Introduction To Dynamical Systems PDF
Introduction To Dynamical Systems PDF
Dynamical
Systems
Michael Brin
and Garrett Stuck
Introduction to Dynamical Systems
MICHAEL BRIN
University of Maryland
GARRETT STUCK
University of Maryland
PUBLISHED BY CAMBRIDGE UNIVERSITY PRESS (VIRTUAL PUBLISHING)
FOR AND ON BEHALF OF THE PRESS SYNDICATE OF THE UNIVERSITY OF
CAMBRIDGE
The Pitt Building, Trumpington Street, Cambridge CB2 IRP
40 West 20th Street, New York, NY 10011-4211, USA
477 Williamstown Road, Port Melbourne, VIC 3207, Australia
http://www.cambridge.org
Introduction page xi
2 Topological Dynamics 28
2.1 Limit Sets and Recurrence 28
2.2 Topological Transitivity 31
2.3 Topological Mixing 33
2.4 Expansiveness 35
2.5 Topological Entropy 36
2.6 Topological Entropy for Some Examples 41
2.7 Equicontinuity, Distality, and Proximality 45
2.8 Applications of Topological Recurrence to
Ramsey Theory 48
vii
viii Contents
3 Symbolic Dynamics 54
3.1 Subshifts and Codes 55
3.2 Subshifts of Finite Type 56
3.3 The Perron–Frobenius Theorem 57
3.4 Topological Entropy and the Zeta Function of an SFT 60
3.5 Strong Shift Equivalence and Shift Equivalence 62
3.6 Substitutions 64
3.7 Sofic Shifts 66
3.8 Data Storage 67
4 Ergodic Theory 69
4.1 Measure-Theory Preliminaries 69
4.2 Recurrence 71
4.3 Ergodicity and Mixing 73
4.4 Examples 77
4.5 Ergodic Theorems 80
4.6 Invariant Measures for Continuous Maps 85
4.7 Unique Ergodicity and Weyl’s Theorem 87
4.8 The Gauss Transformation Revisited 90
4.9 Discrete Spectrum 94
4.10 Weak Mixing 97
4.11 Applications of Measure-Theoretic Recurrence
to Number Theory 101
4.12 Internet Search 103
Bibliography 225
Index 231
Introduction
Ken Berg, Mike Boyle, Boris Hasselblatt, Michael Jakobson, Anatole Katok,
Michal Misiurewicz, and Dan Rudolph. We thank them for their contribu-
tions. We are especially grateful to Vitaly Bergelson for his contributions to
the treatment of applications of topological dynamics and ergodic theory to
combinatorial number theory. We thank the students who used early ver-
sions of this book in our classes, and who found many typos, errors, and
omissions.
CHAPTER ONE
1
2 1. Examples and Basic Concepts
that is preserved by the map f . For example, (X, f ) could be a measure space
and a measure-preserving map; a topological space and a continuous map;
a metric space and an isometry; or a smooth manifold and a differentiable
map.
A continuous-time dynamical system consists of a space X and a one-
parameter family of maps of { f t : X → X}, t ∈ R or t ∈ R+ 0 , that forms a one-
parameter group or semigroup, i.e., f t+s = f t ◦ f s and f 0 = Id. The dynam-
ical system is called a flow if the time t ranges over R, and a semiflow if t
ranges over R+ t
0 . For a flow, the time-t map f is invertible, since f
−t
= ( f t )−1 .
Note that for a fixed t0 , the iterates ( f ) = f form a discrete-time dynam-
t0 n t0 n
ical system.
We will use the term dynamical system to refer to either discrete-time
or continuous-time dynamical systems. Most concepts and results in dy-
namical systems have both discrete-time and continuous-time versions. The
continuous-time version can often be deduced from the discrete-time ver-
sion. In this book, we focus mainly on discrete-time dynamical systems, where
the results are usually easier to formulate and prove.
To avoid having to define basic terminology in four different cases, we
write the elements of a dynamical system as f t , where t ranges over Z, N0 , R,
or R+ , as appropriate. For x ∈ X, we define the positive semiorbit O+f (x) =
0 t
f (x). In the invertible case, we define the negative semiorbit O−f (x) =
t≥0 t + − t
t≤0 f (x), and the orbit O f (x) = O f (x) ∪ O f (x) = t f (x) (we omit the
subscript “ f ” if the context is clear). A point x ∈ X is a periodic point of
period T > 0 if f T (x) = x. The orbit of a periodic point is called a periodic
orbit. If f t (x) = x for all t, then x is a fixed point. If x is periodic, but not
fixed, then the smallest positive T, such that f T (x) = x, is called the minimal
period of x. If f s (x) is periodic for some s > 0, we say that x is eventu-
ally periodic. In invertible dynamical systems, eventually periodic points are
periodic.
For a subset A ⊂ X and t > 0, let f t (A) be the image of Aunder f t , and let
f (A) be the preimage under f t , i.e., f −t (A) = ( f t )−1 (A) = {x ∈ X: f t (x) ∈
−t
Rα x = x + α mod 1.
Lh g = hg and Rh g = gh.
∞
xi
φ:
→ [0, 1], φ((xi )i∈N ) = .
i=1
mi
ments of the set {0, . . . , m − 1}. A subset A ⊂ [0, 1] is dense if and only if
every finite sequence w ∈ Fm occurs at the beginning of the base-m expan-
sion of some element of A. It follows that the set of periodic points is dense
in S1 . The orbit of a point x = 0.x1 x2 . . . is dense in S1 if and only if every
finite sequence from Fm appears in the sequence (xi ). Since Fm is countable,
we can construct such a point by concatenating all elements of Fm.
Although φ is not one-to-one, we can construct a right inverse to φ. Con-
sider the partition of S1 = [0, 1] / ∼ into intervals
Pk = [k/m, (k + 1)/m), 0 ≤ k ≤ m − 1.
−1
Exercise 1.3.1. Prove that λ(Em ([a, b])) = λ([a, b]) for any interval
[a, b] ⊂ [0, 1].
Exercise 1.3.3. Show that the set of points with dense orbits is uncountable.
*Exercise 1.3.5. Show that the set of points with dense orbits under Em has
Lebesgue measure 1.
where n1 < n2 < · · · < nk are indices in Z or N, and ji ∈ Am. Since the preim-
+
age of a cylinder is a cylinder, σ is continuous on
m and is a homeomorphism
of
m. The metric
1 1
1 2 1 1
1 1 0
1 0 1
1 2 3
0 0 1
Figure 1.1. Examples of directed graphs with labeled vertices and the corresponding
adjacency matrices.
+
generates the product topology on
m and
m (Exercise 1.4.3). In
m,
the open ball B(x, 2 ) is the symmetric cylinder Cx−l,−l+1,...,l
−l
−l ,x−l+1 ,...,xl
, and in
m +
,
−l +
B(x, 2 ) = Cx1 ,...,xl . The shift is expanding on
m ; if d(x, x ) < 1/2, then
1,...,l
Exercise 1.4.2. Let Abe an m × m matrix of zeros and ones. Prove that:
(a) the number of fixed points in
A (or
+A ) is the trace of A;
(b) the number of allowed words of length n + 1 beginning with the sym-
bol i and ending with j is the i, jth entry of An ; and
(c) the number of periodic points of period n in
A (or
+ A ) is the trace
of An .
+
Exercise 1.4.3. Verify that the metrics on
m and
m generate the product
topology.
Exercise 1.4.5. Assume that all entries of some power of A are positive.
Show that in the product topology on
A and
+
A , periodic points are dense,
and there are dense orbits.
Figure 1.2 shows the graph of q3 and successive images xi = q3i (x0 ) of a point
x0 .
If µ > 1 and x ∈
/ [0, 1], then qµn (x) → −∞ as n → ∞. For this reason, we
focus our attention on the interval [0, 1]. For µ ∈ [0, 4], the interval
[0, 1]
is forward
invariant under qµ . For µ > 4, the interval (1/2 − 1/4 − 1/µ,
1/2 + 1/4 − 1/µ) maps outside [0, 1]; we show in Chapter 7 that the set of
points µ whose forward orbits stay in [0, 1] is a Cantor set, and (µ , qµ ) is
equivalent to the full one-sided shift on two symbols.
Let X be a locally compact metric space and f : X → X a continuous
map. A fixed point p of f is attracting if it has a neighborhood U such that Ū
is compact, f (Ū) ⊂ U, and n≥0 f n (U) = { p}. A fixed point p is repelling
10 1. Examples and Basic Concepts
0.8
0.6
0.4
0.2
2
x0 q3(x0) q 3(x0)
Exercise 1.5.3. Suppose p is an attracting fixed point for f . Show that there
is a neighborhood U of p such that the forward orbit of every point in U
converges to p.
was studied by C. Gauss, and is now called the Gauss transformation. Note
that ϕ maps each interval (1/(n + 1), 1/n] continuously and monotonically
onto [0, 1); it is discontinuous at 1/n for all n ∈ N. Figure 1.3 shows the graph
of ϕ.
0.8
0.6
0.4
0.2
(1, 1)
(2, 1)
(0, 0)
(3, 0)
√
by 1/λ in the direction of v1/λ = ((1 − 5)/2, 1). The eigenvectors are per-
pendicular because A is symmetric.
Since A has integer entries, it preserves the integer lattice Z2 ⊂ R2 and
induces a map (which we also call A) of the torus T2 = R2 /Z2 . The torus
can be viewed as the unit square [0, 1] × [0, 1] with opposite sides identified:
(x1 , 0) ∼ (x1 , 1) and (0, x2 ) ∼ (1, x2 ), x1 , x2 ∈ [0, 1]. The map A is given in
coordinates by
x1 (2x1 + x2 ) mod 1
A =
x2 (x1 + x2 ) mod 1
Exercise 1.7.4. Show that the number of fixed points of a hyperbolic toral
automorphism A is det(A − I) (hence the number of periodic points of
period n is det(An − I)).
D1 D2 D3 D4 D5 p
R01
R0
R00
R10
R1
R11
R
Figure 1.6. Horizontal rectangles.
of width µ−1 . For any sequence ω−m, ω−m+1 , . . . , ω−1 of zeros and ones,
m −i −m
∞ −i
i=1 f (Rωi ) is a vertical rectangle of width µ , and H− = i=1 f (R)
is the product of a vertical interval (of length 1) and a horizontal Cantor
set C − .
∞
The horseshoe set H = H+ ∩ H− = i=−∞ f i (R) is the product of the
Cantor sets C − and C + and is closed and f -invariant. It is locally maximal,
i.e., there is an open set U containing H such that any f -invariant subset of U
containing H coincides with H (Exercise 1.8.2). The map φ:
2 = {0, 1}Z →
H that assigns to each infinite sequence ω = (ωi ) ∈
2 the unique point
φ(ω) = ∞ i
−∞ f (Rωi ) is a bijection (Exercise 1.8.3). Note that
∞
f (φ(ω)) = f i+1 Rωi = φ(σr (ω)),
−∞
1.9. The Solenoid 17
Exercise 1.8.3. Prove that φ is a bijection, and that both φ and φ −1 are
continuous.
consists of 2n disks of radius λn : two disks inside each of the 2n−1 disks of
F n−1 (T ) ∩ {φ = const}. Slices of F(T ), F 2 (T ), and F 3 (T ) for φ = 1/8 are
shown in Figure 1.8.
The set S = ∞ n
n=0 F (T ) is called a solenoid. It is a closed F-invariant
subset of T on which F is bijective (Exercise 1.9.1). It can be shown that S is
locally the product of an interval with a Cantor set in the two-dimensional
disk.
The solenoid is an attractor for F. In fact, any neighborhood of S con-
tains F n (T ) for n sufficiently large, so the forward orbit of every point in
T converges to S. Moreover, S is a hyperbolic set, and is therefore called a
hyperbolic attractor. We give a precise definition of attractors in §1.13.
∞
Let denote the set of sequences (φi )i=0 , where φi ∈ S1 and φi =
2φi+1 mod 1 for all i. The product topology on (S1 )N0 induces the subspace
topology on . The space is a commutative group under component-wise
addition (mod 1). The map (φ, ψ) →φ − ψ is continuous, so is a topo-
logical group. The map α: → , (φ0 , φ1 , . . .) →(2φ0 , φ0 , φ1 , . . .) is a group
automorphism and a homeomorphism (Exercise 1.9.3).
For s ∈ S, the first (angular) coordinates of the preimages F −n (s) =
−
(φn , xn , yn ) form a sequence h(s) = (φ0 , φ1 , . . .) ∈ . This defines a map
h: S → . The inverse of h is the map (φ0 , φ1 , . . .) → ∞ n=0 F ({φn } × D ),
n 2
that h is a homeomorphism.
Exercise 1.9.3. Show that is a topological group, and α is an automor-
phism and homeomorphism.
Exercise 1.9.4. Find the fixed point of F and all periodic points of period 2.
The energy E of the system is the sum of the kinetic and potential energies,
E(x, y) = 1 − cos x + y2 /2. One can show (Exercise 1.10.2) by differentiating
E(x, y) with respect to t that E is constant along solutions of the differential
equation. Equivalently, if f t is the flow in R2 of this differential equation, then
E is invariant by the flow, i.e., E( f t (x, y)) = E(x, y), for all t ∈ R, (x, y) ∈ R2 .
A function that is constant on the orbits of a dynamical system is called a
first integral of the system.
The fixed points in the phase plane for the undamped pendulum are
(kπ, 0), k ∈ Z. The points (2kπ, 0) are local minima of the energy. The points
(2(k + 1)π, 0) are saddle points.
Now consider the damped pendulum θ̈ + γ θ̇ + sin θ = 0, or the equi-
valent system
ẋ = y,
ẏ = − sin x − γ y.
A simple calculation shows that Ė < 0 except at the fixed points (kπ, 0), k ∈
Z, which are the local extrema of the energy. Thus the energy is strictly
decreasing along every non-constant solution. In particular, every trajec-
tory approaches a critical point of the energy, and almost every trajectory
approaches a local minimum.
The energy of the pendulum is an example of a Lyapunov function, i.e.,
a continuous function that is non-increasing along the orbits of the flow.
Any strict local minimum of a Lyapunov function is an asymptotically sta-
ble equilibrium point of the differential equation. Moreover, any bounded
orbit must converge to the maximal invariant subset M of the set of points
satisfying Ė = 0. In the case of the damped pendulum, M = {(kπ, 0): k ∈ Z}.
Here is another class of examples that appears frequently in applications,
particularly optimization problems. Given a smooth function f : Rn → R,
the flow of the differential equation
ẋ = grad f (x)
In other words, flow along {x} × R+ to (x, c(x)), then jump to ( f (x), 0) and
continue along { f (x)} × R+ , and so on. See Figure 1.9. A suspension flow is
also called a flow under a function.
22 1. Examples and Basic Concepts
ψ t (a)
R+
(f (x), c(f (x)))
a
(x, c(x))
g(a)
A
(x, 0) (f (x), 0) X
Note that φαt is the suspension of the circle rotation Rα with ceiling function 1,
and S1 × {0} is a cross-section for φα with constant return time τ (y) = 1 and
first return map Rα . The flow φαt consists of left translations by the elements
g t = (αt, t) mod 1, which form a one-parameter subgroup of T2 .
∗
Exercise 1.11.3. Suppose 1, s, and αs are real numbers that are linearly
independent over Q. Show that every orbit of the time s map φαs is dense in
T 2.
factor of m if d(x, y) ≤ 1/(2m), so any two points are moved at least 1/2m
apart by some iterate of Em, so Em has sensitive dependence on initial con-
ditions. A typical orbit is dense (§1.3) and is uniformly distributed on the
circle (Proposition 4.4.2).
The simplest non-linear chaotic dynamical systems in dimension one are
the quadratic maps qµ (x) = µx(1 − x), µ ≥ 4, restricted to the forward in-
variant set µ ⊂ [0, 1] (see §1.5 and Chapter 7).
Sensitive dependence on initial conditions is usually associated with
positive Lyapunov exponents. Let f be a differentiable map of an open
subset U ⊂ Rm into itself, and let d f (x) denote the derivative of f at x. For
24 1. Examples and Basic Concepts
If f has uniformly bounded first derivatives, then χ is well defined for every
x ∈ U and every non-zero vector v.
The Lyapunov exponent measures the exponential growth rate of tangent
vectors along orbits, and has the following properties:
Exercise 1.12.3. Compute the Lyapunov exponents for the solenoid, §1.9.
1.13. Attractors 25
Exercise
√ 1.12.4. Using a computer, calculate the first 100 points in the orbit
of 2 − 1 under the map E2 . What proportion of these points is contained
in each of the intervals [0, 14 ), [ 14 , 12 ), [ 12 , 34 ), and [ 34 , 1)?
1.13 Attractors
Let X be a compact topological space, and f : X → X be a continuous map.
Generalizing the notion of an attracting fixed point, we say that a compact
set C ⊂ X is an attractor if there is an open set U containing C, such that
f (Ū) ⊂ U and C = n≥0 f n (U). It follows that f (C) = C, since f (C) =
n≥1 f (U) ⊂ C; on the other hand, C = n≥1 f (U) = f (C), since f (U) ⊂
n n
ẋ = σ (y − x),
ẏ = Rx − y − xz, (1.2)
ż = −bz + xy,
f (x, y) = a − by − x 2 ,
g(x, y) = x,
50
40
30
25
20 20
15
10
5
10 0
-5 y
-20 -15 -10
-10 -5 -15
0 5 -20
10 15 -25
x 20
-1
-2
-2 -1 0 1 2
Figure 1.11. Hénon attractor.
Topological Dynamics
PROPOSITION 2.1.1
1. Let f be a homeomorphism, y ∈ O(x), and z ∈ O(y). Then z ∈ O(x).
2. Let f be a continuous map, y ∈ O+ (x), and z ∈ O+ (y). Then z ∈ O+ (x).
Proof. Exercise 2.1.7.
Let X be compact. A closed, non-empty, forward f -invariant subset Y ⊂
X is a minimal set for f if it contains no proper, closed, non-empty, forward
f -invariant subset. A compact invariant set Y is minimal if and only if the for-
ward orbit of every point in Y is dense in Y (Exercise 2.1.4). Note that a peri-
odic orbit is a minimal set. If X itself is a minimal set, we say that f is minimal.
Exercise 2.1.1. Show that the α- and ω-limit sets of a point are closed in-
variant sets.
Exercise 2.1.5. Show that there are points that are non-recurrent and not
eventually periodic for an expanding circle endomorphism Em.
the Baire category theorem. The forward orbit of any point y ∈ Y enters
each Vi , hence is dense in X.
In most topological spaces, existence of a dense full orbit for a home-
omorphism implies existence of a dense forward orbit, as we show in the
next proposition. Note, however, that density of a particular full orbit O(x)
does not imply density of the corresponding forward orbit O+ (x) (see
Exercise 2.2.2).
