usually, sub-
tle questions require calculations involving derivatives of order higher than
the first order. Moreover, if the dimension of the center manifold is suffi-
ciently high, its presence gives practically no useful information. On the other
hand, if its dimension is low (dim W C = 1, 2, 3, 4), then the presence of a low-
dimensional invariant manifold which captures all trajectories remaining in its
neighborhood can tremendously simplify our understanding of the dynamics
of the system, even if the center manifold is only C 1 -smooth. In this case, one
can, at least, consider a low-dimensional model having some assumed smooth-
ness in order to make conjectures, which must be validated using the original
non-reduced system.
ẋ = F (x, µ) (6.1.1)
6.1. Center manifold for a homoclinic loop 327
(A) Let the system have a structurally stable equilibrium state O of the saddle
type. Assume that the characteristic exponents (λn , . . . , λ1 , γ1 , . . . , γm )
of O are ordered so that
and
(C) the homoclinic trajectory Γ does not lie in the non-leading unstable
submanifold W uu .
Assumption (C) implies that the trajectory Γ leaves the saddle point O
along the eigen-direction corresponding to the leading eigenvalue γ1 , as shown
in Fig. 6.1.1.
Conditions (A), (B) and (C) play different roles: condition (A) does not
involve bifurcations: it merely selects the class of systems under consideration.
If (A) is satisfied by the system itself, then it holds also for any nearby system
(i.e. for any system whose right-hand side is close to F along with the first
derivative). Moreover, once it is satisfied at µ = 0, it remains fulfilled for all
small µ as well.
328 Chapter 6. Center Manifold. Non-Local Case
Fig. 6.1.1. Condition (C) implies that the trajectory Γ leaves the saddle point O along the
eigen-direction corresponding to the leading eigenvalue γ1 .
As for condition (B), it cannot hold for all small µ; it can be shown that
if a system has a homoclinic loop, then for some nearby system the loop may
disappear (W s and W u would not have an intersection). Thus, condition (B)
defines µ = 0 as a bifurcational value for the parameter and specifies the
associated bifurcation phenomenon (the bifurcation of the homoclinic loop).
Generally, for any system whose right-hand side is close to F along with the
first derivative, there would exist a value of µ near zero for which the perturbed
system would also have a homoclinic loop.
Like condition (A), condition (C) does not imply any degeneracy. It just
assumes that the one-parameter family under consideration is in general po-
sition: if it is not satisfied for a given family, it can always be achieved by a
small perturbation of the right-hand side and once it is satisfied, it holds for
any close family as well.
Let q be the largest integer such that qγ1 < Re γ2 . Recall (see Sec. 2.7) that
under assumption (A) there exists an invariant Cmin(q,r) -smooth extended sta-
ble manifold W sE which is tangent at O to the eigenspace E sE corresponding
6.1. Center manifold for a homoclinic loop 329
Fig. 6.1.2. The extended stable manifold W sE which contains the stable manifold W s and is
tangent at the saddle point to the eigenspace corresponding to the characteristic exponents
λn , . . . , λ1 , γ1 . The manifold W sE it is not unique; any two of such manifolds have a common
tangent on W s . The strongly unstable sub-manifold W uu is uniquely embedded into the
smooth invariant foliation F u on W u .
Fig. 6.1.3. Continuation of the extended stable manifold W sE along the backward trajecto-
ries close to the homoclinic loop Γ.
Observe that condition (D) needs to be verified at only one point on the
trajectory Γ because the manifold W sE and the foliation F u are invariant with
respect to the flow defined by the system X0 . Note also that the manifold W sE
and the leaves of the foliation F u have complementary dimensions. Therefore,
our transversality condition (D) is well-posed. Like condition (C), it is a
condition of general position.
Theorem 6.1. If conditions (A), (B), (C) and (D) hold, then there
exists a small neighborhood U of the homoclinic trajectory Γ such that for all
sufficiently small µ the system Xµ possesses an (n + 1)-dimensional invariant
Cmin(q,r) -smooth center stable manifold W sC such that any trajectory which
does not lie in W sC leaves U as t → +∞. The manifold W sC is tangent at O
to the extended stable eigenspace E sE (Fig. 6.1.4).
The next two sections are devoted to the proof of this theorem. Note that
due to the symmetry of the problem with respect to a reversion of time, it
follows that there is a corresponding theorem on the center unstable manifold
6.1. Center manifold for a homoclinic loop 331
(A0 ) Let the characteristic exponents of the point O satisfy the following
condition:
In this case, since the leading stable eigenvalue λ1 is real, there exists an
(n − 1)-dimensional strongly stable sub-manifold W ss ⊂ W s .
Fig. 6.1.5. The extended unstable manifold W uE contains W u and is tangent at the saddle
point to the eigenspace corresponding to the characterstic exponents λ1 , γ1 , . . . , γm . The
manifold W uE is not unique; any two of such manifolds touch each other everywhere on
W u . The strongly stable sub-manifold W ss is uniquely embedded into the smooth invariant
foliation F s on W s .
Fig. 6.1.6. Continuation of the extended unstable manifold W uE along the forward trajec-
tories close to Γ.
6.1. Center manifold for a homoclinic loop 333
Theorem 6.2. If the conditions (A0 ), (B), (C0 ) and (D0 ) hold, there exists a
small neighborhood U of the homoclinic trajectory Γ such that for suffciently
small µ the system has an (m + 1)-dimensional invariant Cmin(p,r) -smooth
center unstable manifold W uC such that any trajectory outside of W uC leaves
U as t → −∞; see Fig. 6.1.7. (here p is the largest integer such that p|λ1 | <
|Re λ2 |). The manifold W uC is tangent at the point O to the eigenspace E uE
which corresponds to the characteristic exponents (γm , . . . , γ1 , λ1 ).
In the case where the conditions of both Theorems 6.1 and 6.2 hold, we
have the following result
the center manifold near a homoclinic loop is equal to the number of the leading
characteristic exponents (both negative and positive) and when this dimension
is greater than two, the bifurcations of such loop may be quite complicated in
some cases.
In the case of a homoclinic loop to a saddle-(1,1), the two-dimensional dy-
namics is relatively simple. Nevertheless, the reduction to the center manifold
requires some caution here. Our first observation is a low smoothness of W c .
Generally, it is only C1 and this may present an obstacle to a straightfor-
ward transcription of two-dimensional results into higher dimensions. Thus,
the two-dimensional theory of bifurcations of a homoclinic loop developed by
E. A. Leontovich produces a hierarchy of more and more degenerate cases
(corresponding to an increasingly large number of limit cycles appearing at bi-
furcation). The study of these cases requires an increasingly higher smoothness
of the system and, of course, the naive idea of simply repeating this hierar-
chy in the multidimensional situation, by referring to Theorem 6.3, would lead
to erroneous results. Unlike the case of local bifurcations, Theorems 6.1–6.3
essentially contain results of a qualitative rather than analytic nature.
