Dynamic Systems Notes
Dynamic Systems Notes
Christian Kuehn
Faculty of Mathematics,
Technical University of Munich,
85748 Garching b. München, Germany
1
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
Preface
These notes are designed to give a concise introduction to dynamical sys-
tems with a view towards techniques for applications. Many monographs
and extended textbooks exist; see bibliography. The lectures are based
upon quite a number of sources [1, 2, 3, 4, 5, 6, 7, 8, 9, 10, 11, 12, 13, 16,
17, 18, 19, 20, 21, 22, 23, 24, 25, 26, 27, 28, 29, 30]. The goal here is to
extract the key ideas for the course “Dynamical Systems” (4 weekly hours)
held first at TU Munich by the author in the summer semesters 2017 and
2018.
The course considered in these lecture notes is the third part of a five-
course series:
Currently not all notes of the four courses are available but may be
updated regularly in the future.
Note carefully: The current version of the notes is a work in progress
and should be used with caution! Please report any errors or inaccuracies
you find to
ckuehn@ma.tum.de
2
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
Contents
1 Definitions & Examples 4
2 Stability 9
3 Stable/Unstable Manifolds 15
4 Equivalence 23
5 Poincaré Maps 27
6 Structural Stability 31
7 Hyperbolic Sets 34
8 Local Bifurcations 39
9 Normal Forms 49
10 Center Manifolds 53
11 Higher Codimension 59
12 Global Bifurcations 62
14 Melnikov’s Method 76
15 Chaos 81
16 One-Dimensional Maps 86
17 Renormalization 89
18 Attractors 93
19 Invariant Measures 99
20 Ergodicity 103
21 Mixing 108
22 Entropy 113
3
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
1 Definitions & Examples
Definition 1.1. A dynamical system is a triplet (X , T , φt ), where X is
the phase space (or state space), T is the time set (or time domain),
and φt : X → X for t ∈ T is the family of evolution operators satisfying
(D1) φ0 = Id,
dx
= x′ = Ax, x = x(t), x(0) = x0
dt
by setting φt (x0 ) = x(t) = etA x0 . A very simple nontrivial example in R2
arises if A is a diagonal matrix with Aii 6= 0, which has a simple visualiza-
tion in the phase plane/space; see Figure 1(b).
g(x)
x2
(a) (b) (c)
φt
x1 x2 x1 x0 x
X
Figure 1: (a) General dynamical system on X . (b) Linear ODE (1.1) with
diagonal matrix A ∈ R2×2 and A11 < 0, A22 > 0. (c) Iterated cubic
map (1.2) including cob-web diagram (dotted lines).
x3 =: g(x), g : R → R, (1.2)
4
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
and define
x 0 1
1
x
2
Figure 2: (a) Sketch of the logistic map (1.4) g(x; p) =: g(x) for p > 2. (b)
Logistic map; draw in I as discussed in the text (exercise!).
5
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
so the inverse is multi-valued. Multi-valued (or set-valued) dynamical
systems will not be discussed here.
Recall the following classical theorem from a first ODE course:
Theorem 1.7. (Existence, uniqueness and smooth dependence) Let f :
Rd → Rd be smooth and consider the ODEs
Then there exists a unique map (t, x0 ) 7→ φt (x0 ), smooth in (t, x0 ) such that
(T1) φ0 (x0 ) = x0 ,
(T2) there exists a (maximal) interval T = (−t1 , t2 ), t1,2 = t1,2 (x0 ) > 0
such that for all t ∈ T , φt (x0 ) = x(t), i.e., the ODE (1.5) is satisfied.
Recall that the proof proceeds via re-writing the ODEs in integral form,
defining a suitable norm, and using the Banach fixed-point theorem; see [2,
10].
Theorem 1.8. (Banach Fixed Point Theorem) Let (X , d) be a non-
empty complete metric space and assume g : X → X is a contraction
mapping
6
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
Definition 1.10. Suppose (Rd , R, φt ) is a dynamical system. If φt (x) is
smooth in (t, x), then we say (X , T , φt ), is a smooth dynamical system.
Remark : Similar definitions can be stated for C k (i.e., k-times differentiable,
k ≥ 1), or just continuous C 0 dynamical systems. Usually we shall assume that
the dynamical system is at least C 0 ; however, many of our examples are actually
even smooth.
Our key goal is to analyze and classify possible dynamics.
Example 1.13. Let p > 0 be a parameter. Consider the planar van der
Pol equation
3
x′ = y − x3 + x,
(1.6)
y ′ = −px,
An equilibrium of (1.6) must satisfy y ′ = 0 so x = 0, which implies using
y
2
3
1 x
−1
− 32
the first equation for x′ that also y = 0. Therefore, (x∗ , y∗ ) = (0, 0) is the
unique equilibrium point; see Figure 3.
7
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
Definition 1.14. If we consider an ODE x′ = f (x), then the curves
ω = {. . . , ω−2 , ω−1 , ω0 , ω1 , ω2 , . . .}
or just written as
ω = · · · ω−2 ω−1 ω0 ω1 ω2 · · · .
Define the (one-step, left-)shift map g : Σ2 → Σ2 by
i.e., we shift each entry one position to the left. Clearly, we have that
are the only two fixed points. However, there are a lot more invariant sets,
which are not fixed points.
for some fixed T0 > 0 and all t ∈ T ; see also Figure 4(a). We call the
minimal such T0 > 0 the period of γ.
8
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
x2
(a) γ (b)
R2 = X
x2 x1
S
x1
Figure 4: (a) Sketch of a periodic orbit for a planar ODE. (b) Sketch of a
positively invariant set S.
2 Stability
Definition 2.1. Let x∗ be a steady state for (X , T , φt ). Then x∗ is called
(Lyapunov) stable if for any given neighbourhood U = U (x∗ ) there exists
another neighbourhood V(x∗ ) = V ⊆ U such that
9
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
From now on, we shall always assume that X has a metric. Since most
examples are related to Rd with Euclidean norm | · |, we just state this case
in our results; generalizations are presented when necessary. For example,
asymptotic stability means
d2 x
+ x = x′′ + x = 0,
dt2
which can be re-written as a first-order system, setting x′ = y,
′
x 0 1 x
′ = . (2.1)
y −1 0 y
x2 x2 x2
(a) (b) (c)
x1 x1 x1
10
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
Definition 2.3. Let x∗ be an equilibrium for (2.2) and U be a neighbour-
hood of x∗ . A function L : U → R is called a Lyapunov function if
(D1) L is continuous,
The Lyapunov function is called strict if equality in (2.3) never occurs; see
Figure 6 for an illustration of the definition.
x1
x2
{x ∈ U : L(x) ≤ δ}
11
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
We argue by contradiction. Therefore, for every n ∈ N, there exists xn ∈
S1/n such that |xn − x∗ | ≥ δ. Since S1/n is connected, we can take xn such
that |xn − x∗ | = δ. Since the sphere is compact, and all xn lie in ∂B(x∗ ; δ),
we extract a convergent subsequence
B(x1 ; r) ⊆ U
of radius r > 0 such that φt1 +ρ (x0 ) ∈ B(x1 ; r) \ Sε for sufficiently small
ρ > 0. However, this implies
Proof. For every x with φt (x) ∈ U (x∗ ) for all t ≥ 0, the limit
12
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
exists by monotonicity. If y ∈ ω(x), it follows that
L′ (φt (x)) = ∇L(φt (x)) · φ′t (x) = ∇L(φt (x)) · f (φt (x)) ≤ 0.
x′ = y = ∂y H(x, y),
′ (2.6)
y = −∇V (x) = −∂x H(x, y),
with equilibrium x∗ = 0.
13
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
(T1) If all eigenvalues λj = aj + ibj of A satisfy Re(λj ) = aj < 0, then 0
is locally asymptotically stable.
y ′ = Jy, y ∈ Rd .
Solving the system explicitly using the matrix exponential φt (x) = etJ x
we have to consider all possible Jordan block structures. For (T1), any
diagonal block has solutions
Definition 2.11. Consider the linear system (2.7) for d = 2 and λ1,2 =
a1,2 + ib1,2 . Then x∗ is called a
• node if b1,2 = 0, a1 a2 > 0;
x 7→ Ax x ∈ Rd , x(0) = x0 .
14
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
3 Stable/Unstable Manifolds
Before we start, a brief introduction (and/or refresher) on elementary def-
initions from differential geometry/topology.
Definition 3.1. A d-dimensional (topological) manifold M is a second-
countable Hausdorff topological space such that each point x ∈ M has a
neighbourhood, which is locally homeomorphic to an open set in Rd .
One may check that topological manifolds have an open covering {Uα },
for Uα ⊆ M of charts, i.e., homeomorphisms hα : Uα → Vα ⊂ Rd , where
each Vα is an open subset in Rd . The collection {(Uα , hα )} of charts is
called an atlas; see Figure 7(a).
x2
(a) (b1) (b2)
Uα Uβ
x1
hα hβ
(b3) (b4)
Vα Vβ
hα ◦ h−1
β
Figure 7: (a) Sketch of a manifold, charts, and transition maps. (b) Ex-
amples of two-dimensional manifolds: Euclidean space R2 , a square in R2 ,
a torus, and a sphere.
hα ◦ h−1 k
β : hβ (Uα ∩ Uβ ) → hα (Uα ∩ Uβ ) is in C , (3.1)
and the atlas is maximal, i.e., if there exists another chart, which is com-
patible in sense of (3.1) with all other charts, then it is included in the
atlas; see also Figure 7(a).
Remark : One may prove that it suffices to exhibit some C k -atlas, where the
smoothness condition for the transition maps is satisfied, to show that the manifold
is C k . Surprising fact: Any topological manifold for d ≤ 3 can be given the
structure of a smooth manifold but this fails for d ≥ 4.
15
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
(d = 1 is just the circle), as well as unit sphere Sd = {x ∈ Rd : |x| = 1}. All
these manifolds are C k (they are even smooth k = ∞, and analytic k = ω)
and can be locally written as graphs of smooth functions (exercise!); see
Figure 7(b).
One naturally obtains the space C k (M1 , M2 ) for maps between smooth
manifolds by postulating that f ∈ C k if it is a smooth map upon application
of coordinate charts.
Tx M
S2
16
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
We now start the main dynamical constructions. As before, we mostly
work in this section with dynamical systems generated by the ODE (2.2),
i.e.,
x′ = f (x), x ∈ Rd . (3.2)
X ′ = Df (0)X = AX, X ∈ Rd
Ps : Rd → E s and Pu : Rd → E u .
However, the nonlinear terms O(|x|2 ) may locally deform the geometry of
the linear spaces near x∗ ; see Figure 9.
17
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
x2 x2
(a) (b)
u
Wloc
Pu
x1 x1
s
Wloc
Eu Eu
Es Es
18
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
(a) (b)
h(x)
Es Es
x
Bεs
Eu Eu
s (0) as a graph with a sufficiently small
Figure 10: (a) Idea to construct Wloc
(ball) domain Bε = B(0; ε) within the stable eigenspace E s = E s (0). (b)
s
Define P (t) = Ps for t > 0 and P (t) = −Pu for t ≤ 0, then (3.7) is re-written
as
Z ∞
tA
x(t) = K(x)(t), K(x)(t) = e xs + e(t−r)A P (t − r)R(x(r)) dr. (3.8)
0
19
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
Since we work in some U , assume |x(t)| ≤ δ for some δ > 0. Observe that
g satisfies
|R(x(t)) − R(y(t))| ≤ ε|x(t) − y(t)| (3.9)
and ε > 0 can be made arbitrarily small if δ > 0 is made small. In
backward time solutions of the linearized system decay exponentially in
E u , and solutions decay in forward time in E s so that the definition of P (t)
implies
ke(t−r)A P (t − r)k ≤ Ce−κ|t−r| (3.10)
for some constant κ > 0, which can be bounded using the spectrum of A
and some constant C > 0. Using (3.9) and (3.10) in (3.8) yields
Z ∞
kK(x) − K(y)k ≤ sup e(t−r)A P (t − r)(R(x(r)) − R(y(r))) dr
t≥0 0
Z ∞
≤ C sup e−κ|t−r| |R(x(r)) − R(y(r))| dr
t≥0 0
Z ∞
2Cε
≤ Cεkx − yk sup e−κ|t−r| dr = kx − yk.
t≥0 0 κ
For ε < κ/(2C), K is a contraction so the Banach Fixed Point Theorem 1.8
s (0) using (3.4).
gives the existence of x(t) and hence also Wloc
Step 2: The next question is differentiability. We know the solution
ψ(t, xs ) is continuous in t. To show continuity ψ ∈ C([0, ∞) × U , Rd ), we
use the triangle inequality
|ψ(t, a) − ψ(s, b)| ≤ |ψ(t, a) − ψ(t, b)| + |ψ(t, b) − ψ(s, b)|, (3.11)
where the second term can be made small by continuity in time. We claim
that the first term is also small. Indeed, note that
20
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
which is continuous since R and P are. The other derivative Da ψ = η, if it
exists, must satisfy the fixed-point equation
Z ∞
tA
η(t, xs ) = e Ps + e(t−r)A P (t − r)η(t, xs ) · ∇R(ψ(r, xs )) dr, (3.12)
0
Dxs h(a)|a=0 = Pu Ps = 0
21
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
The second major proof technique (Hadamard-Perron method) for
invariant manifold theorems will be briefly illustrated in the context of
Theorem 3.9 for x∗ = 0. Consider the set
Bεs := {y ∈ E s (0) : |y| ≤ ε}.