PROPOSITION 2.2.2. Let f : X → X be a homeomorphism of a compact
metric space, and suppose that X has no isolated points. If there is a dense full
orbit O(x), then there is a dense forward orbit O+ (y).
Proof. Since O(x) = X, the orbit O(x) visits every non-empty open set U
at least once, and therefore infinitely many times because X has no isolated
points. Hence there is a sequence nk, with |nk| → ∞, such that f nk (x) ∈
B(x, 1/k) for k ∈ N, i.e., f nk (x) → x as k → ∞. Thus, f nk+l (x) → f l (x) for
any l ∈ Z. There are either infinitely many positive or infinitely many neg-
ative indices nk, and it follows that either O(x) ⊂ O+ (x) or O(x) ⊂ O− (x).
In the former case, O+ (x) = X, and we are done. In the latter case, let
U, V be non-empty open sets. Since O− (x) = X, there are integers i <
j < 0 such that f i (x) ∈ U and f j (x) ∈ V, so f j−i (U) ∩ V =. ∅Hence, by
Proposition 2.2.1, f is topologically transitive.
Exercise 2.2.1. Show that if X has no isolated points and O+ (x) is dense,
then ω(x) is dense. Give an example to show that this is not true if X has
isolated points.
Exercise 2.2.2. Give an example of a dynamical system with a dense full
orbit but no dense forward orbit.
Exercise 2.2.3. Is the product of two topologically transitive systems topo-
logically transitive? Is a factor of a topologically transitive system topologi-
cally transitive?
Exercise 2.2.4. Let f : X → X be a homeomorphism. Show that if f has
a non-constant first integral or Lyapunov function (§1.10), then it is not
topologically transitive.
Exercise 2.2.5. Let f : X → X be a topological dynamical system with at
least two orbits. Show that if f has an attracting periodic point, then it is not
topologically transitive.
Exercise 2.2.6. Let α be irrational and f : T 2 → T 2 be the homeomor-
phism of the 2-torus given by f (x, y) = (x + α, x + y).
2.3. Topological Mixing 33
(a) Prove that every non-empty, open, f -invariant set is dense, i.e., f is
topologically transitive.
(b) Suppose the forward orbit of (x0 , y0 ) is dense. Prove that for ev-
ery y ∈ S1 the forward orbit of (x0 , y) is dense. Moreover, if the set
n n
k=0 f (x0 , y0 ) is -dense, then k=0 f (x0 , y) is -dense.
k k
show that for any two balls B(ω, 2−l1 ) and B(ω , 2−l2 ), there is N > 0 such
that σ n B(ω, 2−l1 ) ∩ B(ω , 2−l2 ) = for
∅ n ≥ N. Elements of σ n B(ω, 2−l1 ) are
sequences with specified values in the places −n − l1 , . . . , −n + l1 . Therefore
the intersection is non-empty when −n + l1 < −l2 , i.e., n ≥ N = l1 + l2 + 1.
+
This proves that (m, σ ) is topologically mixing; the proof for (m , σ ) is an
exercise (Exercise 2.3.4).
COROLLARY 2.3.3. The horseshoe (H, f ) (§1.8) is topologically mixing.
Proof. The horseshoe (H, f ) is topologically conjugate to the full two-shift
(2 , σ ) (see Exercise 1.8.3).
PROPOSITION 2.3.4. The solenoid (S, F) is topologically mixing.
Proof. Recall (Exercise 1.9.2) that (S, F) is topologically conjugate to
(
, α), where
∞
= {(φi ): φi ∈ S1 , φi = 2φi+1 , ∀i} ⊂ S1 = T∞ ,
i=0
α
and (φ0 , φ1 , φ2 , . . .) →(2φ0 , φ0 , φ1 , . . .). Thus, it suffices to show that (
, α)
is topologically mixing. The topology in T∞ has a basis consisting of open
∞
sets i=0 Ik, where the I j s are open in S1 and all but finitely many are equal
to S1 . Let U = (I0 × I1 × · · · × Ik × S1 × S1 × · · ·) ∩
and V = (J0 × J1 ×
· · · × Jl × S1 × S1 × · · ·) ∩
be non-empty open sets from this basis. Choose
m > 0 so that 2m I0 = S1 . Then for n > m + l, the first n − m components of
are S1 , so α n (U) ∩ V =. ∅
Exercise 2.3.1. Show that a circle rotation is not topologically mixing. Show
that an isometry is not topologically mixing if there is more than one point
in the space.
+
Exercise 2.3.4. Prove that (m , σ ) is topologically mixing.
2.4. Expansiveness 35
2.4 Expansiveness
A homeomorphism f : X → X is expansive if there is δ > 0 such that for any
two distinct points x, y ∈ X, there is some n ∈ Z such that d( f n (x), f n (y)) ≥
δ. A non-invertible continuous map f : X → X is positively expansive if there
is δ > 0 such that for any two distinct points x, y ∈ X, there is some n ≥ 0
such that d( f n (x), f n (y)) ≥ δ. Any number δ > 0 with this property is called
an expansiveness constant for f .
Among the examples from Chapter 1, the following are expansive (or
positively expansive): the circle endomorphisms Em, |m| ≥ 2; the full and
one-sided shifts; the hyperbolic toral automorphisms; the horseshoe; and
the solenoid (Exercise 2.4.2). For sufficiently large values of the parameter
µ, the quadratic map qµ is expansive on the invariant set µ . Circle rotations,
group translations, and other equicontinuous homeomorphisms (see §2.7)
are not expansive.
PROPOSITION 2.4.1. Let f be a homeomorphism of an infinite compact
metric space X. Then for every > 0 there are distinct points x0 , y0 ∈ X such
that d( f n (x0 ), f n (y0 )) ≤ for all n ∈ N0 .
Proof [Kin90]. Fix > 0. Let E be the set of natural numbers m for which
there is a pair x, y ∈ X such that
d(x, y) ≥ and d( f n (x), f n (y)) ≤ for n = 1, . . . , m. (2.1)
Let M = sup E if E = ∅, and M = 0 if E = ∅.
If M = ∞, then for every m ∈ N there is a pair xm, ym satisfying (2.1). By
compactness, there is a sequence mk → ∞ such that the limits
lim xmk = x , lim ymk = y
k→∞ k→∞
a j = x0 and b j = y0
for infinitely many positive j and hence d( f n (x0 ), f n (y0 )) < for all n ≥ 0.
Proposition 2.4.1 is also true for non-invertible maps (Exercise 2.4.3).
Exercise 2.4.2. Show that the expanding circle endomorphisms Em, |m| ≥
2, the full one- and two-sided shifts, the hyperbolic toral automorphisms,
the horseshoe, and the solenoid are expansive, and compute expansiveness
constants for each.
Exercise 2.4.3. Show that Proposition 2.4.1 is true for non-invertible con-
tinuous maps of infinite metric spaces.
measures the maximum distance between the first n iterates of x and y. Each
dn is a metric on X, dn ≥ dn−1 , and d1 = d. Moreover, the di are all equivalent
metrics in the sense that they induce the same topology on X(Exercise 2.5.1).
Fix > 0. A subset A ⊂ X is (n, )-spanning if for every x ∈ X there is
y ∈ A such that dn (x, y) < . By compactness, there are finite (n, )-spanning
sets. Let span(n, , f ) be the minimum cardinality of an (n, )-spanning set.
A subset A ⊂ X is (n, )-separated if any two distinct points in A are at
least apart in the metric dn . Any (n, )-separated set is finite. Let sep(n, , f )
be the maximum cardinality of an (n, )-separated set.
Let cov(n, , f ) be the minimum cardinality of a covering of X by sets of
dn -diameter less than (the diameter of a set is the supremum of distances
between pairs of points in the set). Again, by compactness, cov(n, , f ) is
finite.
The quantities span(n, , f ), sep(n, , f ), and cov(n, , f ) count the num-
ber of orbit segments of length n that are distinguishable at scale . These
quantities are related by the following lemma.
htop = h( f ) = lim+ h ( f )
→0
sep(n, , f ), i.e.,
1
h( f ) = lim+ lim log(span(n, , f )) (2.3)
→0 n
n→∞
1
= lim+ lim log(sep(n, , f )). (2.4)
→0 n→∞ n
1
h( f ) = lim+ lim log(cov(n, , f )) (2.5)
→0 n→∞ n
1
= lim+ lim log(span(n, , f )) (2.6)
→0 n→∞ n
1
= lim+ lim log(sep(n, , f )). (2.7)
→0 n→∞ n
1 1
lim lim log(cov (n, δ, f )) ≤ lim+ lim log(cov(n, , f )).
δ→0+ n→∞ n →0 n→∞ n
h( f ) = max h( f |Ai ).
1≤i≤k
Therefore,
1 1 1
lim log (span(n, , f )) ≤ lim log k + lim log max spani (n, , f )
n→∞ n n→∞ n n→∞ n 1≤i≤k
1
= 0 + max lim log (spani (n, , f ))
1≤i≤k n→∞ n
Proof. Fix γ and with 0 < γ < < δ. We will show that h2γ ( f ) = h ( f ).
By monotonicity, it suffices to show that h2γ ( f ) ≤ h ( f ).
By expansiveness, for distinct points x and y, there is some i ∈ Z such
that d( f i (x), f i (y)) ≥ δ > . Since the set {(x, y) ∈ X × X: d(x, y) ≥ γ }
is compact, there is k = k(γ , ) ∈ N such that if d(x, y) ≥ γ , then
2.6. Topological Entropy for Some Examples 41
d( f i (x), f i (y)) > for some |i| ≤ k. Thus if A is an (n, γ )-separated set,
then f −k(A) is (n + 2k, )-separated. Hence, by Lemma 2.5.1, h ( f ) ≥
h2γ ( f ).
REMARK 2.5.8. The topological entropy of a continuous (semi)flow can be
defined as the entropy of the time-1 map. Alternatively, it can be defined using
the analog dT , T > 0, of the metrics dn . The two definitions are equivalent
because of the equicontinuity of the family of time-t maps, t ∈ [0, 1].
Exercise 2.5.1. Let (X, d) be a compact metric space. Show that the metrics
di all induce the same topology on X.
Exercise 2.5.2. Prove the remaining inequalities in Lemma 2.5.1.
Exercise 2.5.3. Let {an } be a subadditive sequence of non-negative real
numbers, i.e., 0 ≤ am+n ≤ am + an for all m, n ≥ 0. Show that limn→∞ an /n =
infn≥0 an /n.
Exercise 2.5.4. Show that the topological entropy of an isometry is zero.
Exercise 2.5.5. Let g: Y → Y be a factor of f : X → X. Prove that h( f ) ≥
h(g).
Exercise 2.5.6. Let Y and Z be compact metric spaces, X = Y × Z, and
π be the projection to Y. Suppose f : X → X is an isometric extension
of a continuous map g: Y → Y, i.e., π ◦ f = g ◦ π and d( f (x1 ), f (x2 )) =
d((x1 ), (x2 )) for all x1 , x2 ∈ Y with π (x1 ) = π(x2 ). Prove that h( f ) = h(g).
Exercise 2.5.7. Prove that the topological entropy of a continuously differ-
entiable map of a compact manifold is finite.
Proof. It suffices to prove the lemma for a Jordan block. Thus without loss
of generality, we assume that B has λs on the diagonal, ones above and zeros
elsewhere. In this setting, Vλ = Ck and in the standard basis e1 , . . . , ek, we
have Be1 = λe1 and Bei = λei + ei−1 , for i = 2, . . . , k. For δ > 0, consider the
basis e1 , δe2 , δ 2 e3 , . . . , δ k−1 ek. In this basis, the linear map B is represented by
the matrix
λ δ
λ δ
. .
Bδ = .. .. .
λ δ
λ
Observe that Bδ = λI + δ A with A ≤ 1, where A = supv=0Av|/v.
Therefore
(|λ| − δ)n v ≤ Bδn v ≤ (|λ| + δ)n v.
Since Bδ is conjugate to B, there is a constant C(δ) > 0 that bounds the
distortion of the change of basis.
LEMMA 2.6.3. Let B be a k × k real matrix and λ an eigenvalue of B. Then
for every δ > 0 there is C(δ) > 0 such that
C(δ)−1 (|λ| − δ)n v ≤ Bn v ≤ C(δ)(|λ| + δ)n v
for every n ∈ N and every v ∈ Vλ (if λ ∈ R) or every v ∈ Vλ,λ̄ (if λ ∈
/ R).
Proof. If λ is real, then the result follows from Lemma 2.6.2. If λ is complex,
then the estimates for Vλ and Vλ̄ from Lemma 2.6.2 imply a similar estimate
for Vλ, λ̄ , with a new constant C(δ) depending on the angle between Vλ and
Vλ̄ and the constants in the estimates for Vλ and Vλ̄ (since |λ| = |λ̄|).
PROPOSITION 2.6.4. Let à be a k × k integer matrix with determinant 1
and with eigenvalues α1 , . . . , αk, where
|α1 | ≥ |α2 | ≥ · · · ≥ |αm| > 1 > |αm+1 | ≥ · · · ≥ |αk|.
Let A: T → Tk be the associated hyperbolic toral automorphism. Then
k
m
h(A) = log |αi |.
i=1
∞ m j j
2 φi − 2mψi k
2mφi − ψi 1
d(α φ, α ψ ) =
m m j
i
< i
+ k
i=0
2 i=0
2 2
k
2k−i 2−(n+2k+1) 1 1
< 2m + < k−1 < .
i=0
2i 2k 2
1
h(α) ≤ lim log cardAn = log 2.
n→∞ n
2.7. Equicontinuity, Distality, and Proximality 45
Exercise 2.6.2. Compute the topological entropy of the full one- and two-
sided m-shifts.
x →x + α mod 1,
y →x + y mod 1.
We view T2 as the unit square with opposite sides identified and use the
metric inherited from the Euclidean metric. To see that this map is distal, let
(x, y), (x , y ) be distinct points in T2 . If x = x , then d(F n (x, y), F n (x , y ))
is at least d((x, 0), (x , 0)), which is constant. If x = x , then d(F n (x, y),
(z, z) ∈ O(x, y). Recall that O(x, y) is minimal (Proposition 2.1.3). Since
(x, y) ∈/ O(z, z), we obtain a contradiction.
COROLLARY 2.7.7. A factor of a distal homeomorphism f of a compact
Hausdorff space X is distal.
Proof. Let g: Y → Y be a factor of f . Then f × f is pointwise almost pe-
riodic by Proposition 2.7.6. Since (g × g) is a factor of f × f , it is pointwise
almost periodic (Exercise 2.7.5), and hence is distal.
The class of distal dynamical systems is of special interest because it is
closed under factors and isometric extensions. The class of minimal distal sys-
tems is the smallest such class of minimal systems: According to Furstenberg’s
structure theorem [Fur63], every minimal distal homeomorphism (or flow)
can be obtained by a (possibly transfinite) sequence of isometric extensions
starting with the one-point dynamical system.
Exercise 2.7.2. Prove the equivalence of the topological and metric defini-
tions of distal and proximal points at the beginning of this section.
Proof [Ber00]. Since G acts minimally, and X is compact, there are ele-
m
ments g1 , . . . , gm ∈ G such that i=1 gi (U) = X (Exercise 2.8.2).
We argue by induction on n. For n = 1, let T be an IP-system and U ⊂ X
be open and not empty. Set V0 = U. Define recursively Vk = T{k} (Vk−1 ) ∩
gik (U), where i k is chosen so that 1 ≤ i k ≤ m and T{k} (Vk−1 ) ∩ gik (U) =. ∅ By
−1
construction, T{k} (Vk) ⊂ Vk−1 and Vk ⊂ gik (U). In particular, by the pigeon-
hole principle, there are 1 ≤ i ≤ m and arbitrarily large p < q such that
Vp ∪ Vq ⊂ gi (U). Choose p so that β = { p + 1, p + 2, . . . , q} > α. Then the
set W = gi−1 (Vq ) ⊂ U is not empty, and
−1
Tβ−1 (W) = gi−1 T{−1 −1
p+1} · · · T{q} (Vq ) ⊂ gi
−1
T{ p+1} (Vp+1 ) ⊂ gi−1 (Vp ) ⊂ U.
Therefore, U ∩ Tβ (U) ⊃ W =. ∅
Assume that the theorem holds true for any n IP-systems in G. Let U ⊂ X
be open and not empty. Let T (1) , . . . , T (n+1) be IP-systems in G. We will
construct a sequence of non-empty open subsets Vk ⊂ X and an increasing
( j) −1
sequence αk ∈ F, αk > α, such that V0 = U, n+1 j=1 (Tαk ) (Vk) ⊂ Vk−1 , and
Vk ⊂ gik (U) for some 1 ≤ i k ≤ m.
By the inductive assumption applied to V0 = U and the n IP-systems
(T (n+1) )−1 T ( j) , j = 1, . . . , n, there is α1 > α such that
−1 (1) −1 (n)
V0 ∩ Tα(n+1)1
Tα1 (V0 ) ∩ · · · ∩ Tα(n+1)
1
Tα1 (V0 ) =. ∅
(n+1)
Apply Tα1 and, for an appropriate 1 ≤ i 1 ≤ m, set
If Vk−1 and αk−1 have been constructed, apply the inductive assumption to
Vk−1 and the IP-systems (T (n+1) )−1 T ( j) , j = 1, . . . , n, to get αk > αk−1 such
that
−1 (1) −1 (n)
Vk−1 ∩ Tα(n+1)
k
Tαk (Vk−1 ) ∩ · · · ∩ Tα(n+1)
k
Tαk (Vk−1 ) =. ∅
(n+1)
Apply Tαk and, for an appropriate 1 ≤ i k ≤ m, set
THEOREM 2.8.5. Let d ∈ N, and let A be a finite subset of Zd . Then for each
finite partition Zd = mk=1 Sk, there are k ∈ {1, . . . , m}, z0 ∈ Z , and n ∈ N such
d
Exercise 2.8.4. Prove that van der Waerden’s Theorem 2.8.1 is equivalent
to the following finite version: For each m, n ∈ N there is k ∈ N such that if
the set {1, 2, . . . , k} is partitioned into m subsets, then one of them contains
an arithmetic progression of length n.