Our second observation is that the manifold W c is not local (it is not
homeomorphic to a disc). Since its tangent at O is the leading plane E L , it
L
coincides locally with one of the saddle leading manifolds Wloc . At µ = 0, this
+
manifold must contain a piece Γ of the homoclinic trajectory Γ which lies in
s
Wloc and a piece Γ− of Γ which lies in Wloc
u
. From a small neighborhood of Γ+ a
L
small piece of Wloc can be continued by the forward trajectories along the loop
Γ until it reaches Γ− . The manifold obtained as a result of continuation must
return to a neighborhood of O in such a way that it can be glued smoothly
L
at this moment to the same local manifold Wloc — in order that a smooth
C
invariant manifold W can be formed. If the orientation is preserved, the
resulting glued manifold is a two-dimensional annulus. If not, the manifold
W c is a Möbius band. In fact, both cases are possible. Thus, in the multi-
dimensional case, the bifurcation of a homoclinic loop to a saddle-(1,1) are
reduced (generically) to a corresponding bifurcation either on the plane, or on
a two-dimensional non-orientable manifold.
In this and the next sections we present the proof of Theorem 6.1 which is
based on a study of the Poincaré map T defined by the trajectories of the
6.2. The Poincaré map 335
Fig. 6.2.1. The Poincaré map represented as a superposition of two maps: the local map
Tloc defined along the trajectories from the cross-section S in to S out near the saddle point
O, and the global map Tglo defined by the trajectories starting from S out and ending on S in
along the global part of the homoclinic loop Γ.
u̇ = Au + f (u, y, w, µ) ,
ẏ = γy + g(u, y, w, µ) , (6.2.1)
ẇ = Bw + h(u, y, w, µ) ,
∂(f, g, h)
(f, g, h)(0, 0, 0, 0) = 0 , = 0. (6.2.4)
∂(u, y, w, µ) (x,y,z,µ)=0
d s
It follows from Theorem 2.4 that dt kuk < 0 on Wloc ; i.e. kuk strictly decreases
s
along the trajectories in Wloc . This implies that for a sufficiently small ξ the
s
surface kuk = ξ is transverse to the trajectories on Wloc and, therefore, to all
close orbits. Consequently, being a part of this cross-section, the area S in is
transverse to the trajectories close to Γ provided that µ is sufficiently small.
Since the trajectory Γ does not lie in the non-leading unstable sub-manifold
W uu (condition (C) of Theorem 6.1), it leaves the saddle O along the leading
direction which coincides with the y-axis. Without loss of generality, we can
assume that Γ leaves O towards positive values of y. In this case, for sufficiently
small y − > 0, the homoclinic trajectory penetrates the surface {y = y − } at
some point M − ∈ Wloc u
. Denote M − = (u− , y − , w− ). Since at µ = 0 the
trajectory Γ is transverse to {y = y − }, it follows that at all small µ the small
area
S out = y = y − , u − u− , w − w − ≤ δ , (6.2.6)
Lemma 6.1. There exist functions uloc and wloc defined on ku0 − u+ k ≤
δ, kw1 − w− k ≤ δ and 0 < y 0 − ψ s (u0 , µ) ≤ δ 0 for some small δ 0 , such that for
two points M 0 ∈ S in and M 1 ∈ S out , the relation M 1 = Tloc M 0 holds if and
only if
u1 = uloc (u0 , y 0 , w1 , µ) , w0 = wloc (u0 , y 0 , w1 , µ) . (6.2.7)
338 Chapter 6. Center Manifold. Non-Local Case
∂uloc n o
≤ C max 1, eγ−λ+ε)τ , (6.2.9)
∂µ
∂uloc
≤C, (6.2.10)
∂ω 1
∂ωloc
≤C, (6.2.11)
∂(u0 , y 0 , µ)
∂ωloc
≤ C e−(η−γ−ε)τ , (6.2.12)
∂ω 1
where C is some positive constant, λ, η and γ satisfy conditions (6.2.2) and
(6.2.3), a small positive ε can be made arbitrarily small if δ is sufficiently small.
Here τ (y 0 , u0 , w1 , µ) is the flight time from M 0 to M 1 ; it tends to infinity as
y 0 → ψ s (u0 , µ) and
∂γ ∂γ
≤ C e(γ+ε)τ , ≤C. (6.2.13)
∂(u0 , y 0 , µ) ∂ω 1
Furthermore,
lim uloc = ψ u (y − , w1 , µ) ,
y 0 →ψ s (u0 ,µ)
∂uloc ∂ψ u − 1
lim = (y , w , µ) ,
y 0 →ψ (u
s 0 ,µ) ∂w 1 ∂w (6.2.14)
∂uloc ∂ψ u − 1
lim = (y , w , µ) if γ < λ,
y 0 →ψ (u
s 0 ,µ) ∂µ ∂µ
∂wloc ∂ϕsE
lim 0 0
= (u0 , y 0 , µ) .
y 0 →ψ (u
s 0 ,µ) ∂(u , y , µ) ∂(u, y, µ)
Proof. As show in Sec. 2.7, for any positive τ > 0 and for any small
(u0 , y 1 , w1 ) there is a unique trajectory (u∗ (t), y ∗ (t), w∗ (t)) of the system which
6.2. The Poincaré map 339
u∗ (0) = u0 , y ∗ (τ ) = y 1 , w∗ (τ ) = w1 .
u1 = u∗ (τ ; u0 , y 1 , w1 , µ, τ ) ,
y 0 = y ∗ (0; u0 , y 1 , w1 , µ, τ ) , (6.2.15)
w0 = w∗ (0; u0 , y 1 , w1 , µ, τ )
(we took into account the fact that the solution (u∗ , y ∗ , w∗ ) depends on the
boundary data (u0 , y 1 , w1 ), on the flight time τ , and on µ; as shown in Sec. 2.8
the dependence is C r -smooth with respect to all variables). The boundary
value problem under consideration is a special case of the boundary-value prob-
lem considered in Sec. 5.2: one should consider u as the z-variable, and (y, w)
as the v-variable in terms of that section. The estimates of Theorem (5.12)
give in our case (one should assume α = λ and β = γ − ε in (5.2.27))
∂u∗
≤ C e−λτ ,
∂(u0 , τ )
∂u∗
≤C
∂(w1 , µ)
(6.2.16)
∂(y ∗ , w∗ )
≤ C e−(γ−ε)τ ,
∂(y 1 , w1 , τ )
∂(y ∗ , w∗ )
≤C
∂(u0 , µ)
y ∗ |τ =+∞ = ψ s (u0 , µ) ,
w∗ |τ =+∞ = ϕs (u0 , µ) , (6.2.17)
∗ u 1 1
u |τ =+∞ = ψ (y , w , µ) .
340 Chapter 6. Center Manifold. Non-Local Case
Moreover,
∂y ∗ ∂ψ s
= (u0 , µ) ,
∂(u0 , µ) τ =+∞ ∂(u0 , µ)
∂w∗ ∂ϕs
= (u0 , µ) , (6.2.18)
∂(u0 , µ) τ =+∞ ∂(u0 , µ)
∂u∗ ∂ψ u
= (y 1 , w1 , µ) .