Let L(L, ε) be the space of Lipschitz-continuous maps h : Bεs → E u (0) with
Lipschitz constant L and h(0) = 0, which is a Banach space under the
supremum norm. The graph transform is constructed as operator
G : L(L, ε) → L(L, ε), G(ϕ) = ϕ̃.
It is defined by considering the maping
ϕ 7→ g −1 (graph(ϕ)). (3.13)
One then proves (nontrivial!) that the resulting set in (3.13) is again the
graph of a Lipschitz map ϕ̃ for sufficiently small ε > 0 over the correct
domain Bεs ; see Figure 10(b). Then one proves (again nontrivial!) that G
is a contraction for suitable choices of (L, ε) and obtains the existence of
s (0) as a fixed point of G.
Wloc
Example 3.10. Consider the (unforced) Duffing equation/oscillator
x′′ + δx′ − x + px3 = 0, x ∈ R,
where p, δ ∈ R are parameters and δ ≥ 0. Re-writing it as a first-order
system gives
x′ = y,
(3.14)
y ′ = −δy + x − px3 .
Obviously, (x, y) = (0, 0) is an equilibrium. If δ = 0, then one checks
dim E s (0) = 1 = dim E u (0). Therefore, there exist one-dimensional stable
and unstable manifolds by Theorem 3.8.
Example 3.11. Consider the linear map (sometimes called Arnold’s cat
map) defined by
21
g(x) = Ax, A= , x ∈ R2 . (3.15)
11
One checks (exercise!) that g preserves the integer lattice Z2 ⊂ R2 and
induces a map on the torus X = T2 = R2 /Z2 ; see Figure 11. The origin
(0, 0) is a hyperbolic saddle fixed point (exercise!) and hence Theorem 3.8
applies to this point. The map g : T2 → T2 is the classical example for a
hyperbolic toral automorphism, i.e., it does not only have a hyperbolic
splitting at (0, 0) but near every point in X it has well-defined contracting
and expanding directions.
22
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
(a) x2 (b) x2
2 2
1 1
0 x1 0 x1
0 1 2 3 0 1 2 3
Figure 11: (a) Mapping of the fundamental domain [0, 1]2 (grey) via (3.15)
to a polygon (black). (b) Identifying the resulting object back onto the
fundamental domain.
4 Equivalence
Definition 4.1. A dynamical system (X , T , φt ) is called topologically
equivalent to a system (X , T , ψt ) if there exists a homeomorphism h : X →
X mapping orbits from the first system to the second system preserving
the direction of time. If the time parametrization is also preserved, the
systems are called conjugate.
x 7→ f (x), x ∈ Rd , y 7→ g(y), y ∈ Rd ,
Proof. We only show that topological conjugacy implies (4.1); the converse
is an easy exercise. Topological equivalence implies that there exists a
homeomorphism h : Rd → Rd such that
Therefore, h(f (x)) = g(y) = g(h(x)) and since h is invertible, the result
follows.
23
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
Proposition 4.3. Consider two continuous-time flows φt and ψt generated
by
x′ = f (x), x ∈ Rd and y ′ = g(y), y ∈ Rd
with f, g ∈ C 1 . Suppose y = h(x) for some C 1 -diffeomorphism h : Rd →
Rd . Then φt , ψt are C 1 -conjugate, and
x2 x2
x1 x1
Figure 12: Generic stable node for a nonlinear system conjugated via h
to a node for a linear system. For generic matrices with spectrum in the
left-half complex plane, there exists a weakest contracting eigendirection
(exercise!); see also Definition 6.9.
24
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
and then comment on the proof of the Hartman-Grobman Theorem; see
also Figure 12.
Theorem 4.6. Suppose 0 is a hyperbolic fixed point of (4.3). Then there
exists a unique homeomorphism h(x) = x + H(x), H(0) = 0, H bounded,
such that in a neighbourhood U = U (0) ⊂ Rd we have
h ◦ A = g ◦ h, (4.4)
α := max(kA−1
u k, kAs k) < 1, |R(x) − R(y)| ≤ ε|x − y|, (4.5)
where we select ε < (1 − α)/2 and the inequality for R holds on U but we
cut off R with a smooth cut-off function outside U so that it holds globally.
The result (4.4) is equivalent to the functional equation
Consider the Banach space C(Rd , Rd ) with the supremum norm. Define
two invertible linear operators
where K preserves the norm, while A has expanding and contracting di-
rections. Setting L := K − A, it follows that (4.6) can be viewed as a
fixed-point problem for L
so that L−1 −1 −1 −1 −1
u = −(Id − Au Ku ) Au , where the existence of Au holds by
−1
assumption and (Id − Au Ku ) is invertible using the Neumann series
∞
X
(Id − A−1
u Ku )
−1
= (A−1
u Ku )
k
k=0
25
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
as kA−1 −1
u Ku k ≤ α < 1. Furthermore, we have kLu k ≤ 1/(1 − α) using the
geometric series. The same argument applies to Ls so
2
L−1 = (Ku − Au )−1 ⊕ (Ks − As )−1 , kL−1 k ≤ .
1−α
So we may now try to solve H(x) = L−1 R(x + H(x)) for H. Since the goal
is to apply Banach Fixed Point Theorem 1.8, we estimate
2
kL−1 R(Id + H1 ) − L−1 R(Id + H2 )k ≤ kR(Id + H1 ) − R(Id + H2 )k
1−α
2ε
≤ kH1 − H2 k.
1−α
This implies we have a contraction by (4.5) and H exists and is continuous.
Step 2: It remains to prove that h = Id + H is invertible. First, note
that we can construct a map h̃ such that
This follows by exchanging the roles of A, g above and repeating the same
line of argument. Therefore, using (4.4) and (4.7) we find
A ◦ h̃ ◦ h = h̃ ◦ g ◦ h = h̃ ◦ h ◦ A. (4.8)
φ1 (x) = eA x + R(x)
26
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
for some remainder R. Upon localizing the vector field near x = 0 with a
cut-off, one checks that R is indeed small so that Theorem 4.6 applies so
we get a homeomorphism h with
h ◦ eA = φ1 ◦ h. (4.9)
hs := φs ◦ h ◦ e−sA
and note that hs also satisfies (4.9) for s = 1 (exercise). Using the unique-
ness part of Theorem 4.6 one obtains hs ≡ h completing the construction.
5 Poincaré Maps
Consider the ODE
x′ = f (x), x ∈ Rd , f ∈ C 1 , (5.1)
with flow φt and suppose Γ = Γ(t) is a periodic orbit. The idea of a Poincaré
map is to record returns to a manifold transversal to Γ; see Figure 13.
P (x)
x
y
Σ := {x ∈ Rd : (x − y) · f (y) = 0} (5.2)
27
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
In summary, the Poincaré map P (x) := φτ (x) (x) is locally well-defined
and C 1 on B(y; δ) ∩ Σ.
Proof. The idea is to get τ via the implicit function theorem. Define the
function
F (t, x) := (φt (x) − y) · f (y) (5.3)
which is in C 1 since φ ∈ C 1 . Furthermore, F (T, y) = 0 by periodicity of Γ.
Differentiating (5.3) in t, we find
Proof. Suppose there are two points y, ỹ, two sections Σ, Σ̃, and coordinates
x, x̃ with Poincaré maps P, P̃ . Then, using the construction of P and P̃ ,
one finds that there exists a diffeomorphism h(x) = x̃ such that P, P̃ are
locally conjugate so
P = h−1 ◦ P̃ ◦ h.
Differentiating and evaluating at x = 0 gives
28
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
Example 5.3. Consider the planar system with parameter p ∈ R given by
x′ = −y + x(p − x2 − y 2 ),
y ′ = x + y(p − x2 − y 2 ),
which turns out to have a limit cycle, where the single multiplier µ can be
calculated exactly (exercise! Hint: polar coordinates). However, in general
cases, this is not possible analytically.
To analyze the stability of a periodic orbit Γ, one may also link this to
classical ODE theory. Indeed, write
and take local coordinates (x1 , . . . , xd−1 , z), xd =: z ∈ R, such that the
Poincaré map P (x) = φτ (x) (x) is given solely in the first (d−1) coordinates.
We write φt (x) = φ(t, x) for clarity and differentiate
29
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
We claim that Dx φ(T, y) = M (T ). Indeed, we have Dx φ(0, y) = Id and we
calculate
(Dx φ(t, y))′ = Dx φ(t, y)f (y)
interchanging derivatives wrt t and x. Therefore, equation (5.5) implies
using (5.4) that
Dx P (y) v1
M (T ) = (5.6)
v2⊤ m
where v1,2 ∈ Rd−1 are some vectors and m ∈ R. We claim that v1 = 0 and
m = 1, which follows if we can prove M (T )f (y) = f (y), i.e., µ = 1 is an
eigenvalue of the monodromy matrix with eigenvector aligned with tangent
vector to the periodic orbit Γ(t) as easily drawn into Figure 13 (exercise!).
We have
Γ′′ (t) = f (Γ(t))′ = A(t)Γ′ (t)
so Γ′ (t) solves u′ = A(t)u. Therefore, we have
Then consider the flow φt on M defined by (x′ , t′ ) = (0, 1) (also called the
suspension). The time-T leaves {t = 0} ⊂ M invariant and induces the
map g; see Figure 14.
30
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
(x0 , T )
t=T
t=0
(x0 , 0) (g(x0 ), 0)
Figure 14: Illustration of the suspension flow construction.
6 Structural Stability
It would be most elegant to analyze all dynamical systems of a certain class
at once, just restricting to those displaying typical dynamics. Consider a
compact d-dimensional set K ⊂ Rd .
31
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
Definition 6.2. Let f ∈ C 1 (M) be a vector field. Then f is called struc-
turally stable if there exists ε > 0 such that for all vector fields g ∈ C 1 (M)
with
kf − gkC 1 < ε,
g is topologically equivalent to f . If the f is not structurally stable, it is
called structurally unstable.
Example 6.3. Consider the unit torus T2 = R2 /Z2 with vector field f
defined by
x ′ = ω1 ,
(6.2)
y ′ = ω2 .
for fixed ω1,2 ∈ R. Then f is structurally unstable since for ω1 /ω2 ∈ Q all
orbits are periodic while for ω1 /ω2 ∈ R \ Q, all orbits are aperiodic, i.e.,
are never closed.
φt (U ) ∩ U =
6 ∅.
The set of such points is called the nonwandering set NW(φ); see Fig-
ure 15(a). Other points are called wandering.
(a) φt (b)
φ(U )
M
Figure 15: (a) Sketch of the idea in the definition of a nonwandering point.
(b) Illustration of one aspect of Peixoto’s Theorem: a saddle-to-saddle
connection (grey) is structurally unstable.
Example 6.5. Consider the flow φ generated by (6.2), then we easily check
(exercise!) that NW(φ) = T2 = X .
32
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
Theorem 6.6 (Peixoto’s Theorem). A C 1 -vector field f on a compact two-
dimensional differentiable manifold M is structurally stable if and only if
(T1) the number of periodic orbits and equilibria is finite, and each is hy-
perbolic;
We shall not prove Theorem 6.6 but see Figure 15(b). Peixoto’s The-
orem for d ≥ 3 is false. We only have a uni-directional statement under
similar assumptions.
(D1) the number of periodic orbits and equilibria is finite, and each is
hyperbolic;
33
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
7 Hyperbolic Sets
In this section, we prove, why hyperbolicity is a such a key property to
expect robustness of dynamics and structural stability. Let M be a C 1
Riemannian manifold, i.e., there exists a Riemannian metric differ-
entiable family of symmetric bilinear forms
h·, ·ix : Tx M × Tx M → R
p
where we set |v| := hv, vi. For a curve γ : [a, b] → M, γ ∈ C 1 , we define
its length by
Z b
|γ ′ (s)| ds
a
which induces a metric d(x, y) on M by taking the infimum of the length
over all curves connecting x = γ(a) and y = γ(b). Next, define a discrete-
time dynamical system via a diffeomorphism
g : M → M. (7.1)
Dgx (v)
g(x)
M
34
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
(D3) (“unstable”) |Dgx−j v u | ≤ Cλj |v u | for every v u ∈ E u (x) and j ≥ 0;
d0 (Df, Dg) ≤ ε1
d0 (ϕ ◦ h, f ◦ ϕ) ≤ δ
35
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
U M
f
Λ
g
ϕ
∃ψ s.t. ϕ ≈ ψ
Df ≈ Dg
h
Y Y
Figure 17: Illustration of Theorem 7.5.
ψ(y) = f ◦ ψ(y)
x ∈ Λε := {x ∈ M : d(x, Λ) < ε}
36
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
.. x3
(a) x0 . (b)
g
g
g x2
g
x1
x2
Λ
g(x0 ) g(x1 )
g
δ
g
d(g(x0 ), x1 ) ≤ δ d(g(x1 ), x2 ) ≤ δ x x1
x0
h : Y → Y, h(xk ) = xk+1
37
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
The previous results already hint at the fact that hyperbolic sets should
obey various forms of “persistence” or “robustness” to perturbations.
Theorem 7.9 (Persistence of Hyperbolicity). Let Λ be a hyperbolic set.
Then the following statements hold:
(T1) There is an open set U (Λ) containing Λ and ε1 > 0 such that if
K ⊂ U (Λ) is a compact invariant set of a diffeomorphism f : M → M
with
d1 (f, g) < ε1 , (7.2)
then K is a hyperbolic set.