*Exercise 2.8.5. For z ∈ Zd , the translation by z in Zd induces a homeomor-
phism (shift) Tz in = {1, . . . , m}Z . Prove Theorem 2.8.5 by considering the
d
Symbolic Dynamics
+
In §1.4, we introduced the symbolic dynamical systems (m, σ ) and (m , σ ),
and we showed by example throughout Chapter 1 how these shift spaces arise
naturally in the study of other dynamical systems. In all of those examples,
we encoded an orbit of the dynamical system by its itinerary through a finite
collection of disjoint subsets. Specifically, following an idea that goes back
to J. Hadamard, suppose f : X → X is a discrete dynamical system. Con-
sider a partition P = {P1 , P2 , . . . , Pm} of X, i.e., P1 ∪ P2 ∪ · · · ∪ Pm = X and
Pi ∩ Pj = ∅ for i
= j. For each x ∈ X, let ψi (x) be the index of the element
of P containing f (x). The sequence (ψi (x))i∈N0 is called the itinerary of x.
i
+
which satisfies ψ ◦ f = σ ◦ ψ. The space m is totally disconnected, and the
map ψ usually is not continuous. If f is invertible, then positive and negative
iterates of f define a similar map X → m = {1, 2, . . . , m}Z . The image of
+
ψ in m or m is shift-invariant, and ψ semiconjugates f to the shift on
the image of ψ. The indices ψi (x) are symbols – hence the name symbolic
dynamics. Any finite set can serve as the symbol set, or alphabet, of a symbolic
dynamical system. Throughout this chapter, we identify every finite alphabet
with {1, 2, . . . , m}.
Recall that the cylinder sets
,...,nk
C nj11,..., jk = ω = (ωl ) : ωni = ji , i = 1, . . . , k ,
+
form a basis for the product topology of m and m , and that the metric
1 The exposition of this section as well as §3.2, §3.4, and §3.5 follows in part the lectures of
M. Boyle [Boy93].
56 3. Symbolic Dynamics
The last proposition implies that “the future is independent of the past” in
an SFT; i.e., with appropriate one-step coding, if the sequences . . . x−2 x−1 x0
and x0 x1 x2 . . . are allowed, then . . . x−2 x−1 x0 x1 x2 . . . is allowed.
Exercise 3.2.1. Show that the collection of all isomorphism classes of sub-
shifts of finite type is countable.
∗
Exercise 3.2.2. Show that the collection of all subshifts of 2 is uncount-
able.
Exercise 3.2.3. Show that every edge shift is an SFT.
Exercise 3.2.4. Show that every edge shift is naturally isomorphic to a ver-
tex shift. What are the vertices?
Exercise 3.3.1. Show that if A is a primitive integral matrix, then the edge
shift eA is topologically mixing.
Exercise 3.3.3. This exercise outlines the main steps in the proof of
Theorem 3.3.4. Let A be a non-negative irreducible matrix, and let B be
the matrix with entries bi j = 0 if ai j = 0 and bi j = 1 if ai j > 0. Let be the
graph whose adjacency matrix is B. For a vertex v in , let d = d(v) be the
greatest common divisor of the lengths of closed paths in starting from v.
Let Vk, k = 0, 1, . . . , d − 1, be the set of vertices of that can be connected
60 3. Symbolic Dynamics
∞
1
ζ f (z) = exp |Fix( f n )|zn .
n=1
n
where the product is taken over all periodic orbits γ of f , and |γ | is the
number of points in γ (Exercise 3.4.4). The generating function g f (z) is
3.4. Topological Entropy and the Zeta Function of an SFT 61
The generating function is related to the zeta function by ζ f (z) = exp(zg f (z)).
The zeta function of the edge shift determined by an adjacency matrix A
is denoted by ζ A. A priori, the zeta function is merely a formal power series.
The next proposition shows that the zeta function of an SFT is a rational
function.
∞
xn 1
exp = exp(− log(1 − x)) = ,
n=1
n 1−x
and that |Fix(σ n | A)| = tr(An ) = λ λn , where the sum is over the eigenval-
ues of A, repeated with the proper multiplicity (see Exercise 1.4.2). There-
fore, if A is N × N,
∞
(λz)n ∞
(λz)n
ζ A(z) = exp = exp = (1 − λz)−1
n=1 λ
n λ n=1
n λ
−1 −1
1 1 1
= N −λ = zN det I−A = (det(I − zA))−1 .
z λ z z
The following theorem addresses the rationality of the zeta function for
a general subshift.
Exercise 3.4.2. Calculate the zeta and generating functions of the full
2-shift.
Exercise 3.4.4. Prove the product formula for the zeta function.
Exercise 3.4.5. Calculate the generating function of an edge shift with ad-
jacency matrix A.
Exercise 3.4.7. Prove that if the zeta function is rational, then so is the
generating function.
THEOREM 3.5.1 (Williams [Wil73]). The edge shifts eA and eB are topolog-
ically conjugate if and only if the matrices Aand B are strong shift equivalent.
Proof. We show here only that strong shift equivalence gives an isomor-
phism of the edge shifts. The other direction is much more difficult (see
[LM95]).
It is sufficient to consider the case when A and B are elementary strong
shift equivalent. Let A= UV, B = VU, and A, B be the (disjoint) directed
graphs with adjacency matrices Aand B. If A is k × k and B is l × l, then U
is k × l and V is l × k. We interpret the entry Ui j as the number of (additional)
edges from vertex i of A to vertex j of B, and similarly we interpret
Vji as the number of edges from vertex j of B to vertex i of A. Since
3.5. Strong Shift Equivalence and Shift Equivalence 63
a−1 a0 a1
v−1 v0 v1
u−1 u0 u1
b−1 b0
Figure 3.1. A graph constructed from an elementary strong shift equivalence.
Apq = lj=1 Upj Vjq , the number of edges in A from vertex p to vertex q is
the same as the number of paths of length 2 from vertex p to vertex q through
a vertex in B. Therefore we can choose a one-to-one correspondence φ
between the edges a of A and pairs uv of edges determined by U and V,
i.e., φ(a) = uv, so that the starting vertex of u is the starting vertex of a, the
terminal vertex of u is the starting vertex of v, and the terminal vertex of v
is the terminal vertex of a. Similarly, there is a bijection ψ from the edges
b of B to pairs vu of edges determined by V and U. For each sequence
. . . a−1 a0 a1 . . . ∈ eA apply φ to get
are strong shift equivalent but not elementary strong shift equivalent. Write
down an explicit isomorphism from ( A, σ ) to ( B, σ ).
64 3. Symbolic Dynamics
Exercise 3.5.2. Show that strong shift equivalence and shift equivalence
are equivalence relations and elementary strong shift equivalence is not.
3.6 Substitutions2
For an alphabet Am = {0, 1, . . . , m − 1}, denote by A∗m the collection of all
finite words in Am, and by |w| the length of w ∈ A∗m. A substitution s: Am →
A∗m assigns to every symbol a ∈ Am a finite word s(a) ∈ A∗m. We assume
throughout this section that |s(a)| > 1 for some a ∈ Am, and that |s n (b)| → ∞
for every b ∈ Am. Applying the substitution to each element of a sequence
or a word gives maps s: A∗m → A∗m and s: m +
→ m+
,
s
x0 x1 . . . →s(x0 )s(x1 ) . . . .
These maps are continuous but not surjective. If s(a) has the same length
for all a ∈ Am, then s is said to have constant length.
Consider the example m = 2, s(0) = 01, s(1) = 10. We have: s 2 (0) =
0110, s 3 (0) = 01101001, s 4 (0) = 0110100110010110, . . . . If w̄ is the word
obtained from w by interchanging 0 and 1, then s n+1 (0) = s n (0)s n (0). The
sequence of finite words s n (0) stabilizes to an infinite sequence
M = 01101001100101101001011001101001 . . .
called the Morse sequence. The sequences M and M̄ are the only fixed
+
points of s in m .
+
PROPOSITION 3.6.1. Every substitution s has a periodic point in m .
Proof. Consider the map a →s(a)0 . Since Am contains m elements, there
are n ∈ {1, . . . , m} and a ∈ Am such that s n (a)0 = a. If |s n (a)| = 1, then the
sequence aaa . . . is a fixed point of s n . Otherwise, |s ni (a)| → ∞, and the
sequence of finite words s ni (a) stabilizes to a fixed point of s n in m +
.
+
If a substitution s has a fixed point x = x0 x1 . . . ∈ m and |s(x0 )| > 1, then
s(x0 )0 = x0 and the sequence s (x0 ) stabilizes to x; we write x = s ∞ (x0 ). If
n
|s(a)| > 1 for every a ∈ Am, then s has at most m fixed points in m.
The closure s (a) of the (forward) orbit of a fixed point s ∞ (a) under the
shift σ is a subshift.
We call a substitution s: Am → A∗m irreducible if for any a, b ∈ Am there is
n(a, b) ∈ N such that s n(a,b) (a) contains b; s is primitive if there is n ∈ N such
that s n (a) contains b for all a, b ∈ Am.
We assume from now on that |s n (b)| → ∞ for every b ∈ Am.
where · is the Euclidean norm. Hence, by Proposition 3.6.3(1), there are
positive constants C1 and C2 such that for all k ∈ N
C1 · λk ≤ ν k ≤ ν̄k ≤ C2 · λk.
Since for every a ∈ Am there are at most ν̄k different words of length n in
s k(ab) with initial symbol in s k(a), we have
C2 2 C2 2 C2 2
ln ≤ m2 ν̄k ≤ C2 λkm2 = m λ C1 λk−1 ≤ m λ ν k−1 ≤ m λ n.
C1 C1 C1
Exercise 3.6.1. Prove that if s is primitive and s(a)0 = a, then each symbol
b ∈ Am appears in s ∞ (a) infinitely often and with bounded gaps.
01
1 u v
02
Figure 3.2. The directed graph used to construct the even system of Weiss.
Exercise 3.7.1. Prove that the even system of Weiss is not a subshift of finite
type.
Exercise 3.7.2. Prove that for any directed labeled graph , the set X is a
sofic shift.
Exercise 3.7.3. Show that there are only countably many non-isomorphic
sofic shifts. Conclude that there are subshifts that are not sofic.
10100110001
100010010010100101001.
This requires twice the length of the track, but results in fewer read/write
errors. The set of sequences produced by the MFM coding is a sofic system
(Exercise 3.8.3).
There are other considerations for storage devices that impose additional
conditions on the sequences used to encode data. For example, the total
magnetic charge of the device should not be too large. This restriction leads
to a subset of (2 , σ ) that is not of finite type and not sofic.
Recall that the topological entropy of the factor does not exceed the
topological entropy of the extension (Exercise 2.5.5). Therefore in any one-
to-one coding scheme, which increases the length of the sequence by a factor
of n > 1, the topological entropy of the original subshift must be not more
than n times the topological entropy of the target subshift.
Exercise 3.8.1. Prove that the sequences produced by MFM have at least
one and at most three 0s between every two 1s.
Exercise 3.8.3. Prove that the set of sequences produced by the MFM cod-
ing is a sofic system.
CHAPTER FOUR
Ergodic Theory
69
70 4. Ergodic Theory
Exercise 4.1.1. Prove that the unit square [0, 1] × [0, 1] with Lebesgue
measure is (measure-theoretically) isomorphic to the unit interval [0, 1] with
Lebesgue measure.
4.2 Recurrence
The following famous result of Poincaré implies that recurrence is a generic
property of orbits of measure-preserving dynamical systems.
72 4. Ergodic Theory
0 otherwise.
1 n−1
lim |µ(T −i (A) ∩ B) − µ(A) · µ(B)| = 0
n→∞ n
i=0
= 0.
lim
f (T i
(x))g(x) dµ − f dµ · g dµ
n→∞ n X X X
i=0
1 t
= 0.
lim
f (T s
(x))g(x) dµ ds − f dµ · g dµ
t→∞ t 0 X X X
PROPOSITION 4.3.3. Mixing implies weak mixing, and weak mixing implies
ergodicity.
Proof. Suppose T is a measure-preserving transformation of the probability
space (X, A, µ). Let A and B be measurable subsets of X. If T is mixing,
then |µ(T −i (A) ∩ B) − µ(A) · µ(B)| converges to 0, so the averages do as
well; thus T is weak mixing.
76 4. Ergodic Theory
Exercise 4.3.3. Show that the two definitions of strong and weak mixing
given in terms of sets and bounded measurable functions are equivalent.
4.4 Examples
We now prove ergodicity or mixing for some of the examples from Chapter 1.
k−1
P C n,n+1,...,n+k
j0 , j1 ,..., jk = q j0 Aji ji+1 .
i=0
m−1
q j = P C n+1
j = P Cin Ai j .
i=0
The pair (A, q) is called a Markov chain on the set {1, . . . , m}.
It can be shown that P extends uniquely to a shift-invariant σ-additive
measure defined on the completion C of the Borel σ-algebra generated by the
cylinders (Exercise 4.4.5); it is called the Markov measure corresponding to
A and q. The measure space (m, C, P) is a non-atomic Lebesgue probability
space. If A is irreducible, this measure is uniquely determined by A.
4.4. Examples 79
A very important particular case of this situation arises when the transi-
tion probabilities do not depend on the initial state. In this case each row of
A is the left eigenvector q, the shift-invariant measure P is called a Bernoulli
measure, and the shift is called a Bernoulli automorphism.
Let A be the adjacency matrix defined by Ai j = 0 if Ai j = 0 and Ai j = 1
if Ai j > 0. Then the support of P is precisely vA ⊂ m (Exercise 4.4.6).
and the measure µ on the Borel subsets of RZ by µ(A) = P(−1 (A)). Since
the sequence ( fi ) is stationary, the shift σ : RZ → RZ defined by (σ x)n = xn+1
preserves µ (Exercise 4.4.8).
Exercise 4.4.1. Prove that the circle rotation Rα is not weak mixing.
1 n−1
lim U i f = P f.
n→∞ n
i=0
n−1 i
Proof. Let Un = n1 i=0 U and L = {g − Ug: g ∈ H}. Note that Land I are
n−1 i
U-invariant, and I is closed. If f = g − Ug ∈ L, then i=0 U f = g − Ung
and hence Un f → 0 as n → ∞. If f ∈ I, then Un f = f for all n ∈ N. We will
show that L ⊥ I and H = L̄ ⊕ I, where L̄ is the closure of L.
Let { fk} be a sequence in L, and suppose fk → f ∈ L̄. Then
Un f
≤
Un ( f − fk)
+
Un fk
≤
Un
·
f − fk
+
Un fk
, and hence Un f → 0
as n → ∞.
Let ⊥ denote the orthogonal complement, and note that L̄⊥ = L⊥ . If
h ∈ L⊥ , then 0 = h, g − Ug = h − U ∗ h, g for all g ∈ H so that h = U ∗ h,
and hence Uh = h, by Lemma 4.5.1. Conversely (again using Lemma 4.5.1),
if h ∈ I, then h, g − Ug = h, g − U ∗ h, g = 0 for every g ∈ H, and hence
h ∈ L⊥ .
Therefore, H = L̄ ⊕ I, and limn→∞ Un is the identity on I and 0 on L̄.
The following theorem is an immediate corollary of the von Neumann
ergodic theorem.
1
N−1
f N+ (x) = f (T n (x)).
N n=0
Proof. If there are no n-leaders, the lemma is true. Otherwise, let ak be the
first n-leader, and p ≥ 1 be the smallest integer for which ak + · · · + ak+ p−1 ≥
0. If k ≤ j ≤ k + p − 1, then a j + · · · + ak+ p−1 ≥ 0, by the choice of p, and
hence a j is an n-leader. The same argument can be applied to the sequence
ak+ p , . . . , am, which proves the lemma.
THEOREM 4.5.5 (Birkhoff Ergodic Theorem). Let T be a measure-preserving
transformation in a finite measure space (X, A, µ), and let f ∈ L1 (X, A, µ).
Then the limit
1 n−1
f¯(x) = lim f (T k(x))
n→∞ n
k=0
Lemma 4.5.4,
m+n−1
0≤ sn (x) dµ = f (T k(x)) dµ. (4.1)
X k=0 Bk
i.e., if and only if the time average equals the space average for every L1
function.
The preceding corollary implies that to check the ergodicity of a measure-
preserving transformation, it suffices to verify (4.2) for a dense subset of
L1 (X, A, µ), e.g., for all continuous functions if X is a compact topological
space and µ is a Borel measure. Moreover, due to linearity it suffices to check
the convergence for a countable collection of functions that form a basis.
COROLLARY 4.5.8. A measure-preserving transformation T of a finite mea-
sure space (X, A, µ) is ergodic if and only if for every A ∈ A, for a.e. x ∈ X,
1 n−1
µ(A)
lim χ A(T k(x)) = ,
n→∞ n
k=0
µ(X)
1 n−1
lim µ T −k(A) ∩ B = µ(A) · µ(B)
n→∞ n
k=0
for any A, B ∈ A.
Exercise 4.5.2. Using the dominated convergence theorem, finish the
proof of Theorem 4.5.5 by showing that the averages n1 n−1
j=0 f converge
to f¯ in L1 .
Exercise 4.5.3. Prove that if T is a measure-preserving transformation,
then UT is an isometry of L p (X, A, µ) for any p ≥ 1.
Exercise 4.5.4. Prove Corollary 4.5.7.
Exercise 4.5.5. Prove Corollary 4.5.8.
Exercise 4.5.6. A real number x is said to be normal in base n if for any
k ∈ N, every finite word of length k in the alphabet {0, . . . , n − 1} appears
with asymptotic frequency n−k in the base-n expansion of x. Prove that almost
every real number is normal with respect to every base n ∈ N.
4.6. Invariant Measures for Continuous Maps 85
S∞
n
f (x) = lim S f (x)
j
j→∞
exists for every f ∈ F. For any g ∈ C(X) and any > 0 there is f ∈ F such
that max y∈X |g(y) − f (y)| < . Therefore, for a large enough j,
nj
Sg (x) − S∞ (x)
≤ S n j (x) +
S n j (x) − S∞ (x)
≤ 2,
f |g− f | f f
n
so Sg j (x) is a Cauchy sequence. Thus, the limit Sg∞ (x) exists for every g ∈
C(X) and defines a bounded positive linear functional Lx on C(X). By the
Riesz representation
theorem, there is a Borel probability measure µ such
that Lx (g) = X g dµ. Note that
nj
Sg (T(x)) − Sgn j (x)
= 1 |g(T n j (x)) − g(x)|.
nj
and hence the sequence X g dµn j converges for every g ∈ C(X). Therefore,
M is compact in the weak∗ topology. It is also convex: tµ + (1 − t)ν ∈ M
for any t ∈ [0, 1] and µ, ν ∈ M. A point in a convex set is extreme if it cannot
be represented as a non-trivial convex combination of two other points. The
extreme points of M are the probability measures supported on points; they
are called Dirac measures.
Let MT ⊂ M denote the set of all T-invariant Borel probability measures
on X. Then MT is closed, and therefore compact, in the weak∗ topology, and
convex.
Recall that if µ and ν are finite measures on a space X with σ-algebra
A, then ν is absolutely continuous with respect to µ if ν(A) = 0 whenever
µ(A) = 0, for A ∈ A. If ν is absolutely continuous with respect to µ, then
the Radon–Nikodym theorem asserts that there is an L1 function dν/dµ,
called the Radon–Nikodym derivative, such that ν(A) = A(dν/dµ)(x) dµ
for every A ∈ A [Roy88].