∂(y 1 , w1 , µ) τ =+∞ ∂(y 1 , w1 , µ)
u1 = u∗∗ (τ ; u0 , y 0 , w1 , µ, τ ) ,
y 1 = y ∗∗ (τ ; u0 , y 0 , w1 , µ, τ ) , (6.2.19)
0 ∗∗ 0 0 1
w = w (0; u , y , w , µ, τ )
where (u∗∗ (t), y ∗∗ (t), w∗∗ (t)) is the solution of the boundary-value problem
for a system obtained from (6.2.1) via a continuation from a small neighbor-
hood of the origin onto the whole space Rn+m (note that once (6.2.15) is
satisfied, the solution stays in a small neighborhood of the origin and when ap-
plying the results of Sec. 5.2 concerning the boundary-value problem (6.2.20)
we should not worry about the influence of this continuation). The prob-
lem (6.2.20) is a particular case of the boundary-value problem considered in
Secs. 5.2, 5.3: now one should denote the variables (u, y) as the z-variable
and w as the v-variable and assume α = γ + ε and β = η. The estimates of
Theorem 5.12 give for this case (see (5.2.26a) and (5.2.26b))
∂(y ∗∗ , u∗∗ )
≤ C e(γ+ε)τ ,
∂(u0 , y 0 , µ, τ )
∂(y ∗∗ , u∗∗ )
≤C
∂w1
(6.2.21)
∂w∗∗
≤ C e−ητ ,
∂(w1 , τ )
∂w∗∗
≤C
∂(u0 , y 0 , µ)
6.2. The Poincaré map 341
It follows from this equation and from (6.2.25) and (6.2.26) that
∂τ ∂y ∗∗ ∂y ∗∗
=− ẏ|M 1 − ẇ . (6.2.27)
∂(u0 , y 0 , w1 , µ) ∂(u0 , y 0 , w1 , µ) ∂w1 M 1
Note that the denominator in this formula does not vanish (it is positive).
Indeed, since the trajectory Γ does not belong to W uu , it leaves the origin
tangentially to the y-axis (see Theorem 2.5). Hence, the value of w − is much
less than y − . In particular, this means that ẏ|M 1 kẇkM 1 and our claim
∗∗
follows from the boundedness of ∂y ∂w1 (note that ẏ|M is positive since it is
1
− −
equal essentially to γy , and y is positive).
∂τ ∂y ∗
Thus, the inverse ∂y 0 to ∂τ exists which proves that the flight time is
∂τ
indeed uniquely defined by (u , y , w1 , µ). The required negativeness of ∂y
0 0
0
∗∗
follows from formula (6.2.27): ∂y ∂y 0 is equal, by definition, to 1 at τ = 0 and
since this derivative cannot vanish at any τ (by virtue of the first equation in
(6.2.26)) it remains positive for all τ .
From the relations (6.2.27) and (6.2.21) we obtain the inequality (6.2.13).
The functions uloc and wloc are defined as
One may check now that the estimates (6.2.16), (6.2.21) and (6.2.13) imply
(6.2.8)–(6.2.12) and the limit relations (6.2.18), (6.2.22) and (6.2.23) im-
ply (6.2.14). This completes our proof of the lemma.
6.2. The Poincaré map 343
The higher derivatives of the functions uloc and wloc , can also be easily esti-
mated using Theorem 5.12 and the identities (6.2.28) and (6.2.27) (in (6.2.27),
the values of ẏ and ẇ at the point M 1 (u1 , y − , w1 ) are evaluated by formulas
(6.2.1)). Omitting the obvious calculations the final result is as follow:
Lemma 6.3. Let us change the coordinates on the cross-sections S in and S out
in the following way:
0
ynew = y 0 − ψ s (u0 , µ) ,
0
wnew = w0 − ϕsE (u0 , y 0 , µ) on S in
u1new = u1 − ψ u (y 1 , w1 , µ) on S out
s
(we straighten the intersections Wloc ∩ S in and Wloc
u
∩ S out and make the inter-
section Wloc ∩ S tangent to {w = 0} at the point M + = Γ ∩ S in at µ = 0).
sE in 0
Points M 0 ∈ S in and M 1 ∈ S out are related by the map Tloc if, and only if,
where the functions uloc and wloc are now defined at y 0 ∈ [0, δ 0 ]; they satisfy
the following inequalities in the new coordinates:
and
∂ k wloc
(u0 , 0, w1 , µ) ≡ 0 , (6.2.33)
∂(u0 , y 0 , µ)k
at k ≤ min(q, r).
This lemma follows immediately from the two previous lemmas. Observe
that it follows from (6.2.32) that
∂uloc
≡0 at y0 = 0 . (6.2.34)
∂w1
Let us now consider the global map Tglo : S out 7→ S in . Since the flight time
from S in to S out is bounded (and it depends smoothly on the initial point)
the map Tglo is a Cr -diffeomorphism. Somehow, it is more convenient for us
−1
to consider the inverse map Tglo . Since it is Cr -smooth as well, one may write
−1
near the point M the map Tglo : S in → S out in the form
+
u0 − u +
u1 − u− (µ) d11 d12 d13 y0
=
w1 − w− (µ) d21 d22 d23
w0
uglo (u0 , y 0 , w0 , µ)
+ . (6.2.35)
wglo (u0 , y 0 , w0 , µ)
−1
Here (u− (µ), w− (µ)) are the coordinates of the image Tglo M + of the point
+ − out
M . At µ = 0 it is the point M = Γ ∩ S . Recall that in the coordinates
6.3. Proof of the center manifold theorem near a homoclinic loop 345
of Lemma 6.3 u− (0) = 0. The constants d11 , d12 , d13 , d21 , d22 and d23 are
matrices of dimensions n×(n−1), n×1, n×(m−1), (m−1)×(n−1), (m−1)×1
and (m − 1) × (m − 1), respectively; the functions uglo and wglo denote the
nonlinear terms.
Recall that by assumption the manifold W sE intersects transversely the
leaves of the strongly unstable foliation (condition (D)) at the points of Γ at
µ = 0. Therefore, the intersection of the tangent to the continuation of W sE
at the point M − with the tangent to the leaf l uu at the same point is zero.
u
The tangent to Wloc at this point is spanned to the tangent to l uu and to the
phase velocity vector (u̇, ẏ, ẇ)|M − which is tangent to Γ at M − . This vector
is also contained in the tangent to W sE (because W sE contains Γ). Thus, the
intersection of the tangents to W sE and Wloc u
at M − is one-dimensional (it is
spanned on the phase velocity vector). This implies that W sE and Wloc u
meet
− sE in −1
transversely at M . Therefore, the image of Wloc ∩ S by the map Tglo must
u
be transverse to the intersection Wloc ∩ S out . In the coordinates of Lemma 6.3
the latter is given by {u = 0} and Wloc ∩ S in is tangent to {w 0 = 0} at µ = 0.
1 sE
two maps that two points (u0 , y 0 , w0 ) and (ū0 , ȳ 0 , w̄0 ) are related by the map
T −1 : (u0 , y 0 , w0 ) 7→ (ū0 , ȳ 0 , w̄0 ) if, and only if,
Here ρ is a small constant. One can see that such multiplications do not change
the estimates of Lemma 6.3 in an essential way, just an additional constant
factor may appear. Observe that the functions uloc and wloc are kept the same
in a small ρ2 -neighborhood of (u+ , 0, 0) whereas they now vanish identically on
the boundary of the domain of definition, whence we may consider them to be
identically zero outside, without any loss of smoothness.
−1
The same procedure may be applied to the map Tglo — the functions uglo
ρ
and wglo may be modified outside the 2 -neighborhood of the point (u0 =
u+ , y 0 = 0, w0 = 0) so that they vanish at a distance ρ of that point, and this
−1
allows one to assume that Tglo is defined at all (u0 , y 0 , w0 ). Recall that uglo
and wglo are nonlinear functions. Hence, if ρ is sufficiently small, then the
−1
modified map Tglo is very close to its linear part everywhere.
6.3. Proof of the center manifold theorem near a homoclinic loop 347
where the function ũloc essentially satisfies the same estimates (given by a
Lemma 6.3) as the function uloc .