(T2) For every open set V containing Λ and every ε > 0, there exists δ > 0
such that for every f : M → M with d1 (f, g) < δ, there exists a
hyperbolic set K ⊂ V of f and a homeomorphism ψ : K → Λ such
that
ψ ◦ f |K = g|Λ ◦ ψ
and d0 (ψ, Id) < ε.
Roughly speaking, (T1) say that hyperbolic sets persist under perturba-
tions of the map, while (T2) means that the dynamics on the original and
the perturbed hyperbolic set is topologically conjugate; cf. the Hartman-
Grobman Theorem.
|Dgx v| ≤ λ|v| for v ∈ Kαs (x) and |Dgx−1 v| ≤ λ|v| for v ∈ Kαu (x)
38
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
for all x ∈ U (Λ). The same contraction/expansion statement also holds,
upon potentially increasing λ but still λ ∈ (0, 1), for f since it is C 1 -close
to g by (7.2). Now one can define the subspaces
\
E s (x) = D−k K s (f k (x)),
f k (x) α
k≥0
\
u
E (x) = Dkf −k (x) Kαu (f −k (x)),
k≥0
ψ ◦ g|Λ = f ◦ ψ.
ψ̃ ◦ f |K = g|Λ ◦ ψ
Proof. Openness follows directly from applying (T1) of Theorem 7.9, while
structural stability follows from (T2).
8 Local Bifurcations
Consider parameters p and the ODE
39
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
Example 8.2. Consider p = 1 = d and the ODE
x′ = px − x3 = x(p − x2 ). (8.2)
Sp≤0
Sp>0
p p
Definition 8.3. The parametrized ODEs (8.1) and (8.3) are called (fiber)
topologically equivalent if
(D1) there exists a homeomorphism ξ : Rm → Rm , ξ(p) = q;
(D2) for fixed ξ(p) = q, the ODEs are topologically equivalent.
Fix a parameter p0 and let Sp0 be the maximally connected subset in
parameter space containing p0 with topologically equivalent phase portraits
and call Sp0 a stratum; see Figure 19(a).
Definition 8.4. A bifurcation diagram is a stratification of the param-
eter space (p ∈ Rm ) induced by topological equivalence, together with a
representative phase portrait for each stratum.
40
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
Definition 8.5. An unfolding is a parametrized family of dynamical sys-
tems containing a bifurcation in a persistent way. The minimal number of
parameters for an unfolding is called its codimension.
Example 8.6. Consider m = 1 = d and the ODE
x′ = p + x2 + O(x3 ) (8.6)
x′ = p + x2 .
41
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
using the implicit function theorem since ∂p g(0, 0) = 1. This implies there
√
exist equilibria y± (p) of (8.7) close to x± (p) = ± p; see Figure 19(b). To
construct the required homeomorphism hp : R → R, define hp (x) = x for
p ≥ 0 and for p < 0 use a linear transformation
!
hp (x) = a(p) + b(p)x, hp (x± (p)) = y± (p),
where a, b can be solved for implicitly using the last two conditions/equations.
Therefore, we have the required (fiber) topological equivalence of Defini-
tion 8.3.
x′ = f (x, p), x ∈ R, p ∈ R,
is locally near (x, p) = (0, 0) topologically equivalent to one of the two fol-
lowing ODEs
y′ = q ± y2, (8.8)
which are also called the topological norm form of the fold bifurcation.
p q q
42
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
Step 1: Consider a linear phase space coordinate change y = x+δ(p) =
x + δ, which yields
f1′ (0)
δ(p) = − p + O(p2 ).
2f2 (0)
for some smooth function µ. Since q(0) = 0, q ′ (0) = f0′ (0) = ∂p f (0, 0) 6= 0
by (A2), the Inverse Function Theorem yields a unique smooth inverse
p = p(q) with p(0) = 0. Therefore, (8.9) can now be written as
43
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
p = 0 = x but wlog we may always shift coordinates in parameter and
phase space). Hence, the main bifurcation condition is
∂x f (0, 0) = 0
∂xx f (0, 0) 6= 0
∂p f (0, 0) 6= 0,
guarantees (exercise!) that the eigenvalue λ = λ(p) passes through the ori-
gin in C at non-zero speed, which is a so-called transversality condition;
see also Figure 21(a) for a transversality condition in another context. In
summary, one refers to (A2) as genericity conditions.
x′ = f (x, p), x ∈ R2 , p ∈ R,
44
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
y2 y2
(a) Im (b1) (b2)
C
Re p p
p<0 p>0
y1 y1
p=0
which are also called the topological norm form of the Hopf bifurca-
tion.
The Hopf and fold bifurcations are both of codimension one. Regarding
the requirement that the setting is generic, the correct transversality condi-
tion µ′ (0) 6= 0 for the Hopf bifurcation is easy to guess. The non-degeneracy
condition is a requirement, which guarantees non-vanishing cubic terms; see
Example 8.11. The transformation to normal form is discussed in Section 9.
Example 8.11. We analyze the dynamics of (8.11) denoting the sign of the
cubic term by l1 ∈ R \ {0}; it is called the first Lyapunov coefficient.
The first Lyapunov coefficient turns out to be actually computable from
the original vector field only [7] and the non-degeneracy condition for the
Hopf bifurcation is l1 6= 0. Two coordinate changes are useful to analyze
the Hopf normal form dynamics. First, consider the complexification
z = y1 + iy2 ∈ C and observe
z ′ = y1′ + iy2′ = q(y1 + iy2 ) + i(y1 + iy2 ) + l1 (y1 + iy2 )(y12 + y22 )
= (q + i)z + l1 z|z|2 ,
45
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
gives us the polar form
r′ = r(q + l1 r2 ),
(8.12)
ϕ′ = 1.
From (8.12), we obtain the two bifurcation diagrams in Figure 21(b):
For a Hopf bifurcation not in normal form, the first Lyapunov coefficient is
computable using derivatives of the vector field up to and including order
three. The non-degeneracy condition is then l1 6= 0.
46
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
with genericity conditions
1 1
(∂xx g(0, 0))2 + ∂xxx g(0, 0) 6= 0 and ∂xp g(0, 0) 6= 0,
2 3
i.e., under these conditions, the map (8.13) is locally topologically conjugate
to (8.15).
where g1 (p) = −(1 + κ(p)) for some smooth function κ with κ(0) = 0 since
µ = −1. Using the transversality assumption
the inverse function theorem implies locally invertibility and we may con-
sider a new parameter q = κ(p). This yields map
x = y + δy 2 , δ = δ(q), (8.17)
y 7→ µy + (a + δµ − δµ2 )y 2
+(b + 2δa − 2δµ(δµ + a) + 2δ 2 µ3 )y 3 + O(y 4 ).
a(q)
δ(q) = µ2 (0) − µ(0) = 2 6= 0.
µ2 (q) − µ(q)
47
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
This implies we obtain a map
2a2
y 7→ µy + b + 2 y 3 + O(y 4 ) = −(1 + q)y + c(q)y 3 + O(y 4 ).
µ −µ
We can compute the smooth function c at q = 0
2
2 1 1
c(0) = a (0) + b(0) = ∂xx g(0, 0) + ∂xxx g(0, 0) 6= 0
2 6
by the non-degeneracy assumption. This implies we may use the scaling
p
ỹ = y |c(q)|
and dropping the tildes yields the result up to higher-order terms, which
can be dropped using the same technique as in Lemma 8.7.
Example 8.14. We analyze the normal form (8.15) with the positive sign
for the cubic term. Fixed points satisfy
This map y 7→ f 2 (y, q) has the trivial fixed point y∗ = 0 but also two
non-trivial fixed points for q > 0 sufficiently small
≈ √
y± = ± q + higher-order terms.
√
In fact, one checks explicitly via (8.19) that the points y± = ± 2 are stable
and constitute a periodic orbit of period two as shown in Figure 22(c).
A similar, yet technically more involved treatment, can be given for the
two-dimensional case
xk+1 = g(xk , p) xk ∈ R2 , p ∈ R,
where a complex conjugate pair of multipliers µ1,2 (q) may pass through
the unit circle |µ1,2 (0)| = 1 (under the assumption µ1 6= µ2 ). In this case,
the bifurcation is called a Neimark-Sacker bifurcation. The Neimark-
Sacker bifurcation leads to a rather complicated invariant curve, which
generically bifurcates at q = 0.
48
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
y g(y) g(y)
(a) (b) (c)
q y y
det(A(p) − Id) = 0
Then one may work out (exercise!) the algebraic conditions for a multiplier
crossing and try to check, when genericity conditions hold.
9 Normal Forms
We now discuss a more general complementary approach to previous tech-
niques to compute a normal form for a smooth ODE x′ = f (x), x ∈ Rd .
Definition 9.1. Let Hk denote the space of d-component vector-valued
homogeneous polynomials of degree k.
49
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
Assume x = 0 is an equilibrium and Taylor-expand the ODE
Lh(k) = f (k) .
50
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
Lemma 9.3. The complement Rk to Hk is orthogonal in a scalar product
h·, ·ik if it is chosen as
ker(L∗ ) = {h ∈ Hk : L∗ u = 0}
where L∗ is the adjoint of L, i.e., hLu, vik = hu, L∗ vik . If there are no
generalized eigenvectors for the zero eigenvalue of L we have that
ker(L) = {h ∈ Hk : Lu = 0}
Proof. Recall that the Fredholm Alternative Theorem says in our con-
text that either Lh = b has a solution b ∈ range(L) or L∗ h∗ = 0 has a solu-
tion h∗ ∈ Hk with hh∗ , bik 6= 0, which proves the first part. For the second
part, no generalized eigenvectors imply that ker(L) ⊕ range(L) = Hk .
Proof. We write (y1 , y2 ) = (x, y) throughout the proof for notational sim-
plicity. Consider quadratic terms (k = 2), fix an (ordered) polynomial
basis {x2 , xy, y 2 } of H2 . View a vector field h = h(2) as a map h : R2 →
T R2 ≃ R2 with basis coordinates ∂x , ∂y ; for example, write (2xy, x2 )⊤ as
(2xy)∂x + (x2 )∂y . Wlog (upon scaling time if necessary) we have
x x
Lh = DA h − (Dh)A
y y
0 −1 h1 ∂x h1 ∂y h1 −y
= − .
1 0 h2 ∂x h2 ∂y h2 x
51
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
for the first component, and for the second component
0 −x2 0 −xy 0 −y 2
L = , L = , L = .
x2 2xy xy y 2 − x2 y2 −2xy
we find
0 −1 0 −1 0 0
+2
0 −2 0 −1 0
0 +1 0 0 0 −1
M (2) = .
+1
0 0 0 −1 0
0 +1 0 +2 0 −2
0 0 +1 0 +1 0
Since det(M (2) ) = −9 6= 0, all quadratic terms can be eliminated. The
procedure can now be carried out for k = 3. We use the ordering of
coordinates in R8
52
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
which does not match (9.2). Using the second part of Lemma 9.3, we find
that ker(M (3) ) is spanned by
The last proof showed that the Poincaré normal form is not unique
due to a choice of basis. We also verified the algebra necessary for The-
orem 8.10, yet Poincaré normal forms do not yield the first Lyapunov
coefficient. The method of normal forms can be immediately applied to
parameter-dependent systems written in the form
x′ = f (x, p),
p′ = 0.
The algebraic structure of (9.3) can be derived from the Poincaré normal
form making the ansatz
′ 01
x = x + f (2) (x) + f (3) (x) + · · · ,
00
10 Center Manifolds
So far, we have crucially assumed to work in the minimal dimension for
the smooth ODE
53
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
where A = Df (0) and x∗ = 0 is an equilibrium. Suppose the spectrum of
A has a center eigenspace
Es Wc
W c (0, 0)
Ec x1 x
Figure 23: (a) Basic idea of center manifold W c . (b) Illustration of center
manifolds for (10.2) showing lack of uniqueness and smoothness. Note that
the sketch is zoom of the situation near (x, y) = (0, 0).
The proof of Theorem 10.1 can be accomplished using similar, yet more
subtle, fixed-point arguments as sused in the proof of the Stable/Unstable
Manifold Theorem 3.8. We shall often write Wloc c (0) as W c (0) or just W c .
54
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
(T2) Center manifolds need not be C ω , even if f ∈ C ω , i.e., center mani-
folds are in general not analytic.
Proof. Consider the counter-example
x′ = x2 ,
(10.2)
y ′ = −y,
with (x(0), y(0)) = (x0 , y0 ). Clearly, (0, 0) is an equilibrium and the eigen-
values of the linearization are λ = 0, −1. The counter-example can be
solved explicitly
1
x(t) = 1 , y(t) = y0 e−t .
x0 − t
See Figure 23(b) for the phase portait. We claim W c (0)p = {(x, y) ∈ R2 :
y = hp (x)} for
p exp x1 for x < 0,
hp (x) :=
0 for x ≥ 0,
is a one-parameter family (for p ∈ R) of center manifolds. Indeed, the
tangent space is TWpc (0) = {y = 0} = E c (0). Each Wpc (0) is invariant
since for p fixed, {y = 0, x > 0} is a trajectory and for x < 0 we may
simply solve for t in x(t) above to obtain that
y = pe1/x
is also a trajectory for x < 0. Hence, W c (0) is not unique and it is not C ω
for p 6= 0.