Exercise 4.6.2. Describe MT and MeT for the homeomorphism of the torus
T(x, y) = (x, x + y) mod 1.
Exercise 4.6.3
(a) Give an example of a map of the circle that is discontinuous at ex-
actly one point and does not have non-trivial finite invariant Borel
measures.
(b) Give an example of a continuous map of the real line that does not
have non-trivial finite invariant Borel measures.
*Exercise 4.6.5. Prove that if σ is the two-sided 2-shift, then Meσ is dense
in Mσ in the weak∗ topology.
lim max
Sn f (x) − f dµ
→ 0.
n→∞ x∈X X
3 The arguments of this section follow in part those of [Fur81a] and [CFS82].
88 4. Ergodic Theory
on Tk. It follows that the last coordinate Pk(n) = P(n) is uniformly dis-
tributed on S1 .
Exercise 4.7.9 finishes the proof.
Exercise 4.7.1. Prove that an irrational circle rotation is uniquely ergodic.
Exercise 4.7.2. Prove that any topologically transitive translation on a com-
pact abelian group is uniquely ergodic.
Exercise 4.7.3. Prove that the diffeomorphism T: S1 → S1 defined by
T(x) = x + a sin2 (π x), a < 1/π, is uniquely ergodic but not topologically
transitive.
Exercise 4.7.4. Prove the remaining statement of Proposition 4.7.1.
Exercise 4.7.5. Prove that the subshift defined by a fixed point a of a prim-
itive substitution s is uniquely ergodic.
Exercise 4.7.6. Let T be a uniquely ergodic continuous transformation of
a compact metric space X, and µ the unique invariant Borel probability
measure. Show that supp µ is a minimal set for T.
Exercise 4.7.7. Let S: X × G → X × G be a G-extension of T: (X, µ) →
(X, µ), and let m be the Haar measure on G. Prove that the product measure
µ × m is S-invariant.
Exercise 4.7.8. Use Fourier series on Tk to prove that T from Proposi-
tion 4.7.4 is ergodic with respect to Lebesgue measure.
Exercise 4.7.9. Reduce the general case of Theorem 4.7.5 to the case where
the leading coefficient is irrational.
For an irrational x ∈ (0, 1], the nth entry an (x) = [1/φ n−1 (x)] of the con-
tinued fraction representing x is called the n-th quotient, and we write
x = [a1 (x), a2 (x), . . .]. The irreducible fraction pn (x)/qn (x) that is equal to
the truncated continued fraction [a1 (x), . . . , an (x)] is called the nth conver-
gent of x. The numerators and denominators of the convergents satisfy the
following relations:
and
The interval b1 ,...,bn is the image of the interval [0, 1) under the map ψb1 ,...,bn
defined by
1
2. lim (a1 (x) + · · · + an (x)) = ∞.
n
n→∞
∞
log k/ log 2
1
3. lim a1 (x)a2 (x) · · · an (x) =
n 1+ .
n→∞
k=1
k2 + 2k
log qn (x) π2
4. lim =
n→∞ n 12 log 2
Proof. 1: Let f be the characteristic function of the semiopen interval
[1/k, 1/(k + 1)). Then an (x) = k if and only if f (φ n (x)) = 1. By the Birkhoff
ergodic theorem, for almost every x,
1 n−1 1
1 1 1 k+ 1
lim f (φ i (x)) = f dµ = µ , = log ,
n→∞ n
i=0 0 k k+ 1 log 2 k
f (x) if f (x) ≤ N,
f N (x) =
0 otherwise.
1 n−1
1 n−1
lim f (φ k(x)) ≥ lim f N (φ k(x))
n→∞ n k=0 n→∞ n k=0
1
n−1
= lim f N (φ k(x))
n→∞ n
k=0
1
1 f N (x)
= dx.
log 2 0 1 + x
1 f N (x)
Since lim N→∞ 0 1+x
dx → ∞, the conclusion follows.
94 4. Ergodic Theory
3: Let f (x) = log a1 (x) = log([ x1 ]). Then f ∈ L1 ([0, 1]) with respect to the
Gauss measure µ (Exercise 4.8.1). By the Birkhoff ergodic theorem,
1
1 n
1 f (x)
lim log ak(x) = dx
n→∞ n
k=1
log 2 0 1 + x
∞ k1
1 log k
= dx
log 2 k=1 k+1 1 1+x
∞
log k 1
= · log 1 + 2 .
k=1
log 2 k + 2k
It follows from the Birkhoff 1 Ergodic Theorem that the first term of (4.9)
converges a.e. to (1/ log 2) 0 log x/(1 + x) dx = −π 2 /12. The second term
converges to 0 (Exercise 4.8.2).
Exercise 4.8.1. Show that log([1/x]) ∈ L1 ([0, 1]) with respect to the Gauss
measure µ.
Exercise 4.8.2. Show that pn (x) = qn (φ(x)) and that
n−1
1 pn−k(φ k(x))
lim log(φ (x)) − log
k
= 0.
n→∞ n
k=0
qn−k(φ k(x))
UT fσ (χ ) = fσ (χId) = fσ (χ ) fσ (Id) = σ fσ (χ ),
Exercise 4.9.2. Suppose that α, β ∈ (0, 1) are irrational and α/β is irra-
tional. Let T be the translation of T2 given by T(x, y) = (x + α, y + β). Prove
that T is topologically transitive and ergodic and has discrete spectrum.
Now
n−1 1 1
1 n−1
1
|νk|2 = e2πikx ν(dx) e−2πiky ν(dy)
n k=0 n k=0 0 0
1 1
1 n−1 2πik(x−y)
= e ν(dx)ν(dy).
0 0 n k=0
The functions n−1 n−1k=0 exp(2πik(x − y)) are bounded in absolute value by
1 and converge to 1 for x = y and to 0 for x =y. Therefore the last integral
tends to the product measure ν × ν of the diagonal of [0, 1) × [0, 1). It fol-
lows that
1 n−1
lim |νk|2 = (ν({x}))2 .
n→∞ n
k=0 0≤x<1
and denote by He (U) the closure of the subspace spanned by the eigenvectors
of U. Both Hw (U) and He (U) are closed and U-invariant.
PROPOSITION 4.10.3. Let U be a (linear) isometry of a separable Hilbert
space H. Then
1. For each v ∈ H, there is a unique finite measure νv on the interval [0, 1)
(called the spectral measure) such that for every n ∈ Z
1
Un v, v = e2πinx νv (dx).
0
Proof. The first statement follows immediately from Lemma 4.10.1 and
the spectral theorem for isometries in a Hilbert space [Hel95], [Fol95]. The
second statement follows from the first (Exercise 4.10.3).
To prove the last statement let v ⊥ He and W = e−2πi x U. Applying the
von Neumann ergodic theorem 4.5.2, let u = limn→∞ n−1 n−1 k=0 W v. Then
k
4.10. Weak Mixing 99
Wu = u. By Proposition 4.10.3,
1 n−1
u, v = lim e−2πi xkU kv, v
n→∞ n
k=0
1
1 n−1 −2πi(x−y)k
= lim e νv (dy) = νv (x).
n→∞ 0 n
k=0
Proof. The splitting follows from Proposition 4.10.3. To prove the remain-
ing statement that d-limU n v, v = 0 if and only if d-limU n v, w = 0, ob-
serve that U n v, w ≡ 0 if v ⊥ U kv for all k ∈ N. If w = U kv, then U n v, w =
U n v, U kv = U n−kv, v.
Recall that if T and S are measure-preserving transformations in finite
measure spaces (X, A, µ) and (Y, B, ν), then T × S is a measure-preserving
transformation in the product space (X × Y, A × B, µ × ν). As in §4.9, we
denote by UT the isometry UT f (x) = f (T(x)) of L2 (X, A, µ).
1. T is weak mixing.
2. T has continuous
spectrum.
3. d- limn X f (T n (x)) f (x) dµ = 0 if f ∈ L2 (X, A, µ) and X f dµ = 0.
4. d- limn X f (T n (x)) g(x) dµ = X f dµ · X g dµ for all functions
f, g ∈ L2 (X, A, µ).
5. T × T is ergodic.
6. T × S is weak mixing for each weak mixing S.
7. T × S is ergodic for each ergodic S.
Proof. The transformation T is weak mixing if and only if Hw (UT )
is the orthogonal complement of the constants in L2 (X, A, µ). Therefore,
by Proposition 4.10.3, 1 ⇔ 2. By Lemma 4.10.4, 1 ⇔ 3. Clearly 4 ⇒ 3. As-
sume that 3 holds. It is enough to show 4 for f with X f dµ = 0. Observe
that 4 holds for g satisfying X f (T k(x)) g(x) dµ = 0 for all k ∈ N. Hence it
suffices to consider g(x) = f (T k(x)). But X f (T n (x)) f (T k(x)) dµ =
X f (T
n−k
(x)) f (x) dµ → 0 as n → ∞ by 3. Therefore 3 ⇔ 4.
Assume 5. Observe that T is ergodic and if UT has an eigenfunction f ,
then | f | is T-invariant, and hence constant. Therefore f (x)/ f (y) is T × T-
invariant and 5 ⇒ 2. Clearly 6 ⇒ 2 and 7 ⇒ 5.
(X × Y, A × B, µ × ν) is spanned
2
Assume 3. To prove 7 observe that L
by functions of the form f (x)g(y). Let X f dµ = Y g dν = 0. Then
f (T n (x))g(Sn (y)) f (x)g(y) dµ × ν
X×Y
= f (T n (x)) f (x) dµ · g(Sn (y))g(y) dν.
X Y
Exercise 4.10.2. Prove that d-lim has the usual arithmetic properties of
limits.
Exercise 4.10.3. Prove the second statement of Proposition 4.10.3.
Exercise 4.10.4. Prove that 3 ⇒ 6 in Theorem 4.10.6.
Exercise 4.10.5. Let T be a weak mixing measure-preserving transforma-
tion, and let S be a measure-preserving transformation such that Sk = T for
some k ∈ N (S is called a kth root of T). Prove that S is weak mixing.
4.11. Measure-Theoretic Recurrence and Number Theory 101
1 bn
Ln ( f ) = f (σ i (ω D)),
bn − an + 1 i=an
1
bnk
L( f ) = lim f (σ i (ω D))
k→∞ bnk − ank + 1
i=an k
visible by m,
1 n−1
U p(mk)
vrat − vrat < 2.
n k=0
n−1
Since (1/n) k=0 U p(mk) w → 0 by Lemma 4.11.3, for n large enough we have
1 n−1
U p(mk)
v − vrat < 2.
n k=0
6 A slight modification of the arguments above yields Proposition 4.11.2 and Theorem 4.11.4
for polynomials with integer values at integer points (rather than integer coefficients).
4.12. Internet Search 103
Theorem 4.11.4 and Proposition 4.11.1 imply the following result in com-
binatorial number theory.
add up to 1. The pair (B, q) is a Markov chain on the vertices of G̃. Google
interprets qi as the PageRank of web page i and uses it together with the
relevance factor of the page to determine how high on the return list this
page should be.
For any initial probability distribution q on the vertices of G̃, the sequence
n
q B converges exponentially to q. Thus one can find an approximation for
q by computing pBn , where q is the uniform distribution. This approach to
finding q is computationally much easier than trying to find an eigenvector
for a matrix with 1.5 billion rows and columns.
Exercise 4.12.1. Let A be an N × N stochastic matrix, and let Ainj be the
entries of An , i.e., Ainj is the probability of going from i to j in exactly n steps
(§4.4). Suppose q is an invariant probability distribution, q A = q.
(a) Suppose that for some j, we have Ai j = 0 for all i = j, and Anjk > 0
for some k = j and some n ∈ N. Show that q j = 0.
(b) Prove that if Ai j > 0 for some j =i and Anji = 0 for all n ∈ N, then
qi = 0.
CHAPTER FIVE
Hyperbolic Dynamics
Emφ(x) = φ( f (x)) for each x (Exercise 5.1.4). This means that any differen-
tiable map that is C 1 -close enough to Em is topologically conjugate to Em.
In other words, Em is structurally stable; see §5.5 and §5.11.
Hyperbolicity is characterized by local expansion and contraction, in com-
plementary directions. This property, which causes local instability of orbits,
surprisingly leads to the global stability of the topological pattern of the
collection of all orbits.
Exercise 5.1.2. Prove that limi→∞ yi0 depends continuously on (xi ) in the
product topology.
Exercise 5.1.3. Prove that if f is C 1 -close to Em, then each infinite -orbit
(xi ) of f is approximated by a unique real orbit of f that depends continu-
ously on (xi ).
1. Tx M = Es (x) ⊕ Eu (x),
2. d fxn v s ≤ Cλn v s for every v s ∈ Es (x) and n ≥ 0,
3. d fx−n v u ≤ Cλn v u for every v u ∈ Eu (x) and n ≥ 0,
4. d fx Es (x) = Es ( f (x)) and d fx Eu (x) = Eu ( f (x)).
Exercise 5.2.7. What are necessary and sufficient conditions for a periodic
orbit to be a hyperbolic set?
5.3 -Orbits
An -orbit of f : U → M is a finite or infinite sequence (xn ) ⊂ U such that
d( f (xn ), xn+1 ) ≤ for all n. Sometimes -orbits are referred to as pseudo-
orbits. For r ∈ {0, 1}, denote by distr the distance in the space of Cr -functions
(see §5.13).
for every > 0 there is δ > 0 such that for any g: O → M with dist1 (g, f ) < 0 ,
any homeomorphism h: X → X of a topological space X, and any continuous
map φ: X → O satisfying dist0 (φ ◦ h, g ◦ φ) < δ there is a continuous map
ψ: X → O with ψ ◦ h = g ◦ ψ and dist0 (φ, ψ) < . Moreover, ψ is unique in
the sense that if ψ ◦ h = g ◦ ψ for some ψ : X → O with dist0 (φ, ψ ) < δ0 ,
then ψ = ψ.
If v is a fixed point of and ψ(x) = φ(x) + v(x), then g̃(ψ(h−1 (x))) = ψ(x).
Observe that g̃(y) ∈ M and hence ψ(x) ∈ M for x ∈ X and g(ψ(h−1 (x))) =
ψ(x). Thus to prove the theorem it suffices to show that has a unique fixed
point near φ, which depends continuously on g.
Since is a hyperbolic set, for some constants λ ∈ (0, 1) and C > 1, we have
for every y ∈ and n ∈ N
n
d f v ≤ Cλn v if v ∈ Es (y), (5.2)
y
−n
d f v ≤ Cλn v if v ∈ Eu (y). (5.3)
y
s n s 1 s u n s
P d g̃ v ≤ v , P d g̃ v ≤ 1 v s , (5.4)
z z
2 100
u n u s n u
P d g̃ v ≥ 2v u , P d g̃ v ≤ 1 v u , (5.5)
z z
100
d g̃ nz v ⊥ = 0. (5.6)
Denote
last point of (xk) is < δ, to obtain a doubly infinite δ-orbit lying in the δ-
neighborhood of . Let X = (xk) with discrete topology, g = f, h be the
shift xk →xk+1 , and φ: X → U be the inclusion, φ(xk) = xk. Since (xk) is a
δ-orbit, dist(φ(h(xk)), f (φ(xk)) < δ. Theorem 5.3.1 applies, and the corollary
follows.
As in Chapter 2, denote by NW( f ) the set of non-wandering points, and
by Per( f ) the set of periodic points of f . If is f -invariant, denote by
NW( f | ) the set of non-wandering points of f restricted to . In general,
NW( f | ) =NW( f ) ∩ .
Exercise 5.3.3. Let T: [0, 1] → [0, 1] be the tent map: T(x) = 2x for 0 ≤
x ≤ 1/2 and T(x) = 2(1 − x) for 1/2 ≤ x ≤ 1. Does T have the shadowing
property?
Exercise 5.3.4. Prove that a circle rotation does not have the shadowing
property. Prove that no isometry of a manifold has the shadowing property.
Exercise 5.3.5. Show that every minimal hyperbolic set consists of exactly
one periodic orbit.
PROPOSITION 5.4.1. For every α > 0 there is = (α) > 0 such that f i ( )
⊂ U(), i = −1, 0, 1, and for every x ∈
◦ ◦
d fx Kαu (x) ⊂ K uα ( f (x)) and d f f−1
(x) Kα ( f (x)) ⊂ K α (x).
s s
and
2. d fx v < v for non-zero v ∈ Kαs (x), and d fx−1 v < v for non-
zero v ∈ Kαu (x).
Then is a hyperbolic set of f .
116 5. Hyperbolic Dynamics
Note that both sets are contained in and that f (s ) ⊂ s , f −1 (u ) ⊂ u .
are small enough. If x ∈ \s , let n(x) ∈ N be such that f n (x) ∈ for n =
−n(x)
0, 1, . . . , n(x) and f n(x)+1 (x) ∈ / , and let Es (x) = d f f n (x) ( Ẽs ( f n(x) (x))).
The continuity of Es on s and the required properties follow from Proposi-
tion 5.4.2. A similar construction with f replaced by f −1 gives an extension
of Eu .
Exercise 5.4.1. Prove that the solenoid (§1.9) is a hyperbolic set.
set g = ∞ n
n=−∞ g (O) is a hyperbolic set of g.
Exercise 5.4.4. Let be a hyperbolic set of f . Prove that if dim Eu (x) > 0
for each x ∈ , then f has sensitive dependence on initial conditions on
(see §1.12).
dist0 (h, Id) < . If one demands that the conjugacy h be C 1 , the definition
becomes vacuous. For example, if f has a hyperbolic fixed point x, then any
small enough perturbation g has a fixed point y nearby; if the conjugation
is differentiable, then the matrices dg y and d fx are similar. This condition
restricts g to lie in a proper submanifold of Diff1 (M).
where Pns (Pnu ) denotes the projection onto Es (n)(Eu (n)) parallel to
Eu (n)(Es (n)),
5.6. Stable and Unstable Manifolds 119
LEMMA 5.6.2. For any L > 0, there exists > 0 such that the graph transform
F is a well-defined operator on (L, ).
Proof. For L > 0 and x ∈ B , let KLs (n) denote the stable cone
Note that d fn−1 (0)KLs (n + 1) ⊂ KLs (n) for any L > 0. Therefore, by the uni-
form continuity of d fn , for any L > 0 there is > 0 such that d fn−1 (x)KLs (n +
1) ⊂ KLs (n) for any n ∈ N0 and x ∈ B . Hence the preimage under fn of
the graph of a Lipschitz-continuous function is the graph of a Lipschitz-
continuous function. For φ ∈ (L, ), consider the following composition
β = Ps (n) ◦ fn−1 ◦ φn , where Ps (n) is the projection onto Es (n) parallel to
Eu (n). If is small enough, then β is an expanding map and its image covers
Bs (n) (Exercise 5.6.1). Hence F(φ) ∈ (L, ).