The substitution of this expression into (6.3.3) and into the first equation
of (6.3.1) represents the map T −1 in the cross-form
One can see, using the estimates of Lemma 6.3, that the functions F and G
satisfy the conditions of Theorem 4.4 (under the weakened smoothness condi-
tions given below that theorem; observe that the function uloc is not smooth
at y 0 = 0). Thus, we immediately have the existence of a C min(q,r) -smooth
manifold w 0 = φ∗ (u0 , y 0 , µ) which is invariant with respect to the modified
map T −1 . The function φ∗ is defined at all u0 , y 0 , µ. Since the modified map
coincides with the original map T −1 in a small neighborhood of (u+ , 0, 0) at
y 0 ≥ 0, it follows that the intersection of the above manifold with this domain
is a smooth invariant manifold of the original map.
By construction (see the proof of the annulus principle in Sec. 4.2), forward
iterations of any point by the modified map T −1 converge exponentially to
348 Chapter 6. Center Manifold. Non-Local Case
the “large” invariant manifold we found. This implies that all points whose
backward iterations are at a bounded distance on this manifold must lie in
this manifold. In terms of the original Poincaré map T this means that all
trajectories whose forward iterations lie in a small neighborhood of the point
(u0 , 0, 0) must belong to the “small” invariant manifold.
The set of trajectories which start from points of this manifold on the
cross-section is an invariant manifold for the system of differential equations
under consideration (one should choose the pieces of the trajectories until they
remain in a small neighborhood U of the homoclinic loop). By construction,
this manifold contains all trajectories which stay in U for all positive times.
s
In particular, it contains the intersection Wloc ∩ U . The point O must also be
included in the resulting invariant manifold. Note that the smoothness of the
above invariant manifold follows from the proven smoothness of its intersection
with the cross-section Sin — everywhere except at the equilibrium state O.
The smoothness of O must be verified separately, but we refrain from giving a
complete proof here because it is irrelevant for our purposes. Just note that the
resulting invariant manifold coincides locally with one of the extended stable
sE
manifolds Wloc (O) from which the smoothness at O follows.
The non-local center manifold theorem which we have proved for a homoclinic
loop admits a straightforward generalization onto a class of heteroclinic cycles.
Namely, suppose a family of C r -smooth dynamical systems
ẋ = X(x, µ) (6.4.1)
We need somewhat more: each wi here should only depend on (ui , yi , µ). The
graphs L∗i of these dependencies would define the invariant manifold on the
(i)
extended cross-sections Sin :
Ti L∗i = L∗i−1 .
To prove that the invariant manifold of the map (6.4.3) has the required
structure, it is sufficient to note that the invariant manifold is obtained in
Theorem 4.4 as the limit of the iterations of an arbitrary Lipschitz manifold.
Thus, if we make as our initial guess the manifold, say, (w1 = 0, . . . , wk = 0)
(1) (k)
which indeed represents the collection of independent surfaces on Sin , . . . , Sin ,
respectively, then just by the essence of the problem all iterations will have the
same structure. Hence their limit will also have the same structure.
The intersection of the derived surfaces L∗i with the original local pieces of
(i)
the cross-sections Sin define the invariant manifold for the original Poincaré
map. So, the set of trajectories which start on any of these surfaces is the
desired invariant manifold of the system itself.
Note that a reversion of time allows one to obtain an analogous center
unstable manifold theorem for the case when the leading negative exponent is
real and simple for each Oi , and a theorem on two-dimensional center manifold
when both positive and negative leading exponents are real and simple, as was
done in Sec. 6.1 for a homoclinic loop.
The heteroclinic cycles under consideration represents one of the simplest
cases among a large variety of possible heteroclinic or homoclinic structures.
For example, a single saddle equilibrium state may have more than one homo-
clinic loop at some value of µ (in a two-parameter family). We distinguish two
generic cases here:
Note that both cases correspond to the case where condition (C) holds for
both homoclinic trajectories: they do not belong to W uu and therefore leave
O along the leading direction, namely the y-axis.
6.4. Center manifold for heteroclinic cycles 351
Fig. 6.4.1. The homoclinic figure-eight for which the non-coincidence conditions are fulfilled:
the separatrix Γ1 intersects only those strongly stable leaves which are not intersected by
the separatrix Γ2 .
Fig. 6.4.2. The homoclinic butterfly composed by two loops Γ1 and Γ2 which does not satisfy
the non-coincidence conditions: the strongly stable leaf of an arbitrary point P 1 ∈ Γ1 lying
near the equilibrium state coincides with the strongly stable leaf of some point P 2 ∈ Γ2 .
352 Chapter 6. Center Manifold. Non-Local Case
Suppose that the transversality condition (D) is fulfilled for both homo-
clinic trajectories (see Sec. 6.1). Again, revisiting the construction of the
previous section one may prove the following result.
(this is the center part of the spectrum); n characteristic exponents lie to the
right of this strip:
Re λ > βiu
and m characteristic exponents lie to the left of this strip:
Re λ < −βis .
To be more accurate, we take into account the gap between the center part
and the strongly stable and strongly unstable parts and write
Re Λss < −βiss < −βis < Re Λc < βiu < βiuu < Re Λuu , (6.4.4)
These conditions are analogous to conditions (D) and (D0 ) from Sec. 6.1.
According to Theorems 5.16, 5.17 and 5.20, a trajectory Γs which tends to an
equilibrium state Oi , or to a periodic trajectory Li , as t → +∞, lies in an
(m + k)-dimensional extended stable manifold W sE of Oi or Li and through
each point of Γs a uniquely defined m-dimensional leaf of the strongly stable
foliation F ss exists; the tangent to W sE is also uniquely defined at each point
of Γs . Analogously, the trajectory Γs in the heteroclinic cycle tends to some
equilibrium state Oj , or to a periodic trajectory Lj , as t → −∞ and this implies
that Γs lies in an (n + k)-dimensional extended unstable manifold W uE of Oj ,
or Lj , (the tangent to W uE is uniquely defined at each point of Γs ) and through
each point of Γs a uniquely defined n-dimensional leaf of the strongly unstable
foliation F uu exists. The manifold W sE is tangent at Oi or Li to the extended
stable invariant subspace E sE of the linearized system, corresponding to the
critical and strongly stable parts of the spectrum of characteristic exponents.
The manifold W uE is tangent at Oj , or Lj , to the extended unstable invariant
subspace E uE corresponding to the critical and strongly unstable parts of the
spectrum. The foliation F ss includes the strongly stable manifold which is
tangent at Oi , or Li , to the strongly stable invariant subspace E ss , and the
foliation F uu includes the strongly unstable manifold which is tangent at Oj
or Lj , to the strongly unstable invariant subspace E uu .
The transversality conditions are:
Observe that due to the invariance of the subspaces with respect to the
linearized system, the transversality must be verified at one point on each
trajectory Γs .
Different choices of the separation of the spectra of characteristic exponents
lead to different manifolds and foliations involved in the above transversality
conditions. For some of our choices transversality may hold, but for some it
may not hold. So these conditions do make an additional selection among
various possible trichotomic separations.
Theorem 6.6. Let q and p be the maximal integers such that βiuu > qβiu ,
βiss > pβis for any equilibrium state or periodic trajectory in the cycle (the β’s
are the separating constants from (6.4.4)).