Remark : Center manifolds do not even have to be C ∞ , even if f is analytic. We
leave it as an exercise to check that the system
x′1 = −x2 x1 − x31 , x′2 = 0, y ′ = −y + x21
is one possible required counterexample.
Separating the critial and non-critical components, we may write (10.1)
x′ = Bx + F (x, y), F = O(x2 , xy, y 2 )
(10.3)
y ′ = Cy + G(x, y), G = O(x2 , xy, y 2 )
where B ∈ Rdc ×dc with spec(B) ⊂ iR and C ∈ R(ds +du )×(ds +du ) with
spec(C) ∩ iR = ∅.
Theorem 10.4 (Center Manifold Reduction). The ODE (10.3) is locally
topologically equivalent near 0 to
x′ = Bx + F (x, H(x)),
(10.4)
y ′ = Cy,
c (0) = {(x, y) ∈ Rdc × Rds +du : y = H(x)} with H = O(x2 ).
where Wloc
55
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
Remark : Essentially analogus center manifold results hold for maps, where the
critical components in E c (0) are associated to multipliers on the unit circle.
In practice, the most important case is du = 0, which we assume for the
rest of this section. We can only apply normal form and bifurcation theory
on W c if we can calculate H, or at least an approximation h : Rdc → Rds .
Making the ansatz y = h(x) in (10.3) yields
and looking for a fixed point of the operator K on the space K of Lipschitz
functions with the sup-norm; cf. proof of Theorem 3.8 and note that (10.7)
essentially removes all decaying directions so only the center directions can
remain. To obtain the approximation h, we study a modified fixed point
56
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
problem. Consider θ ∈ Cc1 (Rdc , Rds ) with θ(x) = h(x) for |x| small, and
define
N (x) := Dθ[Bx + F (x, θ(x))] − Cθ(x) − G(x, θ(x))
and observe N (x) = O(|x|p ) as |x| → 0. We know H is a fixed point of K
and define a new operator by
Lw = K(w + θ) − θ.
x′ = xy,
(10.8)
y ′ = −y + px2 ,
in (10.5) to obtain
So we can select a2 = p, then all quadratic terms vanish, and then take
a3 = 0 to eliminate the cubic term to obtain
c
Wloc (0) = {(x, y) ∈ U = U (0) : y = h(x) = px2 + O(|x|4 )}.
x′ = px3 + O(x5 )
and the phase portraits for p > 0 and p < 0 are shown in Figure 24.
57
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
y y
(a) (b)
x x
Figure 24: Sketch of the center manifolds for Example (10.7). (a) p = p1 >
0. (b) p = p2 < 0 with |p2 | < p1 .
x′ = y − 1,
(10.9)
y ′ = −(y − 1) + px2 + x(y − 1),
at this point has eigenvalues 0 and −1. However, the system (10.9) is not
in standard form. Setting (x, y) = (x̃, ỹ + 1), inserting into (10.9), and
dropping the tiles yields
x′ = y,
(10.10)
y ′ = −y + px2 + xy.
Inserting this coordinate change in (10.10), and dropping the tiles, yields
the standard form
′
x 0 0 x 1 1 0
= + .
y′ 0 −1 y 0 −1 p(x + y)2 − (x + y)y
Now one may proceed analogously to Example 10.7 to find a center manifold
approximation (exercise!).
58
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
11 Higher Codimension
We return to our goal to unfold the Bogdanov-Takens (double-zero eigen-
value) bifurcation. At the bifurcation point q = 0 the normal form (9.3)
is
y1′ = y2 ,
(11.1)
y2′ = y12 ± y1 y2 ,
where we have dropped higher-order terms. Do third-order Taylor terms
matter? Unfortunately, the situation is not as simple as for the fold in
Lemma 8.7 and it helps to consider a more abstract viewpoint using the
following definition.
Definition 11.1. Let F ∈ C l (Rd , Rd ), F (0) = 0, and consider k ≤ l, k ∈ N.
Then the k-jet Jk F (y) is the Taylor polynomial of F up to and including
order k.
Therefore, our question is, whether the 2-jet (y2 , y12 ±y1 y2 ) determines
the topological type of the vector field, i.e., whether including higher-order
jets yield vector fields locally topologically not equivalent to the 2-jet. The
following example illustrates the technique we need to prove the 2-jet de-
terminacy of (11.1).
Example 11.2. Consider the planar ODE
y1′ = y12 − 2y1 y2 =: F1 (y1 , y2 ),
(11.2)
y2′ = y22 − 2y1 y2 =: F2 (y1 , y2 ).
59
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
as division by any power of r just corresponds to a re-parametrization of
time on orbits on the punctured plane (non-trivial exercise!). Yet, one
checks easily that S1 × {r = 0} now contains six equilibria as shown in
Figure 25.
y2
Φ
F F̄
y1
R2 S1 × [0, ∞)
60
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
q2 y2
(a) fold F (b)
q1 y1
homoclinic Hopf
H
Proposition 11.5. The ODE (11.3) has a curve of fold bifurcations given
by
F = {q ∈ R2 : 4q1 − q22 = 0}, (11.4)
and a curve of supercritical Hopf bifurcations
H = {q ∈ R2 : q1 = 0, q2 < 0}
which immediately yields the fold curve (11.4) upon inspection. One checks
that crossing F ∩ {q2 > 0} generates an unstable node y∗− and a saddle y∗+ .
Crossing F ∩ {q2 < 0} yields a stable node y∗− and a saddle y∗+ ; note
that this also explains the name saddle-node bifurcation as a synonym
for the fold bifurcation. It is straightforward to check (exercise!) that
y∗− undergoes a Hopf bifurcation on H but is only a neutral saddle for
{q ∈ R2 : q1 = 0, q2 > 0}.
61
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
easily via Proposition 11.5 except that between H and F ∩ {q2 > 0} we
must somehow destroy the periodic orbit generated in the Hopf bifurcation
without any local bifurcations. Let us, formally, motivate what we expect.
For ν > 0, we translate y∗− to the origin via (ỹ1 , ỹ2 ) = (y1 − (y∗− )1 , y2 ).
Upon dropping the tildes, this yields the ODE
y1′ = y2 ,
(11.5)
y2′ = y1 (y1 − ν) − ((y∗− )1 y2 + y1 y2 ).
Near the bifurcation point, we have to deal with small ν so we use a scaling
12 Global Bifurcations
Definition 12.1. Let (X , T , φt ) for T = R or T = Z be a dynamical
6 y∗ .
system with steady states x∗ , y∗ with x∗ =
(D1) An orbit γ = γ(t) is called a heteroclinic orbit from x∗ to y∗ if
lim γ(t) = x∗ .
t→±∞
62
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
W u (0)
(a) (b) (c)
W s (0)
Σu (0)
γ
x∗ γ
x∗ y∗ 0 Σs (0)
Σ
Figure 27: (a) Heteroclinic orbit. (b) Homoclinic orbit. (c) Sketch for
stable/unstable manifolds of a broken homoclinic and for the construction
of the split function.
63
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
(A1) (“non-resonance”) We have σ(0) := λ1 (0) + λ2 (0) 6= 0.
γ0
Figure 28: Illustration of Theorem 12.3 for σ(0) < 0. (a)-(c) correspond to
a variation of p through the bifurcation point at p = 0 shown in (b)
x′1 = λ1 x1 + f1 (x1 , x2 ),
(12.2)
x′2 = λ2 x2 + f2 (x1 , x2 ).
in (12.2); see Figure 29. Upon scaling coordinates, we obtain that in the
standard rectangle Ω = [−1, 1] × [−1, 1], the dynamics is given by
y1′ = λ1 y1 + y1 g1 (y1 , y2 ),
(12.5)
y2′ = λ2 y1 + y2 g2 (y1 , y2 ).
64
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
Wu
y2 ζ
τ2 Σh
Φ Ws
τ1
y0 (ξ0 , ζ0 )
y1 ξ
Σv
Ω = [−1, 1] × [−1, 1] Ω = [−1, 1] × [−1, 1]
Figure 29: Local construction near the saddle to prove Theorem 12.3.
ξ ′ = λ1 ξ,
(12.7)
ζ ′ = λ2 ζ.
65
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
The solution of the nonlinear system (12.5) can be written as a mild/integral
solution, inserting −τ1 gives
Z −τ1
−λ1 τ1
1 = y1 (−τ1 ) = y1 e + eλ1 (−τ1 −s) y1 (s)g1 (y(s)) ds
0
=: e−λ1 τ1 [y1 + R(τ1 )]
Differentiating y1 = eλ1 τ1 − R(τ1 ) gives
λ1 τ1 ′ λ1 τ1 1 −λ1 τ1 ′
(1, 0) = λ1 e Dτ1 − R Dτ1 = λ1 e 1− e R Dτ1 .
λ1
Solving for Dτ1 , and using (12.9) leads to
1
DΦ1 (y) = h i (1, 0). (12.10)
1 −λ1 τ1 ′
1− λ1 e R
66
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
(a) P (ζ) (b) P (ζ)
q=0
q=0
ζ ζ
Figure 30: Fixed points of the map P (ζ) upon varying q through q = 0
(solid black curve); the diagonal is shown in grey. (a) An unstable nonzero
fixed point (periodic orbit) occurs. (b) A non-zero stable fixed point (peri-
odic orbit) occurs.
67
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
Then for all |p| sufficiently small there exist infinitely many saddle limit
cycles in a neighbourhood of γ0 ∪ {0}.
Proof. We only construct the geometry of the Poincaré map. The final
result is going to follow from results in Section 13. Wlog we may assume
in a neighbourhood U = U (0) that
s
Wloc (0) = {x ∈ U (0) : x1 = 0},
u
Wloc (0) = {x ∈ U (0) : x2 = 0, x3 = 0}.
ζ ′ = λζ,
ξ1′ = αξ1 − ωξ2 , (12.11)
ξ2′ = ωξ1 + αξ2 ,
x2 G
Σh x2
Σv
L
x1 x3
Σv
x3
Figure 31: Sketch of Shilnikov homoclinic bifurcation and its proof. Phase
space on the left and return maps on Σv on the right.
68
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
The global map G can be given to lowest-order by Taylor expansion as
in the proof of the Andronov-Leontovich Theorem 12.3. By the Shilnikov
condition, G maps the spiral outside of the initial rectangle back to Σv as
shown in Figure 31.
69
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
k4 k3
k2 H1
k3
k4
k1
H0
k1 k2 g −1
· · · 01 · · · = · · · a0 a1 · · ·
Figure 32: Inverse g −1 of the standard horseshoe map g (exercise: draw
a suitable sketch for g). The unit square Y is stretched horizontally and
compressed vertically, then folded into a horseshoe shape, and finally pasted
back onto the rectangle. The four solid grey horizontal strips form the set
of points g −2 (Y ∩ g(Y) ∩ g 2 (Y)) staying in Y. The corners kj of Y are also
marked under the transformation g −1 . We also indicate the two vertical
strips in Y of the forward image g(Y) ∩ Y and we show the symbolic coding
via the horizontal strips of one point chosen as an example.
Consider the unit square Y = [0, 1] × [0, 1]. Define a map g by (I) lin-
early stretching Y vertically by a factor µ > 2 and linearly compressing
horizontally by a factor λ ∈ (0, 1/2) and (II) folding the resulting rectangle
back so that g(Y) ∩ Y consists of two disjoint vertical strips V0,1 . In par-
ticular, the folded portion lies outside Y so that on Y ∩ g −1 (Y) the map is
linear. The pre-image g −1 (g(Y) ∩ Y) consists of two horizontal strips H0,1 ;
see Figure 32. Now iterate this map
xk+1 = g(xk ), xk ∈ X := {x ∈ X : xk ∈ X for all k ∈ Z}
which is called the Smale horseshoe map. Observe that Y ∩ g(Y) ∩ · · · ∩
g n (Y) is the union of 2n disjoint vertical strips while Y∩g −1 (Y)∩· · ·∩g −n (Y)
is the union of 2n disjoint horizontal strips. The Jacobian of g is constant
and given on H0,1 by
±λ 0
Dg = (13.1)
0 ±µ
for + on H0 and − on H1 . Most points leave Y under iteration. The remain-
ing points X form a Cantor set (a compact, perfect, totally disconnected
set).
70
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
Let Σ2 be the space of bi-infinite sequences a = {aj }∞
j=−∞ with elements
from {0, 1} with the metric
∞
X
d(a, b) = |aj − bj |2−|j|
j=−∞
σ(a)j = aj+1 .
φ(x) = {aj }∞
j=−∞ with g j (x) ∈ Haj .
71
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
In particular, Theorem 13.2(T4) implies in combination with the per-
sistence of hyperbolic sets Theorem 7.5 that horseshoe dynamics is robust
under perturbations. This observation and Theorem 13.2(T1) almost finish
the proof of Shilnikov’s Theorem 12.6. What we still lack is to generalize
the horseshoe construction to curved structures instead of rectangles for
points (x, y) ∈ Y ⊂ R2 .
and some ρ ∈ (0, 1). Similarly, a horizontal curve y = h(x) is a curve for
which
Given two non-intersecting vertical curves v1 (y) < v2 (y), we may define a
vertical strip by
72
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
(T1) g has a hyperbolic invariant set X .
i.e., all images of some disk in M are eventually C 1 -close to a ball inside
the unstable manifold; see Figure 33.