The next lemma shows that F is a contracting operator for an appropriate
choice of and L.
120 5. Hyperbolic Dynamics
E u (n + 1) fn (cu )
E u (n) φn
fn
φn+1
cu
ψn+1
ψn E s (n + 1)
E s (n)
0 y 0 z
LEMMA 5.6.3. There are > 0 and L > 0 such that F is a contracting oper-
ator on (L, ).
Proof. For L ∈ (0, 0.1), let KLu (n) denote the unstable cone
KLu (n) = {v ∈ Rm: v = v s + v u , v s ∈ Es (n), v u ∈ Eu (n), |v u | ≥ L−1 |v s |},
and note that d fn (0)KLu (n) ⊂ KLu (n + 1). As in Lemma 5.6.2, by the uniform
continuity of d fn , for any L > 0 there is > 0 such that the inclusion
d fn KLu (n) ⊂ KLu (n + 1) holds for every n ∈ N0 and x ∈ B .
Let φ, ψ ∈ (L, ), φ = F(φ), ψ = F(ψ) (see Figure 5.1). For any η > 0
there are n ∈ N0 and y ∈ Bs such that |φn (y) − ψn (y)| > d(φ , ψ ) − η. Let cu
be the straight line segment from (y, φn (y)) to (y, φn (y)). Since cu is parallel
to Eu (n), we have that length( fn (cu )) > λ−1 length(cu ). Let fn (y, ψn (y)) =
(z, ψn+1 (z)), and consider the curvilinear triangle formed by the straight line
segment from (z, φn+1 (z)) to (z, ψn+1 (z)), fn (cu ), and the shortest curve on
the graph of ψn+1 connecting the ends of these curves. For a small enough
> 0 the tangent vectors to the image fn (cu ) lie in KLu (n + 1) and the tangent
vectors to the graph of φn+1 lie in KLs (n + 1). Therefore,
length( fn (cu ))
|φn+1 (z) − ψn+1 (z)| ≥ − L(1 + L) · length( fn (cu ))
1 + 2L
≥ (1 − 4L)length( fn (cu )),
and
d(φ, ψ) ≥ |φn+1 (z) − ψn+1 (z)| ≥ (1 − 4L) length ( fn (cu ))
> (1 − 4L)λ−1 length (cu ) = (1 − 4L)λ−1 (d(φ , ψ ) − η).
Since η is arbitrary, F is contracting for small enough L and .
Since F is contracting (Lemma 5.6.3) and depends continuously on f ,
it has a unique fixed point φ ∈ (L, ), which depends continuously on f
(property 6) and automatically satisfies property 2. The invariance of the
5.6. Stable and Unstable Manifolds 121
stable and unstable cones (with a small enough ) implies that φ satisfies
properties 3 and 4. Property 1 follows immediately from 3 and 4. Since
property 1 gives a geometric characterization of graph(φn ), the fixed point
of F for a smaller is a restriction of the fixed point of F for a larger to
a smaller domain. As → 0 and L → ∞, the stable cone KLs (0, n) (which
contains graph(φn )) tends to Es (n). Therefore Es (n) is the tangent plane to
graph(φn ) at 0 (property 5).
The following theorem establishes the existence of local stable manifolds
for points in a hyperbolic set and in sδ , and of local unstable manifolds
for points in and in uδ (see §5.4); recall that sδ ⊃ and uδ ⊃ .
called the local stable manifold of x s and the local unstable manifold
of x u , are C 1 embedded disks,
2. Tys Ws (x s ) = Es (ys ) for every ys ∈ Ws (x s ), and Tyu Wu (x u ) = Eu (yu )
for every yu ∈ Wu (x u ) (see Proposition 5.4.4),
3. f (Ws (x s )) ⊂ Ws ( f (x s )) and f −1 (Wu ( f (x u ))) ⊂ Wu (x u ),
4. if ys , zs ∈ Ws (x s ), then ds ( f (ys ), f (zs )) < λds (ys , zs ), where ds is the
distance along Ws (x s ),
if yu , zu ∈ Wu (x u ), then du ( f −1 (yu ), f −1 (zu )) < λdu (yu , zu ), where du
is the distance along Wu (x u ),
5. if 0 < dist(x s , y) < and exp−1 x s (y) lies in the δ-cone Kδ (x ), then dist
u s
−1
( f (x ), f (y)) > λ dist(x , y),
s s
if 0 < dist(x u , y) < and exp−1 x u (y) lies in the δ-cone Kδ (x ), then
s u
fn = ψ f n (xs ) ◦ f ◦ ψ −1
f n−1 (x s )
, Es (n) = dψ f n (xs ) (x s )Es ( f n (x s )), and Eu (n) =
dψ f n (x) (x)E ( f (x)), apply Proposition 5.6.1, and set Ws (x) = W0s (). Simi-
u n
PROPOSITION 5.6.5. There is 0 > 0 such that for every ∈ (0, 0 ), for every
x ∈ ,
∞
∞
Ws (x) = f −n Ws ( f n (x) , Wu (x) = f n Wu ( f −n (x)).
n=0 n=0
Bl
graph(φ)
φ(Dn )
In
Rk 0 Bk
LEMMA 5.7.1. Let λ ∈ (0, 1), , δ ∈ (0, 0.1). Suppose f : Bk × Bl → Rm and
φ: Bk → Bl are C 1 maps such that
1. 0 is a hyperbolic fixed point of f ,
2. Wu (0) = Bk × {0} and Ws (0) = {0} × Bl ,
3. d fx (v) ≥ λ−1 v for every v ∈ Kδu whenever both x, f (x) ∈ Bk × Bl ,
4. d fx (v) ≤ λv for every v ∈ Kδs whenever both x, f (x) ∈ Bk × Bl ,
5. d fx (Kδu ) ⊂ Kδu whenever both x, f (x) ∈ Bk × Bl ,
6. d( f −1 )x (Kδs ) ⊂ Kδs whenever both x, f −1 (x) ∈ Bk × Bl ,
7. T(y,φ(y)) graph(φ) ⊂ Kδu for every y ∈ Bk,
Then for every n ∈ N there is a subset Dn ⊂ Bk diffeomorphic to B k and such
that the image In under f n of the graph of the restriction φ| Dn has the following
properties: π u (In ) ⊃ B/2
k
and Tx In ⊂ Kδλ
u
2n for each x ∈ In .
Proof. The lemma follows from the invariance of the cones (Exercise 5.7.2).
The meaning of the lemma is that the tangent planes to the image of
the graph of φ under f n are exponentially (in n) close to the “horizontal”
space Rk, and the image spreads over Bk in the horizontal direction (see
Figure 5.2).
The following theorem, which is also sometimes called the Lambda
Lemma, implies that if f is Cr with r ≥ 1, and D is any C 1 -disk that in-
tersects transversely the stable manifold Ws (x) of a hyperbolic fixed point
x, then the forward images of D converge in the Cr topology to the unstable
manifold Wu (x) [PdM82]. We prove only C 1 convergence. Let BRu be the ball
of radius R centered at x in Wu (x) in the induced metric.
Then for every R > 0 and β > 0 there are n0 ∈ N and, for each n ≥ n0 ,
a subset D̃ = D̃(R, β, n), y ∈ D̃ ⊂ D, diffeomorphic to an open k-disk and
such that the C 1 distance between f n ( D̃)) and BRu is less than β.
Proof. We will show that for some n1 ∈ N, an appropriate subset D1 ⊂
f n1 (D) satisfies the assumptions of Lemma 5.7.1. Since {x} is a hyperbolic
set of f , for any δ > 0 there is > 0 such that Es (x) and Eu (x) can be ex-
tended to invariant distributions Ẽs and Ẽu in the -neighborhood B of
x and the hyperbolicity constant is at most λ + δ (Proposition 5.4.4). Since
f n (y) → x, there is n2 ∈ N such that z = f n2 (y) ∈ B . Since D intersects
Ws (x) transversely, so does f n2 (D). Therefore there is η > 0 such that if
v ∈ Tz f n2 (D), v = 1, v = v s + v u , v s ∈ Ẽs (z), v u ∈ Ẽu (z), and v u =0, then
v u ≥ ηv s . By Proposition 5.4.4, for a small enough δ > 0, the norm
d f n v s decays exponentially and d f n v u grows exponentially. Therefore,
for an arbitrarily small cone size, there exists n2 ∈ N such that Tf n2 (y) f n2 (D)
lies inside the unstable cone at f n2 (y).
Exercise 5.7.1. Prove that if x is a homoclinic point, then x is non-wandering
but not recurrent.
1. f is one-to-one on R;
2. f (R) ∩ R has at least two components 0 , . . . , q−1 ;
3. if z ∈ R and f (z) ∈ i , 0 ≤ i < q, then the sets Giu (z) = f (F u (z)) ∩
5.8. Horseshoes and Transverse Homoclinic Points 125
Rl
R
F u (z) z
f (z)
Rk
d fx−1 Kδ/2
s
⊂ Kδ/2
s
, |d fx v| < λ|v| if v ∈ Kδ/2
s
.
W s (p)
f N (R)
f nu (q)
R
nu +1
f (q)
p W u (p)
f k (R) f −ns (q)
so that β/λk > 10. There is λ ∈ (0, 1) such that if x ∈ R and f N (x) is close to
either f nu (q) or f nu +1 (q) (i.e., f k(x) is close to f −ns (q) or f −ns −1 (q)), then
Kδu is invariant under d fxN and λ |d fxN v| ≥ |v| for every v ∈ Kδu . Similarly,
for an appropriate choice of R and k, the stable δ-cones are invariant under
d f −N and vectors from Kδs are stretched by d f −N by a factor at least (λ )−1 .
To guarantee the correct intersection of f N (R) with R we must choose R
carefully. Choose the horizontal boundary segments of R to be straight line
segments, and let R stretch vertically so that it crosses Wu ( p) near f nu (q) and
f nu +1 (q). By Theorem 5.7.2, the images of these horizontal segments under
f k are almost horizontal line segments. To construct the vertical boundary
segments of R, take two vertical segments s1 and s2 to the left of f −ns −1 (q)
and to the right of f −ns (q), and truncate their preimages f −k(si ) by the
horizontal boundary segments. By Theorem 5.7.2, the preimages are almost
vertical line segments. This choice of R satisfies the definition of a horseshoe.
Exercise 5.8.1. Let f : U → M be a diffeomorphism, p a periodic point
of f , and q a (non-transverse) homoclinic point (for p). Prove that every
128 5. Hyperbolic Dynamics
LEMMA 5.9.2. For every > 0 there is δ > 0 such that if φ: Bk → Rl and
ψ: Bl → Rk are differentiable maps and |φ(x)|, dφ(x), |ψ(y)|, dφ(y) < δ
for all x ∈ Bk and y ∈ Bl , then the intersection graph(φ) ∩ graph(ψ) ⊂ Rk+l
is transverse and consists of exactly one point, which depends continuously
on φ and ψ in the C 1 topology.
Proof. Exercise 5.9.3.
The following property of hyperbolic sets plays a major role in their geo-
metric description and is equivalent to local maximality. A hyperbolic set
has a local product structure if there are (small enough) > 0 and δ > 0 such
that (i) for all x, y ∈ the intersection Ws (x) ∩ Wu (y) consists of at most one
point, which belongs to , and (ii) for x, y ∈ with d(x, y) < δ, the inter-
section consists of exactly one point of , denoted [x, y] = Ws (x) ∩ Wu (y),
and the intersection is transverse (Proposition 5.9.1). If a hyperbolic set
has a local product structure, then for every x ∈ there is a neighborhood
U(x) such that
U(x) ∩ = [y, z]: y ∈ U(x) ∩ Ws (x), z ∈ U(x) ∩ Wu (x) .
to f and denoted by ds and du , the distances along the stable and unstable
leaves.
PROPOSITION 5.10.2
1. Anosov diffeomorphisms form an open (possibly empty) subset in the
C 1 topology (Corollary 5.5.2).
2. Anosov diffeomorphisms are structurally stable (Corollary 5.5.4).
3. The set of periodic points of an Anosov diffeomorphism is dense in the
set of non-wandering points (Corollary 5.3.4).
Here is a more direct proof of the density of periodic points. Let and
δ satisfy Proposition 5.10.1. If x ∈ M is non-wandering, then there is n ∈ N
and y ∈ M such that dist(x, y), dist( f n (y), y) < δ/(2C p ). Assume that λn <
1/(2C p ). Then the map z →[y, f n (z)] is well defined for z ∈ Wδs (y). It maps
Wδs (y) into itself and, by the Brouwer fixed point theorem, has a fixed point
y1 such that ds (y1 , y) < δ, f n (y1 ) ∈ Wu (y1 ) and du (y1 , f n (y1 )) < δ. The map
f −n sends Wδu ( f n (y1 )) to itself and therefore has a fixed point.
Exercise 5.10.1. Prove that the stable and unstable manifolds of an Anosov
diffeomorphism form foliations (see §5.13).
NW( f ) = 1 ∪ 2 ∪ · · · ∪ k,
as a disjoint union of closed f -invariant sets (called basic sets) such that
1. each i is a locally maximal hyperbolic set of f ;
2. f is topologically transitive on each i ; and
j
3. each i is a disjoint union of closed sets i , 1 ≤ j ≤ mi , the diffeomor-
j
phism f cyclically permutes the sets i , and f mi is topologically mixing
j
on each i .
The basic sets are precisely the closures of the equivalence classes of ∼.
For two basic sets, we write i ≤ j if there are periodic points p ∈ i and
q ∈ j such that p ≤ q.
134 5. Hyperbolic Dynamics
p0 p0
A3
B2 B2
p2
p2
B1
p3
A2 A2
p1
p1
A1
p0 p0
A3
p3
p2 Delta
∆ 4
p1
∆2
p0
∆3 p0
∆5
p3 ∆4 p2
p1 ∆1 p1
∆2
p0
p0
∆3
p3
Figure 5.6. The image of the Markov partition under fM .
Exercise 5.12.1. Prove that the stable boundary is forward invariant and
the unstable boundary is backward invariant under fM .
Exercise 5.12.2. Prove that for the toral automorphism f M , the intersec-
tion of the preimages of rectangles i along an allowed infinite sequence
ω consists of exactly one point. Prove that there is a semiconjugacy φ from
σ | A to the toral automorphism fM .
Exercise 5.12.3. Prove that the map ψ defined in the text above satisfies
ψ(x) ∈ A and that φ ◦ ψ = Id.
1. U is a cover of M, and
2. for U, V ∈ U, if U ∩ V = ∅, the mapφU ◦ φV−1 : φV (U ∩ V) → φU (U ∩
V) is C k.
We may take k ∈ N ∪ {∞, ω}, where C ω denotes the class of real analytic
functions.
We write Mm to indicate that M has dimension m. If x ∈ M and U ∈ U
contains x, then the pair (U, φU ), U ∈ U, is called a coordinate chart at x,
and the n component functions x1 , x2 , . . . , xm of φU are called coordinates on
138 5. Hyperbolic Dynamics
for f ∈ C 1 (M). The tangent space at p is the linear space Tp M of all tangent
vectors at p.
Suppose (U, φ) is a coordinate chart, with coordinate functions x1 , . . . , xm,
and let p ∈ U. For i = 1, . . . , m, consider the curves
p
Define (∂/∂ xi ) p to be the tangent vector to αi at p, i.e., for g ∈ C 1 (M),
∂ d p ∂
(g) = g α (t) = (g ◦ φ) .
∂ xi p dt t=0 i
∂ xi φ( p)
141
142 6. Ergodicity of Anosov Diffeomorphisms
The following lemmas can be used to prove the Hölder continuity of in-
variant distributions for a variety of dynamical systems. Our objective is the
Hölder continuity of the stable and unstable distributions of an Anosov dif-
feomorphism, which was first established by Anosov [Ano67]. We consider
only the stable distribution; Hölder continuity of the unstable distribution
follows by reversing the time.
Then
µ
dist(E1 , E2 ) ≤ 3C 2 δ (log µ−log λ)/(log b−log λ) .
λ
Proof. Set Kn1 = {v ∈ R N : L1n v ≤ 2Cλn v}. Let v ∈ Kn1 . Write v = v 1 +
v⊥
1
, where v 1 ∈ E1 and v⊥1
⊥ E1 . Then
1 1 1
L v = L v + v 1 ≥ L1 v 1 − L1 v 1 ≥ C −1 µn v 1 − Cλn v 1 ,
n n ⊥ n ⊥ n ⊥
and hence
n
1
v ≤ Cµ−n L1 v + Cλn v 1 ≤ 3C 2 λ v.
⊥ n
µ
It follows that
n
λ
dist(v, E ) ≤ 3C
1 2
v. (6.1)
µ
Set γ = λ/b < 1. There is a unique non-negative integer k such that
γ k+1 < δ ≤ γ k. Let v 2 ∈ E2 . Then
1 2 2 2 1
L v ≤ L v + L − L2 · v 2
k k k k
≤ Cλkv 2 + bkδv 2
≤ (Cλk + (bγ )k)v 2 ≤ 2Cλkv 2 .
6.1. Hölder Continuity of the Stable and Unstable Distributions 143
W
U1
U2
p
Figure 6.1. Holonomy map p for a foliation W and transversals U1 and U2 .
L = FU (x)
WU (s)
z=r
s
WU (y) p̄s
x y
py
FU (y)
Let py denote the holonomy map along the leaves of W from FU (x) = L
to FU (y), and let qy (·) denote the Jacobian of py . We have
1 A(y, z) δ̄ y (z) dmF(y) (z) = 1 A( py (s))qy (s) δ̄ y ( py (s)) dmL(s),
FU (y) L
Similarly, let p̄s denote the holonomy map along the leaves of F from WU (x)
to WU (s), s ∈ L, and let q̄s denote the Jacobian of p̄s . We transform the
integral over WU (x) into an integral over WU (s) using the change of variables
6.2. Absolute Continuity of the Stable and Unstable Foliations 147
r = py (s), y = p̄−1
s (r ):
1 A( py (s))qy (s) δ̄ y ( py (s)) dmW(x) (y)
WU (x)
= 1 A(r )qy (s) δ̄ y (r )q̄−1
s (r ) dmW(s) (r ).
WU (s)
The last formula together with (6.3) gives the absolute continuity of W.
The converse of Proposition 6.2.2 is not true in general (Exercise 6.2.2).
LEMMA 6.2.3. Let (X, A, µ), (Y, B, ν) be two compact metric spaces
with Borel σ-algebras and σ-additive Borel measures, and let pn : X → Y,
n = 1, 2, . . . , and p: X → Y be continuous maps such that
1. each pn and p are homeomorphisms onto their images,
2. pn converges to p uniformly as n → ∞,
3. there is a constant J such that ν( pn (A)) ≤ J µ(A) for every A ∈ A.