6.4. Center manifold for heteroclinic cycles 355
We do not give the proof of this theorem. It includes the above non-local
center manifold theorems of this chapter as a special cases. For example, when
we consider a single homoclinic loop, the strongly unstable manifold is a par-
ticular leaf of the strongly unstable foliation. If the homoclinic trajectory lies
in this leaf, it intersects the leaf, formally speaking, in continuum points which
prevents of existence of the smooth invariant manifold under consideration.
Thus, conditions (C) and (C0 ) are necessary for the theorems of Sec. 6.1 to be
valid.
When we consider a pair of homoclinic loops, the leaves of the strongly
unstable foliation are surfaces of the kind
(u, y) = ψ(w)
in the coordinates of Sec. 6.2. One may straighten the foliation so that the
leaves are the intersections of the surfaces {y = constant} with the unstable
manifold. In the case of a homoclinic figure-eight, the leaves corresponding to
y > 0 intersect the homoclinic trajectory Γ1 at one point each, and the leaves
corresponding to y < 0 intersect the homoclinic trajectory Γ2 also at only one
point each. The strongly unstable manifold — the leaf corresponding to y = 0
356 Chapter 6. Center Manifold. Non-Local Case
357
358 Appendix A
Let us assume also that the eigenvalues (λ1 , . . . , λm1 ) of the matrix A1 (0)
have the same real part, namely
and that the real parts of the eigenvalues (γ1 , . . . , γn1 ) of the matrix B1 (0) are
equal to each other, i.e.,
With regard to the eigenvalues of the matrices A2 (0) and B2 (0), let us assume
that the real parts of the eigenvalues of A2 (0) are strictly less than λ, and
those of B2 (0) are strictly larger than γ. In this case x and y are the leading
stable and unstable coordinates, respectively, and u and v are the non-leading
coordinates.
where fij , gij are Cr−1 with respect to (x, u, y, v) and their first derivatives
with respect to (x, u, y, v) are Cr−2 with respect to (x, u, y, v, µ), and
x̃ = x − ϕ1s (y, v, µ) ,
ũ = u − ϕ2s (y, v, µ) ,
(A.4)
ỹ = y − ψ1u (x, u, µ) ,
ṽ = v − ψ2u (x, u, µ) ,
where {x = ϕ1s (y, v, µ), u = ϕ2s (y, v, µ)} and {y = ψ1u (x, u, µ), v =
ψ2u (x, u, µ)} are the equations of the stable and the unstable manifolds of the
saddle point, respectively. This transformation does not give us the identities
(A.3); by now we have that the functions fij and gij in (A.2) are Cr−1 -smooth
and vanishing at the origin.
We can also recast system (A.2) into the form
where
Ri = fi1 (x, u, 0, 0, µ)x + fi2 (x, u, 0, 0, µ)u ,
Pi = gi1 (0, 0, y, v, µ)y + gi2 (0, 0, y, v, µ)u ,
ϕ1 = f11 (0, 0, y, v, µ), ϕ2 = f12 (0, 0, y, v, µ) ,
ϕ3 = f21 (0, 0, y, v, µ), ϕ4 = f22 (0, 0, y, v, µ) ,
ψ1 = g11 (x, u, 0, 0, µ), ψ2 = g12 (x, u, 0, 0, µ) ,
ψ3 = g21 (x, u, 0, 0, µ), ψ4 = g22 (x, u, 0, 0, µ) ,
and
Ri (x, u, µ) = R̃i1 (x, u, µ)x + R̃i2 (x, u, µ)u ,
Pi (y, v, µ) = P̃i1 (y, v, µ)y + P̃i2 (y, v, µ)v ,
R̃ij (0, 0, µ) ≡ 0, P̃ij (0, 0, µ) ≡ 0 ,
ϕj (0, 0, µ) ≡ 0, ψj (0, 0, µ) ≡ 0 ,
and the ellipsis denotes the terms which we will hereafter call negligible: in
the first two equations these are the terms of the form f˜(x, u, y, v, µ)x and
360 Appendix A
and in the last two equations these are the terms of the form g̃(x, u, y, v, µ)y
and g̃(x, u, y, v, µ)v such that
(1)
ξ1 = x + h1 (y, v, µ)x, ξ2 = u + h2 (y, v, µ)x ,
η1 = y, η2 = v ,
where hi (0, 0, µ) = 0;
(2)
ξ1 = x, ξ2 = u ,
η1 = y + s1 (x, u, µ)y, η2 = v + s2 (x, u, µ)y ,
where si (0, 0, µ) = 0;
(3)
ξ1 = x + r1 (x, u, µ)x + r2 (x, u, µ)u, ξ2 = u ,
η1 = y, η2 = v ,
where r1 (0, 0, µ) = 0, r2 (0, 0, µ) = 0;
(4)
ξ1 = x, ξ2 = u ,
η1 = y + p1 (y, v, µ)y + p2 (y, v, µ)v, η2 = v ,
where p1 (0, 0, µ) = 0, p2 (0, 0, µ) = 0 .
The change of variables (1) gets rid of the terms ϕ1 and ϕ3 in system (A.5).
By a change of variables (2) we eliminate the terms ψ1 and ψ3 . By a change of
variables (3) we eliminate the terms R1 . Finally by a change of variables (4)
we eliminate the terms P1 , thereby reducing the original system to the desired
form.
Special form of the systems near a saddle 361
(A.8)
362 Appendix A
∂h2 ∂h2
ξ˙2 = u̇ + ẏx + v̇x + h2 (y, v, µ)ẋ
∂y ∂v
= A2 (µ)u + R2 (x, u, µ) + ϕ3 (y, v, µ)x + ϕ4 (y, v, µ)u
∂h2 ∂h2
+ B1 y + P1 (y, v, µ) x + B2 v + P2 (y, v, µ) x
∂y ∂v
+ h2 (y, v, µ) A1 (µ)x + ϕ1 (y, v, µ)x + ϕ2 (y, v, µ)u + . . .
∂h2 ∂h2
+ B1 (µ)y + P1 (y, v, µ) + B2 (µ)v + P2 (y, v, µ)
∂y ∂v
The form of the third and fourth equations is not affected by such change of
variables.
We assume that the functions h1 (y, v, µ) and h2 (y, v, µ) satisfy the following
conditions
A 1 h 1 − h 1 A 1 − ϕ 1 + ϕ 2 h 2 − h 1 ϕ1 + h 1 ϕ2 h 2
∂h1 ∂h1
= (B1 y + P1 ) + (B2 v + P2 ) ,
∂y ∂v
(A.10)
A 2 h 2 − h 2 A 2 − ϕ 3 + ϕ 4 h 2 − h 2 ϕ1 + h 2 ϕ2 h 2
∂h2 ∂h2
= (B1 y + P1 ) + (B2 v + P2 ) .
∂y ∂v
Special form of the systems near a saddle 363
This implies that the expressions inside the square brackets in (A.8) and (A.9)
vanish. Consider next the following system of matrix equations:
Ẋ = A1 X − XA1 − ϕ1 + ϕ2 U − Xϕ1 + Xϕ2 U ,
U̇ = A2 U − U A1 − ϕ3 + ϕ4 U − U ϕ1 + U ϕ2 U ,
(A.11)
ẏ = B1 y + P1 ,
v̇ = B2 v + P2 .