Proof. First, note that iterating g and using the invariance of W s (0), im-
plies that if g n (M) intersects W s (0) transversally for any n ≥ 0. Hence, we
may focus on a neighbourhood of 0 upon picking a submanifold D ⊂ M.
Furthermore, it is easy to check that the result holds for C 0 , i.e.,
73
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
W u (0)
g n (D) M
B(x, r)
D
y W s (0)
0
|v1s | |As v s |
λ1 = u = u 0u ≤ a2 λ0 < aλ0 ,
|v1 | |A v0 |
where we stated the last obvious inequality as we are going to lose one
factor of a in the general case with higher-order terms due to cross-terms.
Similarly, one finds
|v1u |
≥ a−2 > a−1 .
|v0u |
Therefore, for any given δ > 0 and r > 0, there exists n0 such that for all
n ≥ n0 , we find
|vnu |
λn ≤ δ and ≥ r,
|v0u |
which implies the result.
74
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
Proof. (of Theorem 13.5) We present only the main geometric idea in
R2 . We may again assume that the stable and unstable manifolds W s (0)
and W u (0) are the (x1 , x2 )-coordinate axes in a neigbourhood V = V(0);
cf. (12.3)-(12.4). Since q ∈ W s (0) ∩ W u (0), it follows that there exists n0
such that
g n (q) ∈ V and g −n (q) ∈ V for all n ≥ n0 .
Now pick a thin rectangle R with long horizontal sides parallel to the x1 -
axis but separated from W s (0); see Figure 34.
x2
g k (R)
W u (0)
W s (0) q x1
The vertical boundary segments are chosen as straight lines such that
we may find iterates g ns (q) and g ns +1 (q) with ns ≥ n0 close to the hori-
zontal boundary of R. By the λ-Lemma 13.6, the vertical and horizontal
boundary segments of R are transformed to vertical and horizontal curves
for g k (R), which stretches as a thin rectangle parallel to W u (0). Also
for this image, we can pick k to match precisely two iterates g −nu (q) and
g −nu +1 (q) with nu ≥ n0 close to the vertical boundary of R; see Figure 34.
The λ-Lemma 13.6 guarantees that pre-images of vertical and horizontal
boundary segments are transformed to vertical and horizontal curves. This
construction has the required horseshoe geometry viewing R as our basic
domain for the horseshoe analogous to the unit rectangle above. What re-
mains to be checked is that suitable contraction and expansion conditions
hold as prescribed in Theorem 13.4, which is essentially evident (difficult
exercise!) since we may always also ensure by adjusting k that g k is a map
with sufficiently strong contraction and expansion as 0 is a saddle point.
75
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
14 Melnikov’s Method
Although the Smale-Birkhoff Homoclinic Theorem 13.5 provides a crite-
rion, how to prove the existence of a horseshoe, verifying the existence of
transverse homoclinic points is very difficult. In this section, we provide
one special case, where an analytical approach is possible. Consider the
ODE
x′ = f (x) + εg(x, t), x ∈ R2 , (14.1)
where f, g are smooth, g is periodic in t with minimal period T > 0 and
f = (∂x2 H, −∂x1 H)⊤ is Hamiltonian. The key assumption is that the
unperturbed system for ε = 0 contains a homoclinic orbit γ = γ(t) to a
hyperbolic saddle point p, i.e., for ε = 0 we have
Lemma 14.1. For ε > 0 sufficiently small, the Poincaré map has a locally
unique hyperbolic saddle point ptε0 = p + O(ε).
Proof. This just follows from the implicit function theorem applied to F :=
Id − Pεt0 and using that p is hyperbolic. Alternatively, one may just appeal
to Theorem 7.9 for a proof.
76
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
pε
γ(t)
t p
x1
x2 Σt0 L
Figure 35: Phase space for the system (14.2). The solid curves illustrate
the situation in the limit ε = 0, while dashed curves show the breaking of
the family of homoclinic orbits.
Since γεs (t; t0 ) remains in U for any t > t1 , the C k -closeness of the stable
manifolds implies
So the first equation in (14.3) follows. For the second one, the same argu-
ment following an orbit backward in time.
In fact, one even has an equation for the first-order approximation using
a variational equation (cf. Definition 5.4) along γ, i.e.,
77
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
shown in Figure 35. As before, consider γ(0) and define a one-dimensional
cross-section/line L via the normal vector
to γ(t) at γ(0). Let γεs (t0 ) := γεs (t0 ; t0 ) ∈ W s (ptε0 ) and γεu (t0 ) := γεu (t0 ; t0 ) ∈
W u (ptε0 ) be the unique points in L closest to γ(0). Define the (signed)
distance by
d(t0 ) := γεu (t0 ) − γεs (t0 ).
Lemma 14.3. The function d(t0 ) can be uniformly approximated by
f (γ(0)) ∧ (γ1u (t0 ) − γ1s (t0 ))
d(t0 ) = ε + O(ε2 ). (14.5)
|f (γ(0))|
Proof. By Lemma 14.2, we may approximate the difference γεu (t0 ) − γεs (t0 )
to first order by γ1u (t0 ) − γ1s (t0 ). Furthermore, recall the definition of the
wedge product a ∧ b = a1 b2 − b1 a2 , which yields that f ∧ (γ1u − γ1s ) is the
projection of γ1u − γ1s onto L and that |f (γ(0))| is a normalization factor;
see Figure 35.
By Lemma 14.3, we have d(t0 ) = εδ(t0 ; t0 )/|f (γ(0))| + O(ε2 ). The trick is
to find ODEs for δ u (t; t0 ) and δ s (t; t0 ) and solve them by integration. We
start with the stable part and differentiate
d s
δ (t; t0 ) = Df (γ(t − t0 ))γ ′ (t − t0 ) ∧ γ1s (t; t0 ) + f (γ(t − t0 )) ∧ (γ1s )′ (t; t0 ).
dt
78
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
Using the variational equation (14.4), γ ′ = f (γ), and omitting the argu-
ments (t − t0 ) and (t; t0 ) for notational simplicity leads to
However, Tr(Df (γ)) = 0 as the vector field is Hamiltonian and smooth (in
fact: C 2 would suffice here). Integrating the ODE for δ s from t0 to ∞ yields
Z ∞
s s
δ (∞; t0 ) − δ (t0 ; t0 ) = f (γ(t − t0 )) ∧ g(γ(t − t0 ), t) dt. (14.6)
t0
so that adding (14.6) and (14.7) and employing Lemma 14.3 gives us
M (t0 )
d(t0 ) = ε + O(ε2 ), (14.8)
|f (γ(0))|
W s (ptε∗ ) ∩ W s (ptε∗ ) 6= ∅.
x′ = y,
(14.9)
y ′ = x − x3 + εα cos(t) − εy,
79
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
where α is a parameter. The system (14.9) is Hamiltonian for ε = 0 with
Hamiltonian function
1 1 1
H(x, y) = y 2 − x2 + x4 .
2 2 4
There is a hyperbolic saddle point at p = (0, 0) and the level set {H = 0}
consists of two homoclinic orbits γ± ; see Figure 36.
y
Figure 36: Phase portrait showing the double-homoclinic loop in the un-
forced (i.e., ε = 0) Duffing equation (14.9).
In this case, one may even explicit calculate the homoclinic orbits and
check that
√ √
γ+ (t) = 2 sech(t), − 2 sech(t) tanh(t) , γ− (t) = −γ+ (t).
Therefore, we can also give an explicit formula for the Melnikov function.
We focus on γ+ (t) since the calculation is similar for γ− (t) by symmetry.
y
Let γ+ denote the second component of γ+ , then we have
Z ∞
y y
M (t0 ) = γ+ (t) α cos(t + t0 ) − γ+ (t) dt.
−∞
y
After plugging in the explicit formula for γ+ (t), the resulting integral can
actually be evaluated (non-trivial exercise!) and one obtains
4 √ π
M (t0 ) = − + 2απ sech sin(t0 ).
3 2
Therefore, after quite a bit of algebraic manipulation (exercise!), one finds
that there exists an amplitude α > 0 of the forcing such that M (t0 ) has
simple zeros. Theorem 14.5 applies and shows the existence of a transverse
homoclinic point. Therefore, the Smale-Birkhoff Theorem can be used to
show that the return map of the Duffing equation for suitable forcing con-
tains horseshoe dynamics.
80
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
15 Chaos
Although we have already seen the key example of chaotic dynamics, one
may also define the concept abstractly.
which can be taken as one of the simplest examples, where we can expect
sensitive dependence on initial conditions; see Figure 37(a).
(a) g(x) (b) g(x)
1 1
1 x 1 x
0 2 1 0 2 1
Figure 37: (a) Sketch of the doubling map 15.1. (b) Sketch of the tent
map (15.3) for p = 2.
Theorem 15.3 (Sensitive Dependence for the Doubling Map). The dou-
bling map (15.1) has sensitive dependence on initial conditions on X =
[0, 1].
Proof. We are going to show that g is expansive with constant 41 , i.e., for
any x0 6= y0 , there exists k such that
1
|g k (x0 ) − g k (y0 )| ≥ ,
4
81
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
which immediately implies the result. Consider two points x, y such that
x < y and y − x < 14 . We claim that
The logistic map is another example, where one may prove sensitive
dependence for certain parameters. The calculations are easier in the fol-
lowing simplification.
for p ∈ (1, 2]. For example, it is relatively easy to check (exercise!) that the
classical tent map g(x; 2) has sensitive dependence on initial conditions.
Theorem 15.6 (Tent Map Topological Transitivity). The tent map (15.3)
is topologically transitive for p = 2.
Remark : Similar result and proofs also work for other parameter values and var-
ious other classes of dynamical systems; see Section 16.
82
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
Proof. We use symbolic dynamics, similar to the horseshoe analysis.
Let Σ+2 = {s = s0 s1 · · · : sj ∈ {0, 1}} be the space of infinite one-sided
sequences of two symbols. Define the itinerary map h : [0, 1] → Σ+ 2 by
0 if g j (x) ∈ [0, 1/2] =: I0 ,
h(x)j =
1 if g j (x) ∈ [1/2, 1] =: I1 .
We exclude all points x, which are going to eventually hit the point x = 1/2,
which is not going to alter our argument regarding topological transitivity.
Then note that h satisfies h(g(x)) = σ(h(x)), where σ is the left-shift map,
i.e., σ(s)j = sj+1 for j ∈ N0 . We may associate subsets of [0, 1] with subsets
of symbols via
Observe that each Is0 s1 ···sn is an interval and has length 2−(n+1) . The
sequence of compact intervals
I s0 ⊃ I s0 s1 ⊃ I s0 s1 s2 ⊃ · · ·
1
It is not injective since, e.g., H(1, 0, 0, 0, · · · ) = 2 = H(0, 1, 1, 1, . . .).
83
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
Theorem 15.8 (Chaos in Interval Maps). S = [0, 1] is chaotic for the tent
map (15.3) with p = 2, for the doubling map (15.1), and for the logistic
map x 7→ 4x(1 − x).
xk+1 = g(xk ), g : Rd → Rd , xk ∈ Ω ⊆ Rd
for d ≥ 2 turns out to be difficult; the same problem occurs for flows in
dimension d ≥ 3. The following concept is a good practical indicator.
1 (k)
lj (x0 ) = lim ln λj (15.5)
k→∞ k
One may interpret, or define and then prove (15.5) afterwards, Lya-
punov exponents as deformation rates of the unit sphere centered at x0
under the linearized dynamics. In particular, looking at Ak (Sd ) and the
resulting axes of the ellipsoid one; see Figure 38.
S2 Ak
Figure 38: Lyapunov exponents can be viewed as a measure for the defor-
mation of the axes of a unit sphere into the axes of an ellipsoid.
84
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
Theorem 15.11 (Liouville Formula). Suppose lj (x0 ) exist for all j and
| det Dgp | is the same for every p, then
d
X
lj (x0 ) = ln(| det Dgx0 |).
j=1
1
l1 = lim ln µk = ln µ > 0, l2 = ln λ < 0, (15.6)
k→∞ k
which leads to the general observation that Lyapunov exponents are directly
related to multipliers for cases, where Dgx0 is independent of x0 . In fact,
one may prove (non-trivial!) that Lyapunov exponents are independent of
the starting point for certain classes of invariant sets.
Example 15.13. Consider the standard tent map g(x; 2) from (15.3). Ex-
cept at the single point x = 21 , we have l1 (x0 ) = ln(2) > 0.
85
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
16 One-Dimensional Maps
The tent map, the logistic map, and the doubling map have served as ex-
cellent examples to understand chaotic dynamics. In the case of continuous
maps, we can do even better and understand their orbit structure.
For a proof using transition graphs and symbolic dynamics see [14]. In
fact, one may get a lot finer information on the appearance of different
families of periodic orbits.
1 ⊳ 2 ⊳ 4 ⊳ 23 ⊳ · · · ⊳ 2n ⊳ 2n+1 · · ·
· · · ⊳ 9 · 2n+1 ⊳ 7 · 2n+1 ⊳ 5 · 2n+1 ⊳ 3 · 2n+1 ⊳ · · ·
· · · ⊳ 9 · 2n ⊳ 7 · 2n ⊳ 5 · 2n ⊳ 3 · 2n ⊳ · · · ⊳ 9 ⊳ 7 ⊳ 5 ⊳ 3.
Example 16.4. Consider the tent map (cf. Example 15.4), which can also
be defined as
so that Sharkovskii’s Theorem and the Li-Yorke Theorem apply. One can
prove that period three orbits do exist for large ranges of parameter values
as indicated already by Figure 39(a).