Then ν( p(A)) ≤ J µ(A) for every A ∈ A.
Proof. It is sufficient to prove the statement for an arbitrary open ball
Br (x) in X. If δ < r then p(Br −δ (x)) ⊂ pn (Br (x)) for n large enough,
and hence ν( p(Br −δ (x))) ≤ ν( pn (Br (x))) ≤ J µ(Br (x)). Observe now that
ν( p(Br −δ (x))) ν( p(Br (x))) as δ 0.
For subspaces A, B ⊂ R N , set
assume that (a) x = (0, 0), (b) T(0,0) Q1 = {v = 0}, (c) p(x) = (0, v0 ), where
v0 = p̂(x) − x, (d) T(0,v0 ) Q2 is given by the equation v = v0 + Bu, where
B is a k × (N − k) matrix whose norm depends only on δ, and (e) Ê(0, 0) =
{u = 0}, and Ê(w, 0) is given by the equation u = w + A(w)v, where A(w)
is an (N − k) × k matrix which is C 1 in w and A(0) = 0.
The image of (w, 0) under p̂ is the intersection point of the planes v =
v0 + Bu and u = w + A(w)v. Since the norm of B is bounded from above in
terms of ξ , it suffices to estimate the determinant of the derivative ∂u/∂w at
w = 0. We substitute the first equation into the second one,
u = w + A(w)v0 + A(w)Bu;
f n (L1 )
p̂(y1 ) = f n (x2 )
s
W
x2 = pn (x1 )
pn f n (p(x1 ))
x1 p(x1 ) y1 = f n (x1 )
p
Ê(y1 )
s
W
L1 L2 f n (L2 )
dist( f k(x1 ), f k(x2 )) ≤ dist( f k(x1 ), f k( p(x1 ))) + dist( f k( p(x1 )), f k(x2 ))
≤ C3 λk. (6.6)
1n−1
ψk+ (x) = lim ψk( f i (x))
n→∞ n
i=0
152 6. Ergodicity of Anosov Diffeomorphisms
and hence
1n−1
φC (y) − ψk( f (y)) dµ(y)
i
X n i=0
1n−1
1
≤ |φC (y) − ψk( f i (y))| dµ(y) < .
n i=0 X k
Since ψk is uniformly continuous, ψk+ (y) = ψk+ (x) whenever y ∈ V s (x) and
ψk+ (x) is defined. Therefore, there is a null set Nk such that ψk+ exists and is
constant on the stable sets in X\Nk. It follows that φC+ (x) = limk→∞ ψk+ (x)
is constant on the stable sets in X\ Nk. Clearly φC (x) = φC+ (x) mod 0.
Let φ be a µ-measurable f -invariant function. By Lemma 6.3.2, there is
a µ-null set Ns such that φ is constant on the leaves of Ws in M\Ns and
another µ-null set Nu such that φ is constant on the leaves of Wu in M\Nu .
Let x ∈ M, and let U x be a small neighborhood, as in the definition of
absolute continuity for Ws and Wu . Let Gs ⊂ U be the set of points z ∈ U for
which mWs (z) (Ns ∩ Ws (z)) = 0 and z ∈/ Ns . Let Gu ⊂ U be the set of points
z ∈ U for which mWu (z) (Nu ∩ Wu (z)) = 0 and z ∈ / Nu . By Proposition 6.2.1
and the absolute continuity of Wu and Ws (Theorem 6.2.5), both sets Gs
and Gu have full µ-measure in U, and hence so does Gs ∩ Gu . Again, by the
absolute continuity of Wu , there is a full-µ-measure subset of points z ∈ U
such that z ∈ Gs ∩ Gu and mWu (z) -a.e. point from Wu (z) also lies in Gs ∩ Gu .
It follows that φ(x) = φ(z) for µ-a.e. point x ∈ U. Since M is connected, φ is
constant mod 0 on M.
Exercise 6.3.1. Prove that a C 2 Anosov diffeomorphism preserving a
smooth measure is weak mixing.
CHAPTER SEVEN
Low-Dimensional Dynamics
153
154 7. Low-Dimensional Dynamics
Thus, | pm/m − pn /n| < |1/m + 1/n|, so { pn /n} is a Cauchy sequence. It fol-
lows that (F n (x) − x)/n converges as n → ∞.
If G = F + k is another lift of f , then ρ(G) = ρ(F) + k, so ρ( f ) is inde-
pendent of the lift F. Moreover, there is a unique lift F such that ρ(F) = ρ( f )
(Exercise 7.1.1).
Since S1 = [0, 1] mod 1, we will often abuse notation by writing ρ( f ) = x
for some x ∈ [0, 1].
Proof. The “if” part of the first assertion is contained in Theorem 7.1.1.
Suppose ρ( f ) = p/q, where p, q ∈ N. If F and F̃ = F + l are two lifts
of f , then F̃ q = F q + lq. Thus we may choose F to be the unique lift with
p ≤ F q (0) < p + q. To show the existence of a periodic point of f , it suffices
to show the existence of a point x ∈ [0, 1] such that F q (x) = x + k for some
k ∈ N. We may assume that x + p < F q (x) < x + p + q for all x ∈ [0, 1],
since otherwise we have F q (x) = x + l for k = p or k = p + q, and we are
done. Choose > 0 such that for any x ∈ [0, 1], x + p + < F q (x) < x +
p + q − . The same inequality then holds for all x ∈ R, since F q (x + k) =
F q (x) + k for all k ∈ N. Thus
contradicting the fact that F q (x) = x + p . Thus, x is periodic with period q.
Suppose f is a homeomorphism of S1 . Given any subset A ⊂ S1 and a
distinguished point x ∈ A, we define an ordering on A by lifting A to the
interval [x̃, x̃ + 1) ⊂ R, where x̃ ∈ π −1 (x), and using the natural ordering on
R. In particular, if x ∈ S1 , then the orbit {x, f (x), f 2 (x), . . .} has a natural
order (using x as the distinguished point).
LEMMA 7.1.6. Suppose ρ( f ) is irrational. Then for any x ∈ S1 and any dis-
tinct integers m > n, every forward orbit of f intersects the interval
I = [ f m(x), f n (x)].
Proof. It suffices to show that S1 = ∞
k=0 f
−k
I. Suppose not. Then
∞
∞
−(k−1)m+kn
S1
⊂ f −k(m−n) I = f (x), f −km+(k+1)n (x) .
k=1 k=1
Moreover,
Exercise 7.1.1. Show that if F and G = F + kare two lifts of f , then ρ(F) =
ρ(G) + k, so ρ( f ) is independent of the choice of lift used in its definition.
Show that there is a unique lift F of f such that ρ(F) = ρ( f ).
Exercise 7.1.5. Show that Theorems 7.1.1 and 7.1.5 hold under the weaker
hypothesis that f : S1 → S1 is a continuous map such that any (and thus
every) lift F of f is non-decreasing.
160 7. Low-Dimensional Dynamics
where the supremum is taken over all partitions 0 ≤ x1 < · · · < xn ≤ 1, for
all n ∈ N. We say that g has bounded variation if Var(g) is finite. Note that
any Lipshitz function has bounded variation. In particular, any C 1 function
has bounded variation.
THEOREM 7.2.1 (Denjoy). Let f be an orientation-preserving C 1 diffeo-
morphism of the circle with irrational rotation number ρ = ρ( f ). If f has
bounded variation, then f is topologically conjugate to the rigid rotation Rρ .
Proof. We know from Theorem 7.1.9 that if f is transitive, it is conjugate
to Rρ . Thus we assume that f is not transitive, and argue to obtain a contra-
diction. By Proposition 7.1.7, we may assume that ω(0) is a perfect, nowhere
dense set. Then S1 \ω(0) is a disjoint union of open intervals. Let I = (a, b)
be one of these intervals. Then the intervals { f n (I)}n∈Z are pairwise disjoint,
since otherwise f would have a periodic point. Thus n∈Z l ( f n (I)) ≤ 1,
b
where l ( f n (I)) = a ( f n ) (t) dt is the length of f n (I).
LEMMA 7.2.2. Let J be an interval in S1 , and suppose the interiors of the
intervals J, f (J ), . . . , f n−1 (J ) are pairwise disjoint. Let g = log f , and fix
x, y ∈ J . Then for any n ∈ Z,
Var(g) ≥ | log( f n ) (x) − log( f n ) (y)|.
k=0
k=0
n−1
n−1
= log f ( f k(y)) − log f ( f k(x))
k=0 k=0
Fix x ∈ S1 . We claim that there are infinitely many n ∈ N such that the
intervals (x, f −n (x)), ( f (x), f 1−n (x)), . . . , ( f n (x), x) are pairwise disjoint. It
suffices to show that there are infinitely many n such that f k(x) is not in the
interval (x, f n (x)) for 0 ≤ |k| ≤ n. Lemma 7.1.8 implies that the orbit of x is
7.2. Circle Diffeomorphisms 161
ordered in the same way as the orbit of a point under the irrational rotation
Rρ . Since the orbit of a point under an irrational rotation is dense, the claim
follows.
Choose n as in the preceding paragraph. Then by applying Lemma 7.2.2
with y = f −n (x), we obtain
( f n ) (x)
Var(g) ≥ log n = | log(( f n ) (x)( f −n ) (x))|.
( f ) (y)
Thus for infinitely many n ∈ N, we have
= [( f n ) (x) + ( f −n ) (x)] dx
I
≥ ( f n ) (x)( f −n ) (x) dx
I
1
≥ exp (−Var(g)) dx = exp − Var(g) l (I).
I 2
This contradicts the fact that n∈Z l ( f n (I)) ≤ ∞, so we conclude that f is
transitive, and therefore conjugate to Rρ .
THEOREM 7.2.3 (Denjoy Example). For any irrational number ρ ∈ (0, 1),
there is a non-transitive C 1 orientation-preserving diffeomorphism f : S1 → S1
with rotation number ρ.
Proof. We know from Lemma 7.1.8 that if ρ( f ) = ρ, then for any x ∈ S1 ,
the orbit of x is ordered the same way as any orbit of Rρ , i.e., f k(x) < f l (x) <
f m(x) if and only if Rρk(x) < Rρl (x) < Rρm(x). Thus in constructing f , we have
no choice about the order of the orbit of any point. We do, however, have a
choice about the spacing between points in the orbit.
Let {ln }n∈Z be a sequence of positive real numbers such that n∈Z ln = 1
and ln is decreasing as n → ±∞ (we will impose additional constraints later).
Fix x0 ∈ S1 , and define
an = l k, bn = an + ln .
{k∈Z:Rρk (x0 )∈[x0 ,Rρn (x0 )}
The intervals [an , bn ] are pairwise disjoint. Since n∈Z ln = 1, the union of
these intervals covers a set of measure 1 in [0, 1], and is therefore dense.
To define a C 1 homeomorphism f : S1 → S1 it suffices to define a contin-
uous, positive function g on S1 with total integral 1. Then f will be defined
162 7. Low-Dimensional Dynamics
to be the
b integral of g. The function g should satisfy:
1. ann g(t) dt = ln+1 .
To construct such a g it suffices to define g on each interval [an , bn ] so that it
also satisfies:
2. g(an ) = g(bn ) = 1.
3. For any sequence {xk} ⊂ n∈Z [an , bn ], if y = lim xk, then g(xk) → 1.
We then define g to be 1 on S1 \ n∈Z [an , bn ].
There are many such possibilities for g|[an , bn ]. We use the quadratic
polynomial
6(ln+1 − ln )
g(x) = 1 + (bn − x)(x − an ),
ln3
which clearly satisfies condition 1. For n ≥ 0, we have ln+1 − ln < 0, so
6(ln − ln+1 ) ln 2 3ln+1 − ln
1 ≥ g(x) ≥ 1 − 3
= .
ln 2 2ln
For n < 0, we have ln+1 − ln > 0, so
3ln+1 − ln
1 ≤ g(x) ≤ .
2ln
Thus if we choose ln such that (3ln+1 − ln )/2ln → 1 as n → ±∞, then condi-
tion 3 is satisfied. For example, we could choose ln = α(|n| + 2)−1 (|n| + 3)−1 ,
where α = 1/ n∈Z ((|n| + 2)−1 (|n| + 3)−1 ).
x
Now define f (x) = a1 + 0 g(t) dt. Using the results above, it follows
that f : S1 → S1 is a C 1 homeomorphism of S1 with rotation number ρ
(Exercise 7.2.1). Moreover, f n (0) = an , and ω(0) = S1 \ n∈Z (an , bn ) is a
closed, perfect, invariant set of measure zero.
Exercise 7.2.1. Verify the statements in the last paragraph of the proof of
Theorem 7.2.3.
Exercise 7.2.2. Show directly that the example constructed in the proof of
Theorem 7.2.3 is not C 2 .
LEMMA 7.3.2
1. If f (I) ⊃ I, then the closure of I contains a fixed point of f .
2. Fix m ∈ N ∪ {∞}, and suppose that {Ik}1≤k<m is a finite or infinite se-
quence of non-empty closed intervals in [0, 1] such that f (Ik) ⊃ Ik+1 for
1 ≤ k < m − 1. Then there is a point x ∈ I1 such that f k(x) ∈ Ik+1 for
1 ≤ k < m − 1. Moreover, if In = I1 for some n > 0, then I1 contains a
periodic point x of period n such that f k(x) ∈ Ik+1 for k = 1, . . . , n − 1.
Proof. The proof of part 1 is a simple application of the intermediate value
theorem.
To prove part 2, note that since f (I1 ) ⊃ I2 , there are points a0 , b0 ∈ I1
that map to the endpoints of I2 . Let J1 be the subinterval of I1 with
endpoints a0 , b0 . Then f (J1 ) = I2 . Suppose we have defined subintervals
J1 ⊃ J2 ⊃ · · · ⊃ Jn in I1 such that f k(Jk) = Ik+1 . Then f n+1 (Jn ) = f (In+1 ) ⊃
In+2 , so there is an interval Jn+1 ⊂ Jn such that f n+1 (Jn+1 ) = In+2 . Thus we
obtain a nested sequence {Jn } of non-empty closed intervals. The intersection
164 7. Low-Dimensional Dynamics
m−1
i=1 Ji is non-empty, and for any x in the intersection, f k(x) ∈ Ik+1 for
1 ≤ k < m − 1.
The last assertion follows from the preceding paragraph together with
part 1.
A partition of an interval I is a (finite or infinite) collection of closed subin-
tervals {Ik}, with pairwise disjoint interiors, whose union is I. The Markov
graph of f associated to the partition {Ik} is the directed graph with ver-
tices Ik, and a directed edge from Ii to I j if and only if Ii f-covers I j . By
Lemma 7.3.2, any loop of length n in the Markov graph of f forces the
existence of a periodic point of (not necessarily minimal) period n.
As a warmup to the proof of the full Sharkovsky theorem, we prove
that the existence of a periodic point of minimal period three implies the
existence of periodic points of all periods. This result was rediscovered in
1975 by T. Y. Li and J. Yorke, and popularized in their paper “Period three
implies chaos” [LY75].
Let x be a point of period three. Replacing x with f (x) or f 2 (x) if nec-
essary, we may assume that x < f (x) and x < f 2 (x). Then there are two
cases: (1) x < f (x) < f 2 (x) or (2) x < f 2 (x) < f (x). In the first case, we let
I1 = [x, f (x)] and I2 = [ f (x), f 2 (x)]. The associated Markov graph is one
of the two graphs shown in Figure 7.1.
For k ≥ 2, the path I1 → I2 → I2 → · · · → I2 → I1 of length k implies the
existence of a periodic point y of period k with the itinerary I1 , I2 , I2 , . . . ,
I2 , I1 . If the minimal period of y is less than k, then y ∈ I1 ∩ I2 = { f (x)}. But
f (x) does not have the specified itinerary for k
= 3, so the minimal period of
y is k. A similar argument applies to case (2), and this proves the Sharkovsky
theorem for n = 3.
To prove the full Sharkovsky theorem it is convenient to use a sub-
graph of the Markov graph defined as follows. Let P = {x1 , x2 , . . . , xn } be
a periodic orbit of (minimal) period n > 1, where x1 < x2 < · · · < xn . Let
I1 I2
I1 I2
Figure 7.1. The two possible Markov graphs for period three.
7.3. The Sharkovsky Theorem 165
LEMMA 7.3.3. The P-graph of f contains a trivial loop, i.e., there is a vertex
I j with a directed edge from I j to itself.
Proof. Let j = max{i: f (xi ) > xi }. Then f (x j ) > x j and f (x j+1 ) ≤ x j+1 , so
f (x j ) ≥ x j+1 and f (x j+1 ) ≤ x j . Thus [ f (x j ), f (x j+1 )]) ⊃ [x j , x j+1 ].
We will renumber the vertices of the P-graph (but not the points of P)
so that I1 = [x j , x j+1 ], where j = max{i: f (xi ) > xi }. By the proof of the
preceding lemma, I1 is a vertex with a directed edge from itself to itself.
For any two points xi < xk in P, define
k−1
fˆ([xi , xk]) = [ f (xl ), f (xl+1 )].
l=i
I1 In−1
I2 I3 I4 I5 In−2
I1 → I2 → · · · → Ik → I1 (7.1)
I1 → I1 → I2 → · · · → Ik → I1
I1 → I1 → · · · → I1 → I2 → · · · → Im−1 → I1
gives a periodic point of period q. The verification that these periodic points
have minimal period q is left as an exercise (Exercise 7.3.3).
LEMMA 7.3.8. If n is even, then f has a periodic point of minimal period 2.
Proof. Let m be the smallest even period of a non-fixed periodic point, and
let I1 be an interval of the associated partition that fˆ-covers itself. If no other
interval fˆ-covers I1 , then Lemma 7.3.5 implies that m = 2.
Suppose then that some other interval fˆ-covers I1 . In the proof of
Lemma 7.3.6, we used the hypothesis that n is odd only to conclude the
existence of such an interval. Thus the same argument as in the proof of that
lemma implies that the P-graph contains the paths
Then In−1 → In−2 → In−1 implies the existence of a periodic point of minimal
period 2.
Conclusion of the proof of the first assertion of the Sharkovsky
Theorem. There are two cases to consider:
1. n = 2k, k > 0. If q ≺ n, then q = 2l with 0 ≤ l < k. The case l = 0 is
l−1
trivial. If l > 0, then g = f q/2 = f 2 has a periodic point of period
2k−l+1 , so by Lemma 7.3.8, g has a non-fixed periodic point of period
2. This point is a fixed point for f q , i.e., it has period q for f . Since it
is not fixed by g, its minimal period is q.