Here, the matrices X ∈ Rm1 ×m1 and U ∈ Rm2 ×m1 , where m2 is the dimension
of the vector u. System (A.11) has an equilibrium state O1 (0, 0, 0, 0). The
linearized system is
∂ϕ1 ∂ϕ1
Ẋ = A1 X − XA1 − (0, 0, µ)y − (0, 0, µ)v ,
∂y ∂v
∂ϕ3 ∂ϕ3
U̇ = A2 U − U A1 − (0, 0, µ)y − (0, 0, µ)v ,
∂y ∂v
ẏ = B1 y ,
v̇ = B2 v .
The spectrum of characteristic exponents of this system may be represented
as a union of the spectra of the following associated linear operators
X 7→ A1 X − XA1 ,
U 7→ A2 U − U A1 ,
y 7→ B1 y ,
v 7→ B2 v .
Let us now recall a well-known fact from matrix theory (see [39]), namely, for
any square matrices A and B the spectrum of the operator Z 7→ AZ − ZB
(where Z is a rectangular matrix) belongs to the set of numbers generated all
possible differences between the eigenvalues of the matrices A and B.
Then, since the eigenvalues of the matrix A2 lie to the left of the eigen-
values of the matrix A1 , and the latter lie all on the line Re· = λ, it follows
that when µ = 0 the equilibrium state of system (A.11) possesses m21 charac-
teristic exponents on the imaginary axis, m1 · m2 characteristic exponents in
the open left-half plane, and n1 + n2 = n characteristic exponents in the open
right-half plane. Therefore, the equilibrium state of system (A.11) has an in-
variant n-dimensional strongly unstable manifold W̃1uu defined by the equation
364 Appendix A
∂s2 ∂s2
η̇2 = v̇ + ẋy + u̇y + s2 (x, u, µ)ẏ
∂x ∂u
= B2 (µ)v + P2 (y, v, µ) + ψ3 (x, u, µ)y + ψ4 (x, u, µ)v
∂s2 ∂s2
+ A1 x + R1 (x, u, µ) y + A2 u + R2 (x, u, µ) y
∂x ∂u
+ s2 (x, u, µ)v B1 (µ)y + ψ1 (x, u, µ)y + ψ2 (x, u, µ)v + . . .
We choose the functions s1 and s2 such that the expressions inside the square
brackets become identically equal to zero, i.e.
B 1 s 1 − s 1 B 1 − ψ 1 + ψ 2 s 2 − s 1 ψ1 + s 1 ψ2 s 2
∂s1 ∂s1
= (A1 x + R1 ) + (A2 u + R2 ) ,
∂x ∂u
(A.12)
B 2 s 2 − s 2 B 2 − ψ 3 + ψ 4 s 2 − s 2 ψ1 + s 2 ψ2 s 2
∂s2 ∂s2
= (A1 x + R1 ) + (A2 u + R2 ) .
∂x ∂u
To show that such s1 and s2 exist, consider the matrix system
ẋ = A1 x + R1 ,
u̇ = A2 u + R2 ,
(A.13)
Ẏ = B1 Y − Y B1 − ψ1 + ψ2 V − Y ψ1 + Y ψ2 V ,
V̇ = B2 V − V B1 − ψ3 + ψ4 V − V ψ1 + V ψ2 V ,
366 Appendix A
2
where Y ∈ Rn1 and V ∈ Rn1 n2 . For all small µ this system has an equilibrium
state O2 (0, 0, 0, 0). The linearized system is
ẋ = A1 x ,
u̇ = A2 u ,
∂ψ1 ∂ψ1
Ẏ = B1 Y − Y B1 − (0, 0, µ)x − (0, 0, µ)u ,
∂x ∂u
∂ψ3 ∂ψ3
V̇ = B2 V − V B1 − (0, 0, µ)x − (0, 0, µ)u .
∂x ∂u
A1 (µ) 0 B1 (µ) 0
A(µ) = , B(µ) = ,
0 A2 (µ) 0 B2 (µ)
R1 (x, µ)
r(x, µ) = (r1 (x, µ), r2 (x, µ)) , R(x, µ) = ,
R2 (x, µ)
P1 (y, µ)
p(y, µ) = (r1 (y, µ), r2 (y, µ)) , P (y, µ) = .
P2 (y, µ)
In terms of the above new notation, the change of variables (3) assumes the
form
ξ1 = x + r(x, µ)x, ξ2 = u, η1 = y, η2 = v
Let R(x, µ) = R̃(x, µ)x, and, consequently, R1 (x, µ) = R̃1 (x, µ)x and R2 (x, µ) =
R̃2 (x, µ)x). After the change of variables we obtain
Special form of the systems near a saddle 367
∂r
ξ˙1 = ẋ + ẋx + r(x, µ)ẋ = A1 (µ)x + R1 (x, µ) + ϕ2 (y, µ)u
∂x
∂r
+ A(µ)x + R(x, µ) x + r(x, µ) A(µ)x + R(x, µ) + . . .
∂x
"
= A1 (µ)ξ1 + ϕ2 (η1 , η2 , µ)ξ2 + −A1 (µ)r(x, µ) + R̃1 (x, µ)
#
∂r
+ (A(µ)x + R̃(x, µ)x) + r(x, µ)A(µ) + r(x, µ)R̃(x, µ) x + . . . ,
∂x
the right of the imaginary axis, respectively. This implies that for sufficiently
small µ system (A.15) possesses an m-dimensional invariant manifold (strongly
stable) Y = r(x, µ). This gives the existence of the function r which satisfies
condition (A.14).
The transformation (3) with such r(x, µ) brings the system to the form (A.5)
with ϕ1 ≡ 0, ϕ3 ≡ 0, ψ1 ≡ 0, ψ3 ≡ 0 and R1 ≡ 0.
Step 4. Making the change of variables (4), we obtain
#
∂p
+ (B(µ)y + P̃ (y, µ)y) + p(y, µ)B(µ) + p(y, µ)P̃ (y, µ) y + . . . ,
∂y
∂p
(B(µ)y + P̃ (y, µ)y)
∂y (A.16)
= B1 (µ)p(y, µ) − p(y, µ)B(µ) − P̃1 (y, µ) − p(y, µ)P̃ (y, µ) .
where X ∈ Rn1 n and y ∈ Rn . For all µ sufficiently small this system pos-
sesses the equilibrium state O4 (0, 0) whose characteristic exponents comprise
the spectrum of the linear operator
∂ P̃1
X 7→ B1 (µ)X − XB(µ) − (0, µ)y ,
y
y 7→ B(µ)y .
(r − 1), except for the last (r − 1)-th derivative with respect to µ alone which may not exist.
371
372 Appendix B
For example, when there is only one stable leading multipliers (m1 = 1 and
λ1 is real), then λ0 (µ) = λ1 (µ); if there is a pair of complex-conjugate stable
leading multipliers (m1 = 2 and λ1 = λ∗2 is not real), then λ0 (µ) = Re λ1 (µ).
Analogously, γ0 (µ) = γ1 (µ) if n1 = 1; and γ0 (µ) = Re λ1 (µ) if n1 = 2 and
γ1 = γ2∗ is not real.
Since A and B depend smoothly on µ, we also have
∂q ∂q
A1 (µ)j ≤ const · j q λ0 (µ)j , B1 (µ)−j ≤ const · j q γ0 (µ)−j
∂µq ∂µq
(B.4)
at q = 1, . . . , r − 1.