The classical tent map (for p = 2) shows strong similarity to the classical
logistic map yk+1 = 4yk (1 − yk ) (exercise: simulate the logistic map!).
86
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
1 1
x (a) g(x) (b)
0.8 0.8
0.6 0.6
0.4 0.4
0.2 0.2
0 0
0 0.5 1 1.5 2 0 0.2 0.4 0.6 0.8 1
p x
Figure 39: (a) Bifurcation diagram of the tent map (16.1) obtained by
direct simulation for a 500 point equally spaced mesh of the parameter
space p ∈ (0, 2]. Transients have been removed (here: first 500 iterations),
then 40 iterates are plotted for each value of p. (b) Illustration of the
cobweb construction for p = 1.95.
87
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
Proof. Suppose z is a critical point of g ′ , which implies g ′′ (z) = 0. Since
Sg < 0, we have by (16.2) that g ′′′ (z)/g ′ (z) < 0. Therefore, g ′′′ (z) and g ′ (z)
have opposite signs. If g ′ (z) < 0, it follows that g ′′′ (z) > 0 so z is a local
minimum of g ′ , so z is also a local maximum of |g ′ | using g ′ (z) < 0. In the
other possible case we have g ′′′ (z) < 0 and g ′ (z) > 0, which yields that z is
a local maximum of g ′ and hence also of |g ′ |. Since g ′ does not vanish in I
by assumption, the result follows.
Proof. The idea is to check that if there are too many attracting periodic
orbits, then the regions they attract must overlap. Let z be a periodic point
of period m and denote by W (z) the maximal attracting neighbourhood
of z, i.e., g mn (y) → z as n → ∞ for all y ∈ V and W (z) is the maximal
interval inside V (remark: V is also called the basin of attraction of z).
Note we have g m (W (z)) ⊆ W (z) by construction. Suppose
Example 16.9. Consider the logistic map g(x) = px(1−x). Then Sg(x) =
−6/(1 − 2x)2 for x 6= 1/2 and Sg(1/2) = −∞ at the single critical point.
88
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
Since Sg < 0 for x ∈ I = [0, 1], it follows that the logistic map has at most
three locally asymptotically stable periodic orbits.
17 Renormalization
Renormalization is one well-established tool in dynamical systems theory.
Here we illustrate it in two contexts, the second one concerns the question
about generating sequences upon parameter variation for periodic orbits in
unimodal maps as raised in Section 16.
Example 17.2. Consider the fold normal form for a map (cf. Theorem 8.12)
Assume q < 0 is small, so that (17.1) has no fixed points, yet is close to
bifurcation. Iterates passing from y > 0 to y < 0 need a long time as
the map is small near y = 0. This effect is known as critical slowing
down (or intermittency). We want to formally determine the number of
iterates N (q) to pass near zero asymptotically as q ր 0, y → 0. Consider
the second iterate map
g 2 (y; q) = (g ◦ g)(y; q) = q + (q + y − y 2 ) − (q + y − y 2 )2
= (2 − q)q + (1 − 2q)y − (2 − 2q)y 2 + 2y 3 − y 4
= 2q + y − 2y 2 + O(q 2 , qy, qy 2 , y 3 ) ≈ 2q + y − 2y 2 .
89
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
If N (q) iterates are required for g, then approximately N (q)/2 iterates are
required for g 2 for the same passage near y = 0. Let Rα (y) := αy =: Y and
Sβ := βq =: Q be scalings. For α = 2 and β = 4 define a transformation
R2 g 2 ((R2 )−1 y, (S4 )−1 q) = 2 g 2 (Y /2; Q/4) ≈ Q + Y − Y 2 ,
i.e., the scaling has brought back the mapping, at least approximately, into
the (self-)similar fold normal form; cf. to scaling and blow-up methods in
Section 11. However, this means we must have
N (q)
≈ N (4q) ⇒ N (q) = O(q −1/2 ) as q ր 0
2
giving our desired passage-time asymptotic approximation. (R+ , ·) =: G is
a group so we write the RG action on the space of continuous one-parameter
interval maps Z := C 0 (I × R, I × R) as
R
G × G × Z 7→ Z, R(g) = Rα g(Rα−1 ·, Sβ−1 ·), g ∈ Z.
More abstractly, instead guessing the scales α = 2, β = 4, and using ap-
proximations, one may study exact solutions to the functional equation
(D(2) g)(x; p) := g 2 (y; p) = Rα g(Rα−1 y, Sβ−1 q),
90
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
x
d1
P0 P1 P
p0 p1 p2 2 p3
Figure 40: Sketch of the bifurcation diagram of the logistic map (17.2) with
bifurcation points at pj and superstable cycles at Pj ; see also Examples 17.3,
17.5, and 17.7.
The previous construction does not depend upon the precise shape of
the logistic map. It can be carried out for large classes of maps:
91
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
Definition 17.6. Let I = [0, 1]. Then g ∈ C 0 (I, I) is called unimodal
if g(0) = 0 = g(1) and g has a unique local maximum xmax ∈ (0, 1).
If in addition, g has negative Schwarzian derivative on I, it is called S-
unimodal.
(a) g(x; P0 ) (b) g 2 (x; P1 ) (c) g 2 (x; P1 )
xmax xl xmaxxr xl xr
where δF is actually universal, i.e., the constant did not seem to depend
on the S-unimodal map he considered. Second, the constant is linked to
superstable cycles via the result
where κ1 > 0 is not universal, i.e., depends upon the particular map.
Thirdly, even for the phase space measure dn defined in (17.4) one finds a
scaling law
where κ2 > 0 is not universal but αF is. The constants δF , αF are known
as Feigenbaum constants. The key argument to establish Feigenbaum’s
universality observations is again renormalization. Consider Figure 41, and
observe that the graphs of
92
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
look similar once we zoom into a certain region around x = xmax and reflect
the graphs shown in Figure 41. In fact, xmax is a superstable fixed point of
both maps.
Example 17.7 suggests for a general S-unimodal map g = g(x; p) with
superstable cycle sequence Pn to consider the renormalization groups (R− , ·),
(R, +) with actions Rα and Tβ given by multiplication by α and translation
by β. More precisely, we consider the self-similar approximation
x
g(x, P0 ) ≈ αg 2 ; P1 , P1 = P0 + β0 , (17.5)
α
i.e., we claim the second iterate is, upon conjugacy by scaling induced by
Rα and shifting by Tβ0 for β0 chosen to yield the next superstable parameter
value; cf. Example 17.2. Upon iterating (17.5), we get
n
x
g(x, P0 ) ≈ αn g 2 ; Pn , (17.6)
αn
so if we start at P∞ =: Pn , i.e., at the onset of chaos, we get the functional
equation for just the fist step as
One has to augment (17.7) with boundary conditions to solve it. Upon
translating a-priori to the origin and adjusting the scale of x we see that
g ′ (0) = 0 and g(0) = 1 are reasonable conditions. The fixed point problem
for g turns out to be solvable (very difficult proof!) if and only if α = αF ,
thereby determining αF .
Remark : Linearizing the operator on the right-hand side of (17.7) using a Fréchet
derivative on a suitable function space shows that the fixed point g has one unstable
eigenvalue, which is precisely given by δF . Hence, the renormalization group and
the functional equation (17.7) indeed contain full information about the universal
Feigenbaum constants.
18 Attractors
Having seen that the full orbit structure of one-dimensional maps can be
rather involved (Sections 16-17) we now study the attracting parts of phase
space in general dynamical systems more abstractly. In this section, we
always consider a C 0 dynamical system (X , T , φt ), which is well-defined in
forward time, i.e., for all t ≥ 0, e.g., T = [0, ∞) for flows or T = N0 for
maps.
Definition 18.1. A compact set U is called a trapping region provided
that φt (U ) ⊂ int(U ) for all t > 0.
93
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
Example 18.2. Consider the Lorenz system, classically written as
x′1 = σ(x2 − x1 ),
x′2 = ρx1 − x2 − x1 x3 , (18.1)
x′3 = x1 x2 − βx3 ,
Proof. For (T1), one easily sees that A is closed since it is the intersection
of closed sets. Furthermore, if p ∈ A, then p ∈ φt (V) for all t > 0. One can
re-write the definition (18.3) of A as
\ \
A= φt+s (V) = φt (V)
t≥−s t≥−s
94
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
near x∗ . But then we may simply use the definition of the (global) unstable
manifold
[ \[
W u (x∗ ) = u
φt (Wloc (x∗ )) = u
φs ◦ φt (φ−s (Wloc (x∗ )))
t≥0 s≥0 t≥0
\[ \
⊂ φs (φt−s (V)) = φs (V) = A
s≥0 t≥0 s≥0
Although Definition 18.3 is very natural, i.e., just taking the forward
intersection of sets under the dynamics, it may not be appropriate for all
situations as the next example shows.
r′ = 1 − r, (18.4)
θ′ = − θ − 41 θ − 34
(a) (b)
U
φt (U )
A
Figure 42: (a) General idea of an attractor A inside a set U , i.e., the flow
φt contracts everything towards/onto A for t > 0. (b) Illustration of a key
example to understand the different attractor definitions.
95
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
Definition 18.6. Let A be a closed invariant non-empty set. The basin
of attraction is
µ(BA(A)) > 0
This definition fixes some intuition problems but also creates new ones.
Example 18.8. Regarding the ODE (18.4), we see that (1, 1/4) is its
Milnor attractor. However, if we look at
x′ = −x2 , x = x(t) ∈ R,
Hence, we shall simply use the term attractor from Definition 18.3 until
further notice and we shall explicitly write Milnor attractor if we refer to
Definition 18.7. As expected from Sections 13-17, the topology/geometry
of attractors can be extremely intricate.
96
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
Figure 43: Illustration of the Smale-Williams solenoid showing one forward
iteration from the solid torus to the double-wrapped torus of the map (18.5).
One the right-hand side a cross-section is shown.
97
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
where dA is the surface area form for ∂E(t) and ~n is the associated unit
normal vector. Hence, the Divergence Theorem implies
Z
d
V (t) = ∇ · f (x) dx. (18.6)
dt E(t)
d
V (t) = −(σ + 1 + β)V (t), (18.7)
dt
which implies the result by solving (18.7).
Remark : The change-of-volume formula (18.6) under a vector field f is valid in
far more generality if f and E(t) are smooth in Rd . It is then proved using Stokes’
Theorem instead of the Divergence Theorem, which itself is evidently a special
case of Stokes’ Theorem.
Also for the Smale-Williams solenoid, one finds that it has three-dimensional
volume equal to zero. These observations motivate an alternative defini-
tion for a strange attractor A, i.e., that A should be chaotic and A should
satisfy a certain “dimension requirement”. To make this precise, we need
a suitable definition of dimension.
ln N (r, S)
dimB (S) := lim .
r→0 ln(1/r)
For example, we can cover the unit square [0, 1]2 by N = 22j boxes of
side length r = 2−j , so taking the limit j → ∞ yields that dimB ([0, 1]2 ) = 2.
Example 18.14. To show that dimB (S) does capture non-integer consider
the classical middle-third Cantor set. It is obtained by removing the middle
third (1/3, 2/3) from [0, 1], then removing the middle thirds from K1 =
[0, 1/3] ∪ [2/3, 1] to create
1 2 1 2 7 8
K2 = 0, ∪ , ∪ , ∪ ,1 ,
9 9 3 3 9 9
and then repeating the process iteratively to create Kj and the resulting
Cantor set \
K= Kj .
j≥0
98
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
At level j for Kj , we need 2j boxes/intervals of length (1/3)j . Hence, we
obtain
ln(2j ) ln 2
dimB (K) = lim j
= ,
j→∞ ln(3 ) ln 3
which is approximately equal to 0.6309, i.e., certainly not an integer.
Recall that we already encountered Cantor sets as invariant sets for dy-
namics in the case of the Smale-Williams solenoid and the Smale horseshoe.
In addition to box-counting dimension, there are several other notions,
which we shall not discuss here, such as as Lyapunov dimension, Haus-
dorff dimension, and correlation dimension, which can also capture
non-integer dimensions. An alternative definition of a strange attrac-
tor is a chaotic attractor, which is also a fractal set, i.e., of non-integer
dimension.
19 Invariant Measures
Having established the basic theory of hyperbolic, non-hyperbolic and chaotic
dynamical systems, it is difficult to make vast progress using the machinery
we have used so far. Furthermore, the occurrence of chaos naturally leads
to the idea to consider additional structures for a dynamical system, and
we shall look at measure-theoretic tools first.
Certain iterated maps as well as flows (via a time-t map) might generate
measure-preserving maps (or measure-preserving transformations). Note
that sets/spaces of finite measure can always be normalized to have measure
one, so just looking at probability measures is not a major restriction.
99
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
for r ∈ R. Since the map just shifts any measurable set and potentially
decomposes it, one easily checks that it is also a measure-preserving trans-
formation.
An invariant measure can be thought of as describing the distribution
of a typical trajectory of the system.
Theorem 19.4 (Poincaré Recurrence). Let (X , F, µ) be a probability space.
Suppose g : X → X is measure-preserving and fix any B with µ(B) > 0.
Then almost every point in B returns to B infinitely often under forward
iteration of g.