168 7. Low-Dimensional Dynamics
k
2. n = p2k, p odd. The map f 2 has a periodic point of minimal period p,
k
so by Corollary 7.3.7, f 2 has periodic points of minimal period m for
all m ≥ p and all even m < p. Thus f has periodic points of minimal
period m2k for all m ≥ p and all even m < p. In particular, f has a
periodic point of minimal period 2k+1 , so by case 1, f has periodic
points of minimal period 2i for i = 0, . . . , k.
The next lemma finishes the proof of the Sharkovsky theorem.
LEMMA 7.3.9. For any α ∈ NSh , there is a continuous map f : [0, 1] → [0, 1]
such that MinPer( f ) = S(α).
Proof. We distinguish three cases:
1. α ∈ N, α odd,
2. α ∈ N, α even, and
3. α = 2∞ .
Case 1. Suppose n ∈ N is odd, and α = n. Choose points x0 , . . . , xn−1 ∈
[0, 1] such that
0 = xn−1 < · · · < x4 < x2 < x0 < x1 < x3 < · · · < xn−2 = 1,
f D(f ) D(D(f ))
Figure 7.4. Graphs of D ( f ) for f ≡ 1/2.
k
Exercise 7.3.2. Show that there are maps f, g: [0, 1] → [0, 1], each with a
periodic point of period n (for some n), such that the associated P-graphs
are not isomorphic. (Note that for n = 3, all P-graphs are isomorphic.)
Exercise 7.3.3. Verify that the periodic points in the proof of Corollary 7.3.7
have minimal period q.
Exercise 7.3.4. Show that the sequence {gk}k∈N defined near the end of the
proof of Lemma 7.3.9 converges uniformly, and the limit g∞ satisfies g∞ = gk
on [2/3k, 1].
1 Our arguments in this section follow in part those of [CE80] and [MT88].
7.4. Combinatorial Theory of Piecewise-Monotone Mappings 171
LEMMA 7.4.1. The itinerary i(x) is eventually periodic if and only if the
iterates of x converge to a periodic orbit of f .
Proof. If i(x) is eventually periodic, then by replacing x by one of its forward
iterates, we may assume that i(x) is periodic, of period p. If i j (x) = c j for
some j, then c j is periodic, and we are done. Thus we may assume that f k(x)
is contained in the interior of a lap of f for each k. For j = 0, . . . , p − 1,
let J j be the smallest closed interval containing { f k(x): k = j mod p}. Since
the itinerary is periodic of period p, each Ji is contained in a single lap,
so f : J j → J j+1 is strictly monotone. It follows that f p : J0 → J0 is strictly
monotone.
Suppose f p : J0 → J0 is increasing. If f p (x) ≥ x, then by induction,
f kp (x) ≥ f (k−1) p (x) for all k > 0, so { f kp (x)} converges to a point y ∈ J0 ,
which is fixed for f p . A similar argument holds if f p (x) < x.
If f p : J0 → J0 is decreasing, then f 2 p : J0 → J0 is increasing, and by the
argument in the preceding paragraph, the sequence { f 2kp (x)} converges to
a fixed point of f 2 p .
Conversely, suppose that f kq (x) → y as k → ∞, where f q (y) = y. If the
orbit of y does not contain any turning points, then eventually x has the same
172 7. Low-Dimensional Dynamics
itinerary as y. The case where O(y) does contain a turning point is left as an
exercise (Exercise 7.4.1).
Let be a function defined on {I1 , . . . , Il , c1 , . . . , cl } such that (I1 ) =
±1, (Ik) = (−1)k+1 (I1 ), and (ck) = 1 for k = 0, . . . , l. Associated to is a
signed lexicographic ordering ≺ on
, defined as follows. For s ∈
, define
τn (s) = (sk).
0≤k<n
−Il+1 < −cl < −Il < · · · < −I1 < I1 < c1 < I2 < · · · < cl < Il+1 .
To prove part 3, suppose the intervals I(x), f (I(x)), f 2 (I(x)), . . . are not
pairwise disjoint. Then there are integers n ≥ 0, p > 0, such that f n (I(x)) ∩
f n+ p (I(x))
= ∅. Thenf n+kp (I(x)) ∩ f n+(k+1) p (I(x))
=for
∅ all k ≥ 1. It fol-
lows that L = k≥1 f (I(x)) is a non-empty interval that contains no turn-
kp
We claim that Lt and Rt are non-empty. The proof of this claim breaks into
four cases according to the four possibilities for f |∂ I . We prove it in the case
f (c0 ) = f (cl+1 ) = c0 . Then ν0 = i(c0+ ) = (I1 , I1 , . . .), νl+1 = i(cl+1
−
) = (Il+1 ,
I1 , I1 , . . .), (I1 ) = 1, and (Il+1 ) = −1. Note that t
= i (c0 ) = ν0 and t
=
i(cl+1 ) = νl+1 , since t ∈ / i( f ). Thus ν0 ≺ t, so c0 ∈ Lt . If t0 < Il+1 , then t ≺
νl+1 , so cl+1 ∈ Rt , and we are done. So suppose t0 = Il+1 . If t1 > I1 , then
t ≺ νl+1 , and again we are done. If t1 = I1 , then condition 2 implies that
σ ν0 ≺ σ 2 t, which implies in turn that t ≺ νl+1 . Thus νl+1 ∈ Rt .
Let a = sup Lt . We will show that a ∈ / Lt . Suppose for a contradiction that
a ∈ Lt . Since x ∈ / Lt for all x > a, we conclude that i(a) ≺ t i(a + ). This
implies that the orbit of a contains a turning point. Let n ≥ 0 be the smallest
integer such that i n (a) = ck for some k ∈ {1, . . . , l}. Then i j (a) = t j = i j (a + )
for j = 1, . . . , n − 1, and i n (a + ) = Ik or i n (a + ) = Ik+1 . Suppose the latter
holds. Then f n is increasing on a neighborhood of a. Since i(a) ≺ t i(a + )
and i j (a) = t j = i j (a + ) for j = 0, . . . , n − 1, it follows that
and ck ≤ tn ≤ Ik+1 .
If tn = ck, then by condition 1, σ n (t) = i(ck), so t = i(a), contradicting
the fact that t ∈/ i( f ). Thus we may assume that tn = Ik+1 . If f is increasing
on Ik+1 , then condition 2 implies that σ n+1 (t) σ νk. But σ n (t) σ n (i(a + )),
τn (t) = +1 and tn = i n (a + ) imply that
REMARK 7.4.9. One can show that the following extension of Corollary 7.4.8
is also true: Let f and g be l-modal maps of I, and suppose f has no wandering
intervals, no attracting periodic points, and no intervals of periodic points. If
f and g have the same kneading invariants and endpoint itineraries, then f
and g are topologically semiconjugate.
new points { pN j } in that interval are equidistributed. Then we define the map
f N : I → I to be the piecewise-linear map connecting the points ( pnj , pn+1 j ),
j = 0, . . . , l + 1, n = 0, . . . , N − 1. It follows (Exercise 7.4.5) that:
1. f N is l-modal for each N > 0;
2. { f N } converges in the C 0 topology to an l-modal map f with turning
points c1 , . . . , cl ; and
3. the kneading invariants of f are ν1 , . . . , νl .
Exercise 7.4.1. Finish the proof of Lemma 7.4.1.
Exercise 7.4.3. Work out the ordering on the set of itineraries of the
quadratic map qµ for 2 < µ < 3.
Exercise 7.4.4. Show that the tent map has exactly 2n periodic points of
period n, and the set of periodic points is dense in [−1, 1].
Exercise 7.4.5. Verify the last three assertions in the proof of Theorem
7.4.11.
Exercise 7.5.2. Show that any polynomial with distinct real roots has neg-
ative Schwarzian derivative.
Exercise 7.5.3. Prove Lemma 7.5.1.
Exercise 7.5.4. Prove that each Möbius transformation has Schwarzian
derivative 0 and preserves the cross-ratios C and D.
Exercise 7.5.5. Prove that the action of the group of Möbius transforma-
tions on the extended real line is simply transitive on triples of points.
Exercise 7.5.6. Prove the remaining inequality of Proposition 7.5.5.
1.2
0.8
0.6
0.4
0.2
0 I0 a b I1 1
THEOREM 7.6.1. Let µ > 4. Then µ is a Cantor set, i.e., a perfect, nowhere
dense subset of [0, 1]. The restriction qµ |µ is topologically conjugate to the
one-sided shift σ :
2+ →
2+ .
Proof. Let a = 1/2 − 1/4 − 1/µ and b = 1/2 + 1/4 − 1/µ be the two
solutions of qµ (x) = 1, and let I0 = [0, a], I1 = [b, 1]. Then qµ (I0 ) = qµ (I1 ) =
I, and qµ ((a, b)) ∩ I = ∅ (see Figure 7.5). Observe that the images qµn (1/2)
of the critical point 1/2 lie outside I and tend to −∞. Therefore the two
inverse branches f0 : I → I0 and f1 : I → I1 and their compositions are well
defined. For k ∈ N, denote by Wk the set of all words of length k in the
alphabet {0, 1}. For w = ω1 ω2 . . . ωk ∈ Wk and j ∈ {0, 1}, set Iw j = f j (Iw ) and
gw = fωk ◦ · · · ◦ fω2 ◦ fω1 , so that Iw = gw (I).
the intersection n∈N Iω1 ...ωn consists of exactly one point h(ω). The
map h:
2+ → µ is a homeomorphism conjugating the shift σ and qµ |µ
(Exercise 7.6.2).
Exercise 7.6.1. Prove that if µ > 4 and 1/2 − 1/4 − 1/µ < x < 1/2 +
1/4 − 1/µ, then qµn (x) → −∞ as n → ∞.
two types of generic bifurcations: The saddle–node bifurcation (or the fold
bifurcation) may occur if the derivative at a periodic point is 1, and the
period-doubling bifurcation (or flip bifurcation) may occur if the derivative
at a periodic point is −1. We describe these bifurcations in the next two
propositions. See [CH82] or [HK91] for a more extensive discussion of bi-
furcation theory, or [GG73] for a thorough exposition on the closely related
topic of singularities of differentiable maps.
y=x
y
µ = µ0
µ > µ0
µ < µ0
fµ
x1 (µ) x0 x2 (µ) x
α (x0 ) > 0, statements 3 and 4 are satisfied for and δ sufficiently small
(Exercise 7.7.4).
REMARK 7.7.6. The stability of the fixed point ξ (µ) and of the periodic points
x1 (µ) and x2 (µ) depend on the signs of the derivatives in the third and fourth
hypotheses of Proposition 7.7.5. Proposition 7.7.5 deals with only one of the
four possible generic cases when the derivatives do not vanish. The other three
cases are similar, and we do not consider them here (Exercise 7.7.5).
Proof. Since
∂f
(x0 , µ0 ) = −1
=
1,
∂x
we can apply the implicit function theorem to f (x, µ) − x = 0 to obtain a
differentiable function ξ such that f (ξ (µ), µ) = ξ (µ) for µ close to µ0 and
ξ (µ0 ) = x0 . This proves statement 1.
Differentiating f (ξ (µ), µ) = ξ (µ) with respect to µ gives
d ∂f ∂f
f (ξ (µ), µ) = (ξ (µ), µ) + (ξ (µ), µ) · ξ (µ) = ξ (µ),
dµ ∂µ ∂x
and hence
∂f
∂µ
(ξ (µ), µ) 1 ∂f
ξ (µ) = ∂f
, ξ (µ0 ) = (x0 , µ0 ).
1 − ∂ x (ξ (µ), µ) 2 ∂µ
Therefore
d ∂f ∂2 f 1 ∂f ∂2 f
(ξ (µ), µ) = (x0 , µ0 ) + (x0 , µ0 ) · (x0 , µ0 ) = η,
dµ µ=µ0 ∂ x ∂µ ∂ x 2 ∂µ ∂ x2
and assumption 2 yields statement 2.
To prove statements 3 and 4 consider the change of variables y = x −
ξ (µ), 0 = x0 − ξ (µ0 ) and the function g(y, µ) = f ( f (y + ξ (µ), µ), µ) −
ξ (µ). Observe that fixed points of f ( f (·, µ), µ) correspond to solutions of
g(y, µ) = y. Moreover,
∂g ∂2g
g(0, µ) ≡ 0, (0, µ0 ) = 1, (0, µ0 ) = 0,
∂y ∂ y2
i.e., the graph of the second iterate of f (·, µ0 ) is tangent to the diagonal
at (x0 , µ0 ) with second derivative 0. (See Figure 7.7.) A direct calculation
shows that, by assumption 3, the third derivative does not vanish:
2
2
∂3g ∂3 f ∂f
(0, µ0 ) = −2 3 (x0 , µ0 ) − 3 (x0 , µ0 ) = ζ < 0.
∂ y3 ∂x ∂ x2
Therefore
1 3
g(y, µ0 ) = y + ζ y + o(y3 ).
3!
Since ξ (µ) is a fixed point of f (·, µ), we have that g(0, µ) ≡ 0 in an interval
188 7. Low-Dimensional Dynamics
µ < µ0 µ = µ0 µ > µ0
Figure 7.7. Period-doubling bifurcation: the graph of the second iterate.
Exercise 7.7.3. State the analog of Proposition 7.7.3 for the remaining
three generic cases when the derivatives from assumption 3 do not vanish.
Exercise 7.7.4. Prove statements 3 and 4 of Proposition 7.7.3.
Exercise 7.7.5. State the analog of Proposition 7.7.5 for the remaining three
generic cases when the derivatives from assumptions 3 and 4 do not vanish.
Exercise 7.7.6. Prove that a period-doubling bifurcation occurs for the
family fµ (x) = 1 − µx 2 at µ0 = 3/4, x0 = 2/3.
fµ1
B1
fµn
fµn+1 Bn
Bn+1
g W s (g)
W u (g) fµ
Figure 7.8. Fixed point and stable and unstable manifolds for the Feigenbaum
map .
Complex Dynamics
1 Many of the proofs in this chapter follow the corresponding arguments from [CG93].
191
192 8. Complex Dynamics
8.2 Examples
The global dynamics of a rational map R depends heavily on the behavior of
the critical points of R under its iterates. In most of the examples below, the
Fatou set consists of finitely many components, each of which is a basin of
attraction. Some of the assertions in the following examples will be proved
in later sections of this chapter. Proofs of most of the assertions that are not
proved here can be found in [CG93].
Let qa : C̄ → C̄ be the quadratic map qa (z) = z2 − a, and denote by S1
the unit circle {z ∈ C: |z| = 1}. The critical points of qa are 0 and ∞, and
the critical values are −a and ∞; if a
=0, the only superattracting periodic
(fixed) point is ∞. In the examples below, we observe drastically different
global dynamics depending on whether the critical point lies in the basin of
a finite attracting periodic point, or in the basin of ∞, or in the Julia set.
1. q0 (z) = z2 . There is a superattracting fixed point at 0, whose basin of
attraction is the open disk 1 = {z ∈ C: |z| < 1}, and a superattract-
ing fixed point at ∞, whose basin of attraction is the exterior of S1 .
There is also a repelling fixed point at 1, and for each n ∈ N there
are 2n repelling periodic points of period n on S1 . The Julia set is
S1 ; the Fatou set is the complement of S1 . The map q0 acts on S1 by
8.2. Examples 195
2 The pictures in this chapter were produced with mandelspawn; see http://www.araneus.fi/
gson/mandelspawn/.
196 8. Complex Dynamics
and hence the sequence φn converges uniformly in U if (|λ| + )2 < |λ|.
By construction, φn ( f (z)) = λφn+1 (z). Therefore the limit φ = limn→∞ φn
is the required conjugation.
COROLLARY 8.4.2. Let ζ be a repelling fixed point of a meromorphic map
f . Then there is a neighborhood U of ζ and an analytic map φ: U → C̄ that
conjugates f and z → λ(ζ )z in U, i.e., φ( f (z)) = λ(ζ )φ(z) for z ∈ U.
Proof. Apply Theorem 8.4.1 to the branch g of f −1 with g(ζ ) = ζ .
PROPOSITION 8.4.3. Let ζ be a fixed point of a meromorphic map f .
Assume that λ = f (ζ ) is not 0 and is not a root of 1, and suppose that an
analytic map φ conjugates f and z → λz. Then φ is unique up to multiplica-
tion by a constant.
8.4. Periodic Points 199
Exercise 8.4.3. Prove that every rational map R
=Id of degree d ≥ 1 has
d + 1 fixed points in C̄ counted with multiplicity.
The following proposition shows that the Julia set possesses self-similarity,
a characteristic property of fractal sets.
PROPOSITION 8.5.6. The Julia set of a rational map of degree >1 is perfect,
i.e., it does not have isolated points.
Proof. Exercise 8.5.3.
PROPOSITION 8.5.7. Let R: C̄ → C̄ be a rational map of degree >1. Then
J (R) is the closure of the set of repelling periodic points.
Proof. We will show that J (R) is contained in the closure of the set Per(R)
of the periodic points of R. The result will follow, since J (R) is perfect and
there are only finitely many non-repelling periodic points.
Suppose ζ ∈ J (R) has a neighborhood U that contains no periodic points,
no poles, and no critical values of R. Since the degree of R is >1, there are
distinct branches f and g of R−1 in U, and f (z)
=g(z), f (z)
=R n (z), and
g(z)
=R n (z) for all n ≥ 0 and all z ∈ U. Hence the family
The postcritical set of a rational map R is the union of the forward orbits
of all critical points of R, and is denoted CL(R).
Measure-Theoretic Entropy
where the minimum is taken over all permutations of m elements. The ax-
ioms of distance are satisfied by d (Exercise 9.1.1).
Partitions ζ and ζ are independent, and we write ζ ⊥ ζ , if µ(C ∩ C ) =
µ(C) · µ(C ) for all C ∈ ζ and C ∈ ζ .
For a transformation T and partition ξ = {C1 , . . . , Cm}, let T −1 (ξ ) =
{T (C1 ), . . . , T −1 (Cm)}.
−1
where |
i | <
, and from now on log denotes logarithm base 2 with 0 log 0 =
0. It follows that for µ-a.e. ω ∈
m
1 m
lim log µ(ηn (ω)) = qi log qi ,
n→∞ n
i=1
(recall that 0 log 0 = 0). Note that −x log x is a strictly concave continuous
function on [0, 1], i.e., if xi ≥ 0, λi ≥ 0, i = 1, . . . , n, and i λi = 1, then
n n n
− λi xi · log λi xi ≥ − λi xi log xi (9.1)
i=1 i=1 i=1
210 9. Measure-Theoretic Entropy
with equality if and only if all xi s are equal. For x ∈ X, let m(x, ζ ) denote
the measure of the element of ζ containing x. Then
H(ζ ) = − log m(x, ζ ) dµ.