Let us also introduce the quantities λ0 and γ 0 , satisfying λ20 < λ0 < λ0 and
γ0 < γ 0 < γ02 , such that for all j ≥ 0
and the same estimates hold true for all the derivatives with respect to µ.
where the terms o(λk0 ) and o(γ0−k ) are Cr−1 -smooth and all their derivatives
with respect to (x0 , u0 , y 1 , v 1 ) are also of order o(λk0 ) and o(γ0−k ) respectively,
the derivatives which involve differentiation q times with respect to µ are
estimated, respectively, as o(k q λk0 ) and o(k q γ0−k ) (q = 0, . . . , r − 2).
j−1
X
xj = Aj1 x0 + Aj−s−1
1 f1 (xs , us , ys , vs , µ) ,
s=0
j−1
X
uj = Aj2 u0 + Aj−s−1
2 f2 (xs , us , ys , vs , µ) ,
s=0
k−1
(B.9)
−(k−j) 1
X −(s+1−j)
yj = B1 y − B1 g1 (xs , us , ys , vs , µ) ,
s=j
k−1
X
−(k−j) 1 −(s+1−j)
vj = B 2 v − B2 g2 (xs , us , ys , vs , µ)
s=j
computed as
j−1
X
(n+1)
xj = Aj1 x0 + Aj−s−1
1 f1 (x(n) (n) (n) (n)
s , us , ys , vs , µ) ,
s=0
j−1
X
(n+1)
uj = Aj2 u0 + Aj−s−1
2 f2 (x(n) (n) (n) (n)
s , us , ys , vs , µ) ,
s=0
k−1
(B.11)
(n+1) −(k−j) 1
X −(s+1−j)
yj = B1 y − B1 g1 (x(n) (n) (n) (n)
s , us , ys , vs , µ) ,
s=j
k−1
X
(n+1) −(k−j) 1 −(s+1−j)
vj = B2 v − B2 g2 (x(n) (n) (n) (n)
s , us , ys , vs , µ) ,
s=j
(n) (n)
kxj k ≤ Kx λj0 , kuj k ≤ Ku λj0 ,
(B.13)
(n) (n)
kyj k ≤ Ky γ0j−k , kvj k ≤ Kv γ0j−k .
(n+1) (n+1)
kxj k ≤ Kx λj0 , kuj k ≤ Ku λj0 ,
(B.14)
(n+1) (n+1)
kyj k ≤ Ky γ0j−k , kvj k ≤ Kv γ0j−k .
Trajectories near a saddle fixed point 375
and
j−1 j−1 s
X
0 −s
X λ0
(λ ) f2 (x(n) (n) (n) (n)
s , us , ys , vs , µ) ≤ [ϕ̃2 (s) + δϕ2 (s)] .
s=0 s=0
λ0
and
j j−1 s
λ0 1 X λ0
ϕ2 (j) = (ε + [ϕ̃2 (s) + δϕ2 (s)]) . (B.21)
λ0 λ0 s=0 λ0
It is well known, that the sum of the kind
k−1
X
αs ϕ(k − s)
s=0
tends to zero as k → +∞ for any α < 1 and any sequence ϕ which tends to
zero as k − s → +∞. Therefore, Eq. (B.20) defines indeed a converging to zero
sequence ϕ1 (k) provided ϕ2 (s) tends to zero as s → +∞.
The sequence ϕ2 (j) is given by (B.21) which can be rewritten as
λ0 δ 1
ϕ2 (j + 1) = 1 + 0 · ϕ2 (j) + · ϕ̃2 (j) .
λ0 λ λ0
Trajectories near a saddle fixed point 377
0
Since λλ0 1 + λδ0 < 1 for sufficiently small δ and since ϕ̃2 (j) → 0 as j → ∞, it
follows from this formula that ϕ2 (j) tends to zero indeed.
By the symmetry of the problem, appropriate functions ψ1 and ψ2 are
found in an absolutely analogous way. We have proved the estimates (B.10).
To complete the lemma we need to show that analogous estimates hold for all
derivatives of the solution (xj , uj , yj , vj ) of (B.9).
It is shown in Sec. 3.7 that the successive approximations converge to the
solution of the boundary-value problem along with all derivatives. Thus, we
may assume that the n-th approximation satisfies2
(n) (p)
kDp xj − Dp A1 (µ)j x0 k ≤ k p2 λj0 ϕ1 (k) ,
(n) (p)
kDp uj k ≤ k p2 λj0 ϕ2 (j) ,
(B.22)
(n) −(k−j) (p)
kDp yj − Dp B1 (µ)−(k−j) y 1 k ≤ k p2 γ0 ψ1 (k) ,
(n) −(k−j) (p)
kDp vj k ≤ k p2 γ0 ψ2 (k − j) ,
for some converging to zero sequences ϕ1,2 and ψ1,2 which are independent
of n but may depend on the order |p| of the derivative. Then, based on this
assumption, we must show that the derivatives of the next approximation
(n+1) (n+1) (n+1) (n+1) k
{(xj , uj , yj , vj )}j=0 satisfy the same estimates.
(n+1) (n+1)
In fact, we need to check these estimates only for xj and uj ; the
(n+1) (n+1)
analogous conclusion concerning yj and vj will follow from the sym-
metry of the problem.
The differentiation of (B.11) gives
j−1
(n+1) X
Dp x j = Dp A1 (µ)j x0 + Dp A1 (µ)j−s−1 f1 (x(n) (n) (n) (n)
s , us , ys , vs , µ) ,
s=0
j−1
(n+1) X
Dp u j = Dp A2 (µ)j u0 + Dp A2 (µ)j−s−1 f2 (x(n)
s , u (n) (n) (n)
s , y s , v s , µ) .
s=0
2 We ∂ p1 +p2
use a notation Dp = ∂(x0 ,u0 ,y 1 ,v 1 )p1 ∂µp2
(here p = (p1 , p2 )).
378 Appendix B
j−1
X
× λ−s
0 Dp0 f1 (x(n) (n) (n) (n)
s , us , ys , vs , µ) ,
s=0
(B.23)
X j−1
X
(n+1) 0 −s
kDp uj k ≤ const · (λ0 )j 1 + (λ )
p01 =p1 , p02 =0,...,p2 s=0
(n) (n) (n) (n)
× D p0 f2 (xs , us , ys , vs , µ) .
Now, in the same way as before, to prove the lemma we must check that
the estimates analogous to (B.18) and (B.19) hold for the derivatives Dp f1,2
for any p:
s−k s p2
kDp f1 (x(n) (n) (n) (n) 2s
s , us , ys , vs , µ)k ≤ [β1 (k − s)λ0 + β2 (s)γ0 λ0 ]k (B.24)
and
(p)
kDp f2 (x(n) (n) (n) (n) s p2
s , us , ys , vs , µ)k ≤ [β3 (s) + δϕ2 (s)]λ0 k (B.25)
where δ may be taken arbitrarily small by decreasing the size of the neigh-
borhood of the saddle fixed point under consideration and β1,2,3 are some se-
quences converging to zero; moreover, β3 is independent of the specific choice
(p) (p) (p)
of the estimators ϕ1,2 and ψ1,2 in (B.22), and β1,2 are independent of ϕ1
(p)
and ψ1 (nevertheless, β1,2,3 may depend on ϕ and ψ corresponding to the
derivatives of lower orders).