Proof. For N ≥ 0, define
∞
[ ∞
\
BN := g −n (B) and B∗ := BN
n=N N =0
where the second conclusion follows since g ni −nj (g nj (x)) ∈ B for all j ∈ N;
cf. the proof of Theorem 18.4 (T3). Next, observe that the set B ∩ B∗ could
have small measure but we would like to show that µ(B ∩ B∗ ) = µ(B).
Indeed, we see that g −1 (BN ) = BN +1 implies µ(BN ) = µ(BN +1 ) as g is
measure-preserving. Upon iterating the argument, this yields µ(BN ) =
µ(B0 ) for all N . Furthermore, we observe
∞
!
\
B0 ⊇ B1 ⊇ B2 ⊇ · · · ⇒ µ BN = µ(B0 ),
N =0
100
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
(a) (b) g(x)
1
X
...
B
1 1 x
3 2 1
exists for all f ∈ F ∗ . To prove the claim (19.2) one uses a standard di-
agonal argument. Label the elements of F ∗ by f1 , f2 , etc. and first use
that {Sfm1 (x)}∞
m=1 is bounded to extract a subsequence m1 (k) such that
m (k)
Sf1 1 (x) converges as k → ∞. Out of this subsequence extract a further
m (k)
subsequence m2 (k) such that Sf2 2 (x) converges and continue this pro-
cess. Finally, pick the diagonal sequence by setting nj = mj (j) and verify
that (19.2) holds. Next, we use density of F ∗ , which implies that for any
h ∈ C(X ) and any ε > 0, there exists f ∈ F ∗ such that
max |h(y) − f (y)| < ε. (19.3)
y∈X
A direct triangle inequality step now yields that for sufficiently large j, we
may combine (19.2) and (19.3) to obtain
n n n
Sh j (x) − Sf∞ (x) ≤ S|fj−h| (x) + Sf j (x) − Sf∞ (x) ≤ 2ε.
n
Hence, Sh j (x) is a Cauchy sequence and the limit Sh∞ (x) exists for every
h ∈ C(X ) and defines a bounded linear functional
Lx (h) := Sh∞ (x), (19.4)
101
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
which is also positive in the sense that h ≥ 0 implies Lx (h) ≥ 0 by the
construction (19.1), and kLh k = 1. The Riesz Representation Theorem
implies that for a bounded positive linear functional on C(X ), there exists
a measure µx such that
Z
Lx (h) = h(y) dµx (y). (19.5)
X
102
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
To prove that (19.7) is invariant for (19.6), note that the pre-image
of an interval (a, b) consists of infinitely many intervals. If we restrict to
(1/(n + 1), 1/n), then the pre-image is (1/(n + b), 1/(n + a)). Now one
calculates
∞ !
−1
[ 1 1
µ(g ((a, b))) = µ ,
n+b n+a
n=1
∞
1 X 1+a+n b+n
= ln · = µ((a, b))
ln 2 a+n 1+b+n
n=1
20 Ergodicity
Of course, it may happen that a measure-preserving transformation de-
composes into several invariant subsets. However, then it makes sense to
restrict to each subset and study indecomposable objects; see Figure 45(a).
103
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
(a) (b) lim f (x) + f (g(x)) + f (g 2 (x)) + . . .
n→∞
X
=
R
X f (x)dµ
g2 g(x)
x
g1 g
g 2 (x)
X
Figure 45: (a) Illustration of the notion of ergodicity: either we can split
the phase space into two invariant regions so that g restricted to the re-
gions leaves them invariant, or this is not possible and we are ergodic. (b)
Illustration of the Birkhoff Ergodic Theorem 20.7.
104
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
One may study Lg on various function spaces. As an important case,
consider L1 = L1 (X , R; µ) and observe that
Z Z
1 1
Lg L ⊂ L , kLg f kL1 = f (g(x)) dµ(x) = f (y) dµ(y) = kf kL1 .
X X
f0 := 0, fn := f + Lf + L2 f + · · · + Ln−1 f FN := max fn ≥ 0.
0≤n≤N
R
Then it follows that {x:FN (x)>0} f (x) dµ(x) ≥ 0.
where one uses FN = 0 on X \ A for the first term. The second term on the
right-hand side in (20.2) can be estimated from below since FN ≥ 0 implies
LFN ≥ 0, which gives
Z Z Z
f dµ ≥ FN dµ − LFN dµ = 0, (20.3)
A X X
105
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
Proof. Just restrict g to A so that we can simply apply Theorem 20.5 with
A = X and f = h − α to obtain the result.
R R
(T2) f ∗ ◦ g = f ∗ and f dµ = X f ∗ dµ. X
R
(T3) If g is ergodic, then f ∗ ≡ X f dµ so that
n Z
1X
lim f (g j (x)) = f dµ, (20.6)
n→∞ n X
j=1
106
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
Combining (20.7)-(20.8) gives
αµ(Eα,β ) ≤ βµ(Eα,β ) ⇒ µ(Eα,β ) = 0 if β < α.
1 Pn−1
This shows (I) and to check (II) let hn (x) := n j=0 f (g j (x)) so Fatou’s
Lemma implies f ∗ ∈ L1 since
Z n−1 Z Z
1X j
lim inf hn dµ ≤ lim inf |f ◦ g | dµ = |f | dµ.
X n→∞ n→∞ n X X j=0
The first statement in (T2), i.e., f ∗ = f ∗ ◦ g is now easy to check. For the
second part, let Dkn := {x ∈ X : k/n ≤ f ∗ (x) ≤ (k + 1)/n}, which are sets
to decompose f ∗ at “scale n” for k ∈ Z, n ∈ N. One may select ε > 0
sufficiently small so that Dkn ∩ B(k/n)−ε = Dkn , which makes Corollary 20.6
applicable again
Z Z
k n k
f dµ ≥ − ε µ(Dk ) ⇒ f dµ ≥ µ(Dkn ).
n
Dk n n
Dk n
107
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
Proof. Fix ε > 0. The proof is a standard ε/3-argument to check that
Sn f is a Cauchy sequence so we provide only the main steps. Consider
h ∈ L∞ such that kh − f kLp ≤ ε/3. Since X is a probability space, it
follows that h ∈ Lp . Therefore, the Birkhoff Ergodic Theorem implies that
Sn h converges in Lp . This yields the existence of a sufficiently large n ∈ N
such that kSn h − Sn+k hkLp ≤ ε/3. Hence, the triangle inequality and
kSn f kLp ≤ kf kLp give
21 Mixing
The next result hints at a crucial consequence of ergodicity, when we are
interested in return probabilities.
Theorem 21.1 (Limit of Cesàro Means). Let (X , F, µ) be a probability
space and suppose g : X → X is a measure-preserving map. Then g is
ergodic if and only if for all A, B ∈ F we have
n−1
1X
lim µ(g −j (A) ∩ B) = µ(A)µ(B). (21.1)
n→∞ n
j=0
The converse of our result assumes (21.1). Suppose S is any invariant set,
g −1 S = S, and let A = S = B. Then (21.1) implies µ(S) = µ(S)2 , so
µ(S) = 0 or µ(S) = 1, which just means ergodicity holds.
108
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
We know by Theorem 21.1 that the sets µ(g −j (A) ∩ B) become asymp-
totically independent “on average” for ergodic maps. Hence, it is very
natural to ask, whether we can do better than just average properties.
The converse implications in Proposition 21.3 are not true; see e.g. Ex-
ample 21.10.
The considerations the last example can easily be turned into a more
general proof of the following result.
109
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
Proposition 21.5. Expanding circle endomorphisms are strong-mixing.
For more general maps, proving mixing and/or working with proper-
ties directly becomes almost impossible. However, an operator-theoretic
viewpoint provides a nice characterization.
Definition 21.6. Let Lp (X , R; µ) =: Lp for p ∈ [1, ∞]. View the Koopman
operator as Lg = L : L∞ → L∞ and let Pg = P : L1 → L1 be its adjoint
defined by
Z Z
!
hPf, hiL2 = (Pf )h dµ = f (Lh) dµ = hf, LhiL2 . (21.5)
X X
for f ∈ L1
and h ∈ L∞ . The operator P is called the Perron-Frobenius
operator.
Theorem 21.7 (Operator-Characterization of Mixing). Let (X , F, µ) be a
probability space and g : X → X be measure-preserving. Then g is strong-
mixing if and only if
lim hP n f, hi = hf, 1ih1, hi, f ∈ L1 , h ∈ L∞ , (21.6)
n→∞
so the desired equality holds for indicator functions, and by linearity also
for linear combinations of indicator functions, which are just the standard
simple functions. We approximate h ∈ L∞ by simple functions hk such
that hk → h, and we approximate f ∈ L1 by simple functions such that
fk → f . Then we have
|hP n f, hi − hf, 1ih1, hi| ≤ |hP n f, hi − hPfk , hk i|
+|hP n fk , hk i − hfk , 1ih1, hk i| (21.8)
+|hfk , 1ih1, hk i − hf, 1ih1, hi|
so that a standard ε/3-argument yields the result as all three terms on the
right-hand side of (21.8) can be made small. The converse of the result is
easier since if (21.6) holds then setting f = 1A and h = 1B gives strong-
mixing.
110
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
The last results naturally point into the direction of using invariants of
the Koopman operator L (respectively the Perron Frobenius operator P)
to characterize/classify measure-preserving transformations. Observe that
Lg = L : L2 → L2 is a unitary operator since it preserves the L2 inner
product. Hence, if Lf = λf for some λ ∈ C, i.e., λ is an eigenvalue and f
is an eigenfunction, then λ ∈ {z ∈ C : |z| = 1}.
Definition 21.8. Consider a probability space (X , F, µ) and let L2 =
L2 (X , F, µ). The Koopman operator L = Lg : L2 → L2 , defined by
Lf (x) = f (g(x)), has:
(D1) pure-point spectrum if the eigenfunctions of L span L2 ;
(D2) continuous spectrum if 1 is the only eigenvalue.
Remark : Of course, many linear operators have both types of spectra, i.e., some
eigenvalues and then some remaining part consisting of continuous spectrum. This
is highly relevant for the dynamical systems analysis of PDEs [15].
The next result illustrates that one can characterize a lot of dynamical
effects of g : X → X purely by looking at the spectrum of Lg = L.
Theorem 21.9 (Spectral Characterization). The following results hold:
(T1) g is ergodic if and only if 1 is a simple eigenvalue of L.
(T2) If g is ergodic, then every eigenvalue of L is simple.
(T3) Suppose g is invertible then g is weak-mixing if and only if g has
continuous spectrum.
Proof. (T1) is easy to check (exercise!). Regarding (T2), if g is ergodic,
then every eigenfunction is constant a.e., so that if there are two eigenfunc-
tions f1 and f2 with the same eigenvalue then f1 /f2 is also an eigenfunction
with eigenvalue 1 so f1 /f2 is also constant a.e., which proves the result. For
(T3), g weak-mixing implying continuous spectrum is relatively easy (exer-
cise!). The converse is difficult, i.e., we now assume that g has continuous
spectrum. An equivalent characterization of weak-mixing is
n−1
1X
lim |hLj f, f i − hf, 1ih1, f i|2 = 0 for f ∈ L2 .
n→∞ n
j=0
111
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
Since L is a unitary operator on the Hilbert space L2 , we may apply the
Spectral Theorem, which guarantees for every f ∈ L2 the existence of a
unique finite Borel measure µf on the spectrum K such that
Z
n
hL f, f i = λn dµf (λ). (21.10)
K
Since for points λ in the spectrum we have λ̄ = λ−1 , we may re-write the
sum in (21.11) as
n−1 Z n−1 Z Z
1X n 1X 2 n −n
| λ dµf (λ)| = λ dµf (λ) λ dµf (λ)
n K n K K
j=0 j=0
n−1 Z
1X
= (λτ̄ )n d(µf × µf )(λ, τ )
n K×K
j=0
Z n−1
1X
= (λτ̄ )n d(µf × µf )(λ, τ ) (21.12)
K×K n
j=0
where the second equality step follows from a re-labelling of the second
variable and using Fubini’s Theorem; essentially, the last argument is a
classical case of doubling variables. For points off the diagonal in K × K,
we get
n−1
1X n 1 1 − (λτ̄ )n
lim (λτ̄ ) = lim = 0. (21.13)
n→∞ n n→∞ n 1 − (λτ̄ )
j=0
Since µf is non-atomic, the diagonal has measure zero, the limit in (21.13)
holds a.e., and the Dominated Convergence Theorem can be applied to (21.12)
to get the result.
112
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
If r ∈ R\Q, then the eigenfunctions span L2 and Theorem 21.9(T3) implies
that the irrational circle rotation is not weak-mixing. However, it is shown
in Example 20.3 that irrational circle rotations are ergodic; cf. Proposi-
tion 21.3.
22 Entropy
Spectral properties provide one possible way to distinguish two dynamical
systems. Hence, a natural question to ask is, which other useful invariants
one may define? Borrowing the concept of entropy from statistical physics
has turned to yield one very helpful invariant. There are two related ap-
proaches to entropy, a measure-theoretic and a topological. We start with
the measure-theoretic and consider a measure-preserving transformation
g:X →X
113
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
so information is additive; in fact, using a logarithm (up to a constant)
is the only choice if one requires the last observations from a definition of
information.
µ(C1 ∩ C2 )
µ(C1 |C2 ) := .
µ(C2 )
Hence, we can now average the usual entropy conditionally over another
partition.