X
and summation over j finishes the proof of the inequality. The equality is
achieved if and only if xi = κi j /µi does not depend on i for each j, which is
equivalent to the independence of ξ and η.
The entropy of a partition has a natural interpretation as the “average
information of the elements of the partition.” For example, suppose X
represents the set of all possible outcomes of an experiment, and µ is the
probability distribution of the outcomes. To extract information from the
experiment, we devise a measuring scheme that effectively partitions X into
finitely many observable subsets, or events, C1 , C2 , . . . , Cn . We define the
information of an event C to be I(C) = − log µ(C). This is a natural choice
given that the information should have the following properties:
Exercise 9.1.1. Prove: (i) d(ξ, η) ≥ 0 with equality if and only if ξ = η mod 0
and (ii) d(ξ, ζ ) ≤ d(ξ, η) + d(η, ζ ).
The quantity H(ξ |η) is the average entropy of the partition induced by ξ
on an element of η. If C(x) ∈ ξ and D(x) ∈ η are the elements containing x,
then
H(ξ |η) = − log µ(C(x)|D(x)) dµ.
X
µ(Ai ∩ Bj ∩ Ck)
H(ξ ∨ η|ζ ) = − µ(Ai ∩ Bj ∩ Ck) · log
i, j,k
µ(Ck)
µ(Ai ∩ Ck)
=− µ(Ai ∩ Bj ∩ Ck) log
i, j,k
µ(Ck)
µ(Ai ∩ Bj ∩ Ck)
− µ(Ai ∩ Bj ∩ Ck) log
i, j,k
µ(Ai ∩ Ck)
= H(ξ |ζ ) + H(η|ξ ∨ ζ ),
and the first equality follows. The second equality follows from the first one
with ζ = ν.
The remaining statements of Proposition 9.2.1 are left as exercises
(Exercise 9.2.1).
The function ρ, which is called the Rokhlin metric, defines a metric on the
space of equivalence classes of partitions (Exercise 9.2.2).
PROPOSITION 9.2.2. For every
> 0 and m ∈ N there is δ > 0 such that if
ξ and η are finite partitions with at most m elements and d(ξ, η) < δ, then
ρ(ξ, η) <
.
Proof ([KH95], Proposition 4.3.5). Let partitions, ξ = {Ci : 1 ≤ i ≤ m}, η =
m
{Di : 1 ≤ i ≤ m} satisfy d(ξ, η) = i=1 µ(Ci
Di ) = δ. We will estimate
H(η|ξ ) in terms of δ and m.
If µ(Ci ) > 0, set αi = µ(Ci \Di )/µ(Ci ). Then
µ(Ci ∩ Di )
−µ(Ci ∩ Di ) log ≤ −µ(Ci )(1 − αi ) log(1 − αi )
µ(Ci )
µ(Ci ∩ Dj )
− µ(Ci ∩ Dj ) log ≤ −µ(Ci )αi (log αi − log(m − 1)).
j=i
µ(Ci )
9.3. Entropy of a Measure-Preserving Transformation 213
Proof. Since H(ξ |η) ≥ H(ξ |ζ ) for η ≤ ζ , the sequence H ζ |T −1 (ζ n ) is
non-increasing in n. Since H(T −1 ξ ) = H(ξ ) and H(ξ ∨ η) = H(ξ ) + H(η|ξ ),
we get
H(ξ n |ηn ) = H(ξ ∨ T −1 (ξ n−1 )|ηn ) = H(ξ |ηn ) + H(T −1 (ξ n−1 )|ξ ∨ ηn )
≤ H(ξ |η) + H(T −1 (ξ n−1 )|ηn )
≤ H(ξ |η) + H(T −1 (ξ )|T −1 (η)) + H(T −2 (ξ n−2 )|ηn )
..
.
≤ nH(ξ |η).
Therefore
1 1
H(ξ n ) ≤ H(ηn ) + H(ξ |η),
n n
and statement 3 follows.
The remaining statements of Proposition 9.3.2 are left as exercises
(Exercise 9.3.2).
9.3. Entropy of a Measure-Preserving Transformation 215
m
µ Cin ∩ Di
H(η|ξn ) = − µ Cin ∩ Di · log n
i=1
µ Ci
n
m
n µ Cm+1 ∩ Dj
− µ Cm+1 ∩ Dj · log n
j=1
µ Cm+1
m
n µ Cin ∩ Dj
− µ Ci ∩ Dj · log .
i=1 j=i
µ Cin
The first sum tends to 0 because µ(Cin ∩ Di ) → µ(Cin ). The second and third
sums tend to 0 because µ(Cin ∩ Dj ) → 0 for j =i.
A sequence (ζn ) of finite partitions is called refining if ζn+1 ≥ ζn for n ∈ N.
A sequence (ζn ) of finite partitions is called generating if for every finite
partition ξ and every δ > 0 there is n0 ∈ N such that for every n ≥ n0 there
n
is a partition ξn with ξn ≤ i=1 ζi and d(ξn , ξ ) < δ, or equivalently if every
n
measurable set can be approximated by a union of elements of i=1 ζi for a
large enough n.
Every Lebesgue space has a generating sequence of finite partitions
(Exercise 9.3.3). If X is a compact metric space with a non-atomic Borel
measure µ, then a sequence of finite partitions ζn is generating if the maxi-
mal diameter of elements of ζn tends to 0 as n → ∞ (Exercise 9.3.4).
n
with melements such that ξn ≤ i=1 ζi and d(ξn , ξ ) < δ. By Proposition 9.2.2,
ζk = (ξk × ν) ∨ (µ × ηk)
Exercise 9.3.3. Prove that every Lebesgue space has a generating sequence
of partitions.
218 9. Measure-Theoretic Entropy
Shifts. Let σ :
m →
m be the one or two-sided shift on m symbols, and
m
let p = ( p1 , . . . , pm) be a non-negative vector with i=1 pi = 1. The vector
p defines a measure on the alphabet {1, 2, . . . , m}. The associated product
measure µ p on
m is called a Bernoulli measure. For a cylinder set, we have
,...,nk k
µ p C nj11,..., jk = p ji .
i=1
k−1
P C n,n+1,...,n+k
j0 , j1 ,..., jk = q j0 Aji ji+1 .
i=0
−1 n
For C = C 0j0 ∈ ξ and D = C 1,...,n
j1 ,..., jn ∈ σ (ξ ), we have
n−1
n−1
P(C ∩ D) = q j0 Aji ji+1 and P(D) = q j1 Aji ji+1 .
i=0 i=1
220 9. Measure-Theoretic Entropy
Thus
m
n−1
−1 q j0 Aj0 , j1
H(ξ |σ (ξ )) = −
n
q j0 Aji , ji+1 log
j0 , j1 ,..., jn =1 i=1
q j1
m
n−1
= − q j0 Aji , ji+1 (log Aj0 , j1 + log q j0 − log q j1 ).
j0 , j1 ,..., jn =1 i=1
(9.2)
n n
Using the identities i=1 qi Ai,k = qk and k=1 Ai,k = 1, we find that
n−1
q j0 Aji , ji+1 log Aj0 , j1 = q j0 Aj0 , j1 log Aj0 , j1 , (9.3)
j0 , j1 ,..., jn i=0 j0 , j1
n−1
q j0 Aji , ji+1 log q j0 = q0 log q j0 , (9.4)
j0 , j1 ,..., jn i=0 j0
m
n−1
q j0 Aji , ji+1 log q j1 = q j1 log q j1 . (9.5)
j0 , j1 ,..., jn =1 i=1 j1
There are many Markov measures for a given subshift. We now construct a
special Markov measure, called the Shannon–Parry measure, that maximizes
the entropy. By the results of the next section, a Markov measure maximizes
the entropy if and only if the metric entropy with respect to the measure is
the same as the topological entropy of the underlying subshift.
Let B be a primitive matrix of zeros and ones. Let λ be the largest positive
eigenvalue of B, and let q be a positive left eigenvector of B with eigenvalue
λ. Let v be a positive right eigenvector of B with eigenvalue λ normalized so
that q, v = 1. Let V be the diagonal matrix whose diagonal entries are the
coordinates of v, i.e., Vi j = δi j v j . Then A = λ−1 V −1 BV is a stochastic matrix:
all elements of Aare positive, and the rows sum to 1. The elements of A are
Ai j = λ−1 vi−1 Bi j v j . Let p = qV = (q1 v1 , . . . , qn vn ). Then p is a positive left
n
eigenvector of A with eigenvalue 1, and i=1 pi = q, v = 1.
The Markov measure P = PA, p is called the Shannon–Parry measure for
the subshift σ A. Recall that while P is defined on the full shift space
,
its support is the subspace
A. Thus h P (σ A) = h P (σ ). Using the properties
9.5. Variational Principle 221
1 The proof of the variational principle below follows the argument of M. Misiurewicz [Mis76];
see also [KH95] and [Pet89].
222 9. Measure-Theoretic Entropy
LEMMA 9.5.1. Let µ, ν ∈ M and t ∈ (0, 1). Then for any measurable parti-
tion of ξ of X,
t Hµ (ξ ) + (1 − t)Hν (ξ ) ≤ Htµ+(1−t)ν (ξ ).
Similarly,
LEMMA 9.5.3. Let En be an (n,
)-separated set, νn = (1/|En |) x∈En δx , and
n−1
µn = n1 i=0 f∗i νn . If µ is any weak∗ accumulation point of {µn }n∈N , then µ is
f -invariant and
1
lim log |En | ≤ hµ ( f ).
n→∞ n
it follows that
a(k)−1
log |En | = Hνn (ξ n ) ≤ Hνn f −(rq+k) ξ q + Hνn f −i ξ
r =0 i∈T
a(k)−1
≤ Hf∗rq+kνn (ξ q ) + 2q log |ξ |.
r =0
1 1
≥ lim (log |β n | − n log 2) ≥ lim Hµ (β n ) − log 2
n→∞ n n→∞ n
= hµ ( f, β) − log 2 ≥ hµ ( f, ξ ) − log 2 − 1
We conclude that hµ ( f ) = hµ ( f n )/n ≤ n1 (htop ( f n ) + log 2 + 1) for all
n > 0. Letting n → ∞, we see that hµ ( f ) ≤ htop ( f ) for all µ ∈ M, which
proves the theorem.
Exercise 9.5.1. Prove Lemma 9.5.1.
Exercise 9.5.2. Let f be an expansive map of a compact metric space with
expansiveness constant δ0 . Show that f has a measure of maximal entropy,
i.e., there is µ ∈ M f such that hµ ( f ) = htop ( f ). (Hint: Start with a measure
supported on an (n,
)-separated set, where
≤ δ0 .)
Bibliography
[AF91] Roy Adler and Leopold Flatto. Geodesic flows, interval maps, and symbolic
dynamics. Bull. Amer. Math. Soc. (N.S.), 25(2):229–334, 1991.
[Ahl73] Lars V. Ahlfors. Conformal invariants: topics in geometric function theory.
McGraw-Hill Series in Higher Mathematics. McGraw-Hill Book Co., New
York, 1973.
[Ano67] D. V. Anosov. Tangential fields of transversal foliations in y-systems. Math.
Notes, 2:818–823, 1967.
[Ano69] D. V. Anosov. Geodesic flows on closed Riemann manifolds with negative cur-
vature. American Mathematical Society, Providence, RI, 1969.
[Arc70] Ralph G. Archibald. An introduction to the theory of numbers. Charles E.
Merrill Publishing Co., Columbus, OH, 1970.
[AS67] D. V. Anosov and Ya. G. Sinai. Some smooth ergodic systems. Russian Math.
Surveys, 22(5):103–168, 1967.
[AW67] R. L. Adler and B. Weiss. Entropy, a complete metric invariant for automor-
phisms of the torus. Proc. Nat. Acad. Sci. U.S.A., 57:1573–1576, 1967.
[BC91] M. Benedicks and L. Carleson. The dynamics of the Hénon map. Ann. of Math.
(2), 133:73–169, 1991.
[Bea91] Alan F. Beardon. Iteration of rational functions. Springer-Verlag, New York,
1991.
[Ber96] Vitaly Bergelson. Ergodic Ramsey theory – an update. In Ergodic theory of Zd
actions (Warwick, 1993–1994), pages 1–61. Cambridge Univ. Press, Cambridge,
1996.
[Ber00] Vitaly Bergelson. Ergodic theory and Diophantine problems. In Topics in sym-
bolic dynamics and applications (Temuco, 1997), pages 167–205. Cambridge
Univ. Press, Cambridge, 2000.
[BG91] Carlos A. Berenstein and Roger Gay. Complex variables. Springer-Verlag,
New York, 1991.
[Bil65] Patrick Billingsley. Ergodic theory and information. John Wiley & Sons Inc.,
New York, 1965.
[BL70] R. Bowen and O. E. Lanford, III. Zeta functions of restrictions of the shift
transformation. In Global Analysis (Proc. Sympos. Pure Math., Vol. XIV,
Berkeley, CA, 1968), pages 43–49. American Mathematical Society, Prov-
idence, RI, 1970.
225
226 Bibliography
[KR99] K. H. Kim and F. W. Roush. The Williams conjecture is false for irreducible
subshifts. Ann. of Math. (2), 149(2):545–558, 1999.
[Kre85] Ulrich Krengel. Ergodic theorems. Walter de Gruyter & Co., Berlin, 1985.
With a supplement by Antoine Brunel.
[KvN32] B. O. Koopman and J. von Neumann. Dynamical systems of continuous spectra.
Proc. Nat. Acad. Sci. U.S.A., 18:255–263, 1932.
[Lan84] Oscar E. Lanford, III. A shorter proof of the existence of the Feigenbaum
fixed point. Comm. Math. Phys., 96(4):521–538, 1984.
[LM95] Douglas Lind and Brian Marcus. An introduction to symbolic dynamics and
coding. Cambridge University Press, Cambridge, 1995.
[Lor63] E. N. Lorenz. Deterministic non-periodic flow. J. Atmos. Sci., 20:130–141, 1963.
[LY75] Tien Yien Li and James A. Yorke. Period three implies chaos. Amer. Math.
Monthly, 82(10):985–992, 1975.
[Mañ88] Ricardo Mañé. A proof of the C 1 stability conjecture. Inst. Hautes Études Sci.
Publ. Math., 66:161–210, 1988.
[Mat68] John N. Mather. Characterization of Anosov diffeomorphisms. Nederl. Akad.
Wetensch. Proc. Ser. A 71: Indag. Math., 30:479–483, 1968.
[Mil65] J. Milnor. Topology from the differentiable viewpoint. University Press of
Virginia, Charlottesville, VA, 1965.
[Mis76] Michal Misiurewicz. A short proof of the variational principle for a n+ action
on a compact space. In International Conference on Dynamical Systems in
Mathematical Physics (Rennes, 1975), pages 147–157. Astérisque, No. 40. Soc.
Math. France, Paris, 1976.
[Mon27] P. Montel. Leçons sur les families normales de fonctions analytiques et leurs
applications. Gauthier-Villars, Paris, 1927.
[Moo66] C. C. Moore. Ergodicity of flows on homogeneous spaces. Amer. J. Math.,
88:154–178, 1966.
[MRS95] Brian Marcus, Ron M. Roth, and Paul H. Siegel. Modulation codes for digital
data storage. In Different aspects of coding theory (San Francisco, CA, 1995),
pages 41–94. American Mathematical Society, Providence, RI, 1995.
[MT88] John Milnor and William Thurston. On iterated maps of the interval. In Dy-
namical systems (College Park, MD, 1986–87), Lecture Notes in Mathematics,
volume 1342, pages 465–563. Springer-Verlag, Berlin, 1988.
[PdM82] Jacob Palis, Jr., and Welington de Melo. Geometric theory of dynamical sys-
tems. Springer-Verlag, New York, 1982. An introduction.
[Pes77] Ya. Pesin. Characteristic Lyapunov exponents and smooth ergodic theory.
Russian Math. Surveys, 32:4:55–114, 1977.
[Pet89] Karl Petersen. Ergodic theory. Cambridge University Press, Cambridge, 1989.
[Pon66] L. S. Pontryagin. Topological groups. Gordon and Breach Science Publishers,
Inc., New York, 1966. Translated from the second Russian edition by Arlen
Brown.
[Que87] Martine Queffélec. Substitution dynamical systems – spectral analysis. Lecture
Notes in Mathematics, volume 1294. Springer-Verlag, Berlin, 1987.
[Rob71] J. W. Robbin. A structural stability theorem. Ann. of Math. (2), 94:447–493,
1971.
[Rob76] Clark Robinson. Structural stability of C 1 diffeomorphisms. J. Differential
Equations, 22(1):28–73, 1976.
Bibliography 229
231
232 Index
cone differentiable
stable, 115 flow, 107
unstable, 115 manifold, 106, 137
conformal map, 153 map, 107
conjugacy, 3 vector field, 107
measure-theoretic, 70 differential equation, 19
topological, 28, 39, 55 Diffk(M), 138
conjugate, 3 Dirac measure, 86
constant-length substitution, 64 direct product, 3
continued fraction, 12, 91 directed graph, 8, 56, 67
continuous (semi)flow, 28 discrete spectrum, 94–97
continuous spectrum, 95, 97 discrete-time dynamical system, 1
continuous-time dynamical system, 2 dist, 140
convergence in density, 99 distal
convergent, 91 extension, 46
coordinate chart, 137 homeomorphism, 45
coordinate system, local, 106 points, 45
covering map, 153 distribution, 139
cov(n, , f ), 37 Hölder-continuous, 143
critical point, 192 integrable, 139
critical value, 192 stable, 108
cross-ratio, 180 unstable, 108
cross-section, of a flow, 22 Douady, 205
Curtis, 55 Douady–Hubbard theorem,
cutting and stacking, 96 206
Cvitanović–Feigenbaum equation, doubling operator, 169
189 dual group, 95
cylinder, 54 dynamical system, 2
cylinder set, 7, 78 chaotic, 23
continuous-time, 2
d-lim, 99 deterministic, 23
deg(R), 193 differentiable, 106
degree, of a rational map, 193 discrete-time, 1
dendrite, 196 ergodic, 4
Denjoy, 205 hyperbolic, 22, 106
example, 161 minimal, 4, 29
theorem, 160 symbolic, 54
dense topological, 28
forward orbit, 32 topologically mixing, 33
full orbit, 32 topologically transitive, 31
orbit, 23, 32 dynamics
dense, -, 4 symbolic, 6, 54
derivative transformation, 72 topological, 28
deterministic dynamical system,
23 edge shift, 57
diameter, 37 elementary strong shift equivalence,
of a partition, 218 62
diffeomorphism, 106 embedded submanifold, 139
Anosov, 130, 141 embedding of a subshift, 55
Axiom A, 133 endomorphism
group, 138 expanding, 5
of the circle, 160 group, 5
234 Index