(n) (n) (n) (n)
By the chain rule, the derivatives Dp fi (xs , us , ys , vs , µ) are estimated
by the sum
X ∂ q1 +q2 +q3 fi
const · q1 ∂(y, v)q2 ∂µq3
(x(n) (n) (n) (n)
s , us , ys , vs , µ)
q1 ,q2 ,q3
∂(x, u)
(n) (n) (n) (n) (B.26)
× kDl1 (xs , us )k · · · kDlq1 (xs , us )k
(n) (n) (n) (n)
× kDlq1 +1 (ys , vs )k · · · kDlq1 +q2 (ys , vs )k
where q1,2,3 are nonnegative integers such that q1 + q2 + q3 ≤ p1 + p2 and
l’s are pairs of nonnegative integers such that l11 + · · · + lq1 +q2 ,1 = p1 and
l12 + · · · + lq1 +q2 ,2 + q3 = p2 .
Trajectories near a saddle fixed point 379
(n) (n)
By assumption, the estimates for the derivatives kDl us k and kDl vs k are
given by (B.22). Since ϕ1 and ψ1 are independent of j, there exists a constant
C independent of the specific choice of ϕ and ψ such that when (B.22) is
fulfilled
−(k−s) l2
kDl x(n) s l2
s k ≤ Cλ0 k , kDl ys(n) k ≤ Cγ0 k (B.27)
for all sufficiently large k.
Thus, the estimate (B.26) is rewritten as
X ∂ q1 +q2 +q3 fi
const · (x(n) , u(n) (n) (n)
s , ys , vs , µ)
q1 ,q2 ,q3
∂(x, u)q1 ∂(y, v)q2 ∂µq3 s
q (s−k) p2 −q3
× λq01 s γ02 k . (B.28)
as x → 0). Hence, the terms with q1 = 0 and q2 ≥ 1 in the estimate (B.28) for
f1 , and all terms with q1 = 0 in the estimate for f2 also fit (B.24) and (B.25),
respectively.
The case q1 = 0, q2 = 0 corresponds to the differentiation with respect to
µ alone (i.e. p1 = 0 and p2 = q3 ). Recall that the (r − 1)-th derivative with
q3
respect to µ may not exist, therefore we must estimate the derivatives ∂∂µqf31
only at q3 ≤ r − 2. These derivatives are smooth with respect to (x, u, y, v),
therefore we may write (using that f1 ≡ 0 at (y, v) = 0; see (B.2))
∂ q3 f1 ∂ ∂ q3 f1
≤ ky, vk · sup .
∂µq3 ∂(y, v) ∂µq3
Thus, the term under consideration is estimated exactly like other terms with
q1 = 0.
The last remaining terms to examine in (B.26) are (q1 = 1)
and
∂ ∂ q2 +q3 fi q (s−k)
· γ0 2 o(λs0 )k p2 −q3 .
∂u ∂(y, v)q2 ∂µq3
0
Note that fix → 0 as (x, u) → 0 (see (B.2)). Hence, both the terms above
q (s−k)
are estimated by γ02 o(λs0 )k p2 ; i.e. they fit (B.25) and, if q2 ≥ 1, they fit
(B.24).
It remains to consider the case q1 = 1, q2 = 0 for f1 . To satisfy (B.24) we
have to show that
∂ ∂ q3 f1
· λ−s
0
∂(x, u) ∂µq3
tends to zero as k − s → +∞, but this obviously follows from (B.16) because
f1 = f11 x + f12 y and both f1i vanish at (y, v) = 0 (see (B.2)).
(n) (n) (n) (n)
We have proved that the derivatives Dp fi (xs , us , ys , vs , µ) satisfy
(n) (n)
estimates (B.24) and (B.25). Note that for the derivatives of xs and ys we
used only estimates (B.27) which are independent of the choice of ϕ and ψ
in (B.22). Thus, the estimators β1,2 in (B.24) are independent of ϕ1 and ψ1
indeed. The only terms in (B.26) which might give a contribution in (B.25)
(p) (p)
dependent on ϕ1,2 and ψ1,2 are
0
kf2u k · kDp u(n)
s k and 0
kf2v k · kDp vs(n) k .
(p)
The first term here is estimated as δλs0 ϕ2 (s)k p2 where δ may be taken arbi-
trarily small. The second term is estimated as
(p)
k p2 ψ2 (k − s) · (kf21v
0
kkx(n) 0 (n)
s k + kf22v kkus k)
which gives, for sufficiently large k, the estimate o(λs0 )k p2 (see (B.16),(B.2))
(p) (p)
independently of the choice of ϕ1,2 and ψ1,2 . All this is in a complete agreement
with (B.25).
(n+1)
Now, the validity of estimates (B.22) for the next approximation {(xj ,
(n+1) (n+1) (n+1)
uj , yj , vj )}kj=0 follows from (B.24), (B.25) exactly in the same way
like the validity of (B.10) follows from (B.18), (B.19). The lemma is proved.
Remark. In the same way, slightly better estimates where o(λk0 ) and
o(γ0−k ) terms are replaced, respectively, by O((λ0 )k ) and O((γ 0 )−k ) in (B.6)
and (B.7) may be obtained for the functions ξ, η and their derivatives up to
the order (r − 2) in case the map is at least C3 -smooth (see Gonchenko and
Shilnikov [27]).
Bibliography
381
382 Bibliography
[11] Belitskii, G. R. [1961] “An algorithm for finding all vertices of convex
polyhedral sets,” J. SIAM 9(1), 72–88.
[13] Bendixson, J. [1901] “Sur les courbes definies par les equations differen-
tielles,” Acta Mathem. 24.
[22] Denjoy, A. [1932] “Sur les courbes définies par les équations différentielles
á la surface du tore,” J. Math. 17(IV), 333–375.
[35] Hirsch, M., Pugh, C. and Shub, M. [1977] Invariant Manifolds, Lecture
Notes in Mathematics, Vol. 583 (Springer-Verlag: New York).
[51] Petrowsky, I. [1934] “Uber das Verhalten der Integralkurven eines Systems
gewöhnlicher Differentialgleichungen in der Nähe eines singularen Punk-
tes. Matemat. Sbornik 41, 108–156.
[54] Poincaré, H. [1921] “Analyse des travans de Henri Poincaré faite par
luimeme,” Acta mathematica 38, 36–135.
[59] Shashkov, M. V. and Turaev, D. V. [1997] “On a proof of the global center
invariant manifolds,” to appear.
[76] Shilnikov, L. P. [1994] “Chua’s Circuit: Rigorous results and future prob-
lems,” Int. J. Bifurcation and Chaos 4(3), 489–519.
[77] Mira, C. [1997] “Chua’s Circuit and the qualitative theory of dynamical
systems,” Int. J. Bifurcation and Chaos 7(9), 1911–1916.
[78] Madan, R. N. [1993] Chua’s Circuit: A Paradigm for Chaos (World Sci-
entific: Singapore).
[79] Pivka, L., Wu, C. W. and Huang, A. [1996] “Lorenz equation and Chua’s
equation,” Int. J. Bifurcation and Chaos 6(12B), 2443–2489.
[80] Wu, C. W. and Chua, L. O. [1996], “On the generality of the unfolded
Chua’s Circuit,” Int. J. Bifurcation and Chaos 6(5), 801–832.
389
390 Index
theorem Birkhoff, 10
theorem on the leading manifold, 141
theorem on the non-leading manifold,
137
time-reverse, 5
topological classification, 56
topological conjugacy, 128
topological saddles, 64
topological type, 63, 133, 207
topologically conjugate fixed points, 133
topologically equivalent, 17, 59
trajectory, 7
trajectory equivalent, 18
trajectory of the Poincaré map, 114
trajectory Poisson-stable, 9
trajectory special, 17