114
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
(T3c) C ⊆ B ⇒ H(A|C) ≥ H(A|B);
Using these last two results, multiplying both sides by µ(Cj ) and summing
over the indices i, j we get
X µ(Ai ∩ Cj ) X µ(Ai ∩ Bk ) µ(Ai ∩ Bk )
µ(Ai ∩ Cj ) ln ≤ µ(Bk ∩ Cj ) ln
µ(Cj ) µ(Bk ) µ(Bk )
i,j i,j,k
X µ(Ai ∩ Bk ) µ(Ai ∩ Bk )
= µ(Bk ) ln ,
µ(Bk ) µ(Bk )
i,k
115
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
Proof. As above, proofs are lengthy but not conceptually difficult. Let us
just show that, how the results of Theorem 22.6 lead to (T3). We want to
use the definition in (22.2) and start with the estimate
n−1
_ (T2c) n−1
_ n−1
_
H g −j A ≤ H g −j A ∨ g −j B
j=0 j=0 j=0
n−1
_ n−1
_ n−1
_
(T1c)
= H g −j B + H g −j A g −j B .
j=0 j=0 j=0
The second term on the right can be be estimated from above as follows
n−1
_ n−1
_ (T4c) n−1
X n−1
_
H g −j A g −j B ≤ H g −j A g −j B
j=0 j=0 j=0 j=0
(T3c) n−1
X
≤ H(g −j A|g −j B)
j=0
(T5c)
= nH(A|B).
Hence, combining this estimate with the previous one, dividing by n, and
taking the limit n → ∞ yields the result.
Finally we can introduce a crucial theorem, how we may calculate en-
tropy in certain cases as it allows us to select certain sub-algebras respec-
tively certain types of partitions.
Theorem 22.8 (Kolmogorov-Sinai). Suppose g is a measure-preserving
invertible transformation on (X , F, µ). Let A be a finite sub-algebra of F
such that
n
_
lim g j A =: lim An = F,
n→∞ n→∞
j=−n
i.e., A is generating, then we may conclude h(g) = h(g, A).
Proof. The idea of the proof is essentially an approximation procedure in-
volving multiple steps with error terms in the form of conditional entropies,
which have to be controlled. So let us first find the error term.
Step 1: (error term identification) To conclude our result, it is going to
sufficefrom the definition of h(g) to consider any finite partition B ⊆ F and
prove that h(g, B) ≤ h(g, A). Using Theorem 22.7, we the main estimate
(T 3)
h(g, B) ≤ h(g, An ) + H(B|An )
(T 4)
= h(g, A) + H(B|An ).
116
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
So if we could show that limn→∞ H(B|An ) = 0 for a generating partition
A, then the result would follow.
Step 2: (partition approximation) Recall the symmetric difference
between two sets C1 , C2
117
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
which is going to follow once we have verified the symmetric difference
condition in (22.3). Note that the statement (22.4) is already very close to,
what we want, i.e., it gives us control over arbitrary subalgebras just using
our generating sequence An . To prove (22.4), first let us select (A∞ )j such
that µ(Bj ∩ (A∞ )j ) < δ̃ for some δ̃ > 0, which can be chosen sufficiently
small in the end of the argument since A∞ does generate F by assumption.
Then we set [
Dj := (A∞ )j \ ((A∞ )i ∩ (A∞ )k ),
i6=k
118
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
which is a measure-preserving transformation. If we want to compute the
entropy via the Kolmogorov-Sinai Theorem 22.8 requires the choice of a
helpful generating subalgebra. One natural choice is
since the sets are simple and one verifies quite quickly that we can obtain
all cylinder sets upon using shift dynamics so that
∞
_
g j A = F.
j=−∞
Observe that the typical elements of An have measure that are easy to
calculate
µ Aj0 ∩ g −1 Aj1 ∩ g −(n−1) Ajn−1 = µ ({x : x0 = j0 , . . . xn−1 = jn−1 })
= pj0 · · · pjn−1 .
119
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
References
[1] K.T. Alligood, T.D. Sauer, and J.A. Yorke. Chaos: An Introduction to Dynamical Systems.
Springer, 1996.
[2] V.I. Arnold. Ordinary Differential Equations. MIT Press, 1973.
[3] V.I. Arnold. Geometrical Methods in the Theory of Ordinary Differential Equations.
Springer, New York, NY, 1983.
[4] M. Brin and G. Stuck. Introduction to Dynamical Systems. CUP, 2002.
[5] C. Chicone. Ordinary Differential Equations with Applications. Texts in Applied Mathe-
matics. Springer, 2nd edition, 2010.
[6] P. Glendinning. Stability, Instability and Chaos. CUP, 1994.
[7] J. Guckenheimer and P. Holmes. Nonlinear Oscillations, Dynamical Systems, and Bifur-
cations of Vector Fields. Springer, New York, NY, 1983.
[8] J.K. Hale. Ordinary Differential Equations. Dover, New York, NY, 2009.
[9] J.K. Hale and H. Kocak. Dynamics and Bifurcations. Springer, New York, NY, 1991.
[10] M.W. Hirsch, S. Smale, and R. Devaney. Differential Equations, Dynamical Systems, and
an Introduction to Chaos. Academic Press, 2nd edition, 2003.
[11] G. Iooss and D.D. Joseph. Elementary Stability and Bifurcation Theory. Springer, 1997.
[12] A. Katok and B. Hasselblatt. Introduction to the Modern Theory of Dynamical Systems.
CUP, 1995.
[13] C. Kuehn. Multiple Time Scale Dynamics. Springer, 2015. 814 pp.
[14] C. Kuehn. Dynamical Systems I: ODEs & Introduction to Nonlinear Systems. AMS Open-
Math Notes, 2018.
[15] C. Kuehn. PDE Dynamics: An Introduction. SIAM, 2019.
[16] Yu.A. Kuznetsov. Elements of Applied Bifurcation Theory. Springer, New York, NY, 3rd
edition, 2004.
[17] A. Lasota and M.C. Mackey. Chaos, Fractals and Noise. Springer, 1994.
[18] G.C. Layek. An Introduction to Dynamical Systems and Chaos. Springer, 2015.
[19] A.J. Lichtenberg and M.A. Lieberman. Regular and Chaotic Dynamics. Springer, 1992.
[20] J.D. Meiss. Differential Dynamical Systems. SIAM, 2007.
[21] L. Perko. Differential Equations and Dynamical Systems. Springer, 2001.
[22] R.C. Robinson. An Introduction to Dynamical Systems: Continuous and Discrete. AMS,
2013.
[23] R. Seydel. Practical Bifurcation and Stability Analysis. Springer, 1994.
[24] S. Sternberg. Dynamical Systems. Dover, 2010.
[25] S.H. Strogatz. Nonlinear Dynamics and Chaos. Westview Press, 2000.
[26] G. Teschl. Ordinary Differential Equations and Dynamical Systems. AMS, 2012.
[27] F. Verhulst. Nonlinear Differential Equations and Dynamical Systems. Springer, 2006.
[28] P. Walters. An Introduction to Ergodic Theory. Springer, 2000.
[29] S. Wiggins. Global Bifurcations and Chaos. Springer, 1998.
[30] S. Wiggins. Introduction to Applied Nonlinear Dynamical Systems and Chaos. Springer,
New York, NY, 2nd edition, 2003.
120
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
Index
ε/3-argument, 108 Cesàro mean, 108
chaotic
adjoint, 50 attractor, 97
Anosov diffeomorphism, 35 chaotic set, 83
aperiodic, 32 chart, 15
Arnold’s cat map, 22 circle endomorphism, 109
atlas, 15 circle rotation, 8, 104
attracting set, 94 cobweb, 5
attractor, 94 codimension, 44
chaotic, 97 common refinement, 113
hyperbolic, 97 complexification, 45
Milnor, 96 conditional
strange, 97, 99 entropy, 114
Banach Fixed Point Theorem, 6 probability, 114
basin of attraction, 88, 96 cones
bifurcation, 39 stable/unstable, 38
Bogdanov-Takens, 53 conjugate, 23
condition, 44 constant of motion, 13
diagram, 40 continued fraction expansion, 102
flip, 46 continuous spectrum, 111
fold, 41, 42, 46 continuous-time, 5
function, 49, 63 contraction mapping, 6
Hopf, 45 critical slowing down, 89
Neimark-Sacker, 48 curve, 16
period-doubling, 46 cycle, 8
pitchfork, 40 cylinder set, 118
saddle-node, 41 decay of correlations, 109
transcritical, 44 determinacy, 59
unfolding, 41 diagonal argument, 101
Birkhoff Ergodic Theorem, 106 differentiable manifold, 15
blow-down, 60 dimension
blow-up, 6, 60 box-counting, 98
Bogdanov-Takens discrete-time, 5
bifurcation, 53 doubling map, 81
box-counting dimension, 98 doubling operator, 90
Duffing equation, 22, 79
cat map, 22
dynamical system, 4
center, 14
manifold, 54 eigenfunction, 111
121
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
eigenvalue harmonic oscillator, 10
of an operator, 111 Hartman-Grobman Theorem, 24
entropy, 113, 114 heteroclinic orbit, 62
equilibrium, 7 homological equation, 50
equivalence, 23 Hopf
ergodic, 103 bifurcation, 45
Ergodic Theorem, 106, 107 horizontal curve, 72
expanding circle endomorphism, 109 horizontal strip, 72
hyperbolic, 17
Feigenbaum constants, 92 attractor, 97
fiber topological equivalence, 40 set, 34
first integral, 13 toral automorphism, 22
first Lyapunov coefficient, 45
fixed point, 6, 7 inclination lemma, 73
flip indecomposable, 94
bifurcation, 46 information, 113
flow, 5 intermittency, 89
focus, 14 invariance equation, 56
fold invariant, 8, 113
bifurcation, 42, 46 negatively, 8
fold bifurcation, 41 positively, 8
Fourier series, 104 invariant measure, 99
fractal, 99 iterated map, 5
Fredholm alternative, 51
functional equation, 25, 90 jet, 59
fundamental solution, 29 join, 113
122
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
Lyapunov period, 8
coefficient, 45 period-doubling
exponent, 84 bifurcation, 46
function, 11 period-doubling sequence, 90
stability, 9 periodic
stability criterion, 11 orbit, 8
Lyapunov-Perron method, 18 Perron-Frobenius operator, 110
phase
manifold, 15 portrait, 9
stable/unstable, 18, 21 phase space, 4
maximal, 94 pitchfork bifurcation, 40
Maximal Ergodic Theorem, 105 Poincaré
measure-preserving, 99 map, 28
Melnikov function, 78 normal form, 50
Milnor attractor, 96 section, 27
minimum principle, 87 Poincaré
mixing, 109 recurrence, 100
strong, 109 point
weak, 109 spectrum, 111
Morse-Smale flow, 33 positive operator, 105
multi-valued, 6 positively invariant, 8
multipliers, 14 pseudo-orbit, 36
near identity, 50 rectification theorem, 29
negatively invariant, 8 recurrence, 100
Neimark-Sacker refinement, 113
bifurcation, 48 renormalization, 89
Neumann series, 25 residual set, 33
node, 14 resonant terms, 50
non-degeneracy condition, 44 Riemannian
nonresonance condition, 67 manifold, 34
nonwandering, 32 metric, 34
normal form, 42, 45 Riesz Representation Theorem, 102
Poincaré, 50 Rokhlin metric, 117
nullcline, 8 rotation map, 8, 104
observable, 101 route-to-chaos, 90
ODE S-unimodal, 92
linear, 4 saddle, 14
orbit, 7 saddle quantity, 64
partition, 113 saddle-node bifurcation, 41
Peixoto’s Theorem, 33 scaling, 62, 89
123
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
Schwarzian derivative, 87 vector, 16
self-similar, 89 tangent bundle, 35
semiflow, 5 tangent map, 34
sensitive dependence, 81 tangent space, 16
shadowing, 36 tent map, 82
theorem, 36 time, 4
Sharkovskii time average, 106
ordering, 86 time-T map, 30
Sharkovskii Theorem, 86 time-one map, 26
shift map, 8 topological
Singer’s Theorem, 88 normal form, 42, 45
Smale horseshoe, 70 topological equivalence, 23
Smale-Williams solenoid, 96 fiber, 40
solenoid, 96 topologically transitive, 82
space average, 106 trajectory, 7
Spectral Theorem, 112 transcritical
spectrum, 111 bifurcation, 44
split function, 63 transversality
stability condition, 44
Lyapunov, 9 transverse
stable homoclinic point, 73
cone, 38 transverse intersection, 28
distribution, 35 trapping region, 93
eigenspace, 17 travelling wave, 69
manifold, 18, 21
structurally, 32 unfolding, 41
state space, 4 unimodal, 92
steady state, 7 unit tangent bundle, 38
strange universal, 92
attractor, 97 unstable, 9
strange attractor, 99 cone, 38
stratum, 40 distribution, 35
strict Lyapunov function, 11 eigenspace, 17
strong-mixing, 109 manifold, 18, 21
structurally stable, 32 structurally, 32
subcritical Hopf, 46
van der Pol, 7
supercritical Hopf, 46
variation-of-constants, 18
superstable, 91
variational equation, 29
suspension, 30
vector field, 6
symmetric difference, 117
vertical curve, 72
tangent vertical strip, 72
124
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22
Von Neumann Ergodic Theorem, 107
wave
speed, 69
weak-mixing, 109
125
AMS Open Math Notes: Works in Progress; Reference # OMN:201709.110713; Last Revised: 2019-01-05 10:35:22