• Non ci sono risultati.

This is a consequence of the following Proposition 4.1.1: Let f : X → X be an endomorphism of a probability space (X, A, µ)

N/A
N/A
Protected

Academic year: 2021

Condividi "This is a consequence of the following Proposition 4.1.1: Let f : X → X be an endomorphism of a probability space (X, A, µ)"

Copied!
23
0
0

Testo completo

(1)

Chapter 4

Ergodic theory

4.1 The Poincar´e recurrence theorem

Ergodic theory deals with dynamics in a measure space. Let us begin with a definition.

Definition 4.1.1: Let (X,A, µ) and (Y, B, ν) be two measure spaces, and f: X → Y a measurable map. We shall denote by fµ the measure on (Y,B) given by fµ(B) = µ¡

f−1(B)¢

for all B ∈ B. We say that f is measure-preserving if fµ = ν. If (X,A, µ) = (Y, B, ν) and f is measure-preserving, we say that µ is f -invariant, or that f is an endomorphism of (X,A, µ).

The first result in ergodic theory is Poincar´e’s recurrence theorem, which states that in a probability Borel measure space almost every point is recurrent. This is a consequence of the following

Proposition 4.1.1: Let f : X → X be an endomorphism of a probability space (X, A, µ). Given A ∈ A, let ˜A⊆ A be the set of points x ∈ A such that fj(x)∈ A for infinitely many j ∈ N, that is

A = A˜

\ n=0

[

j≥n

f−j(A).

Then ˜A∈ A and µ( ˜A) = µ(A).

Proof : Let Cn={x ∈ A | fj(x) /∈ A for all j ≥ n}. We clearly have A = A˜ \ [

n=1

Cn;

therefore it suffices to prove that Cn∈ A and µ(Cn) = 0 for all n≥ 1. Now, we have Cn = A\ [

j≥n

f−j(A), (4.1.1)

and thus Cn∈ A. Furthermore, (4.1.1) implies Cn [

j≥0

f−j(A)\ [

j≥n

f−j(A), and thus

µ(Cn)≤ µ

[

j≥0

f−j(A)

 − µ

[

j≥n

f−j(A)

 .

But since

[

j≥n

f−j(A) = f−n

[

j≥0

f−j(A)

and f is measure-preserving, it follows that µ

[

j≥0

f−j(A)

 = µ

[

j≥n

f−j(A)

 ,

and we are done.

(2)

As a consequence we have

Theorem 4.1.2: (Poincar´e’s recurrence theorem) Let X be a separable metric space, f : X → X a Borel- measurable map, and µ an f -invariant probability Borel measure on X. Then µ-almost every point of X is f -recurrent.

Proof : Since X is separable, we can find (exercise) a countable basis of open sets{Un}n∈Nsuch that

n→+∞lim diam(Un) = 0

and

∀m ≥ 0 [

n≥m

Un= X.

Set ˜Un ={x ∈ Un | fj(x)∈ Un for infinitely many j > 0}. Proposition 4.1.1 says that each ˜Un is a Borel set and µ(Un\ ˜Un) = 0. Put

X =˜

\ m=0

[

n≥m

U˜n,

so that x∈ ˜X if and only if there are infinitely many n≥ 0 so that fj(x)∈ Un for infinitely many j > 0.

We have

µ(X\ ˜X) = µ

[

m=0

X \ [

n≥m

U˜n

 = µ

[

m=0

[

n≥m

Un\ [

n≥m

U˜n

 ≤ µ

[

m=0

[

n≥m

(Un\ ˜Un)

 = 0.

So it suffices to show that every point of ˜X is recurrent. Given ε > 0, choose m≥ 0 so that diam(Un)≤ ε if n≥ m. If x ∈ ˜X, there must exist n≥ m such that x ∈ ˜Un, and thus fj(x)∈ Unfor infinitely many j ≥ 0.

But this implies d¡

x, fj(x)¢

< ε for infinitely many j≥ 0, and thus (being ε arbitrary) x is recurrent.

Example 4.1.1. If f : X→ X has a periodic point p ∈ X of exact period n, then

µ = 1 n

n−1X

j=0

δfj(p)

is an f -invariant Borel probability measure, where δxis the Dirac measure concentrated at x∈ X.

Example 4.1.2. The Lebesgue measure of S1 is Rα-invariant for any α∈ R.

Exercise 4.1.1. Let (X,A, µ) and (Y, B, ν) be σ-finite measure spaces, and f: X → Y a measurable map.

Prove if there exists a generating sub-algebraB0⊆ B such that µ¡

f−1(B)¢

= ν(B) for all B∈ B0 then f is measure-preserving.

Example 4.1.3. The Lebesgue measure of S1is Em-invariant for any m∈ Z, because the inverse image of an interval of length ` small enough is the union of|m| disjoint intervals of length `/|m|, and we can apply the previous exercise. Poincar´e’s recurrence theorem applied to E10 then says that there exists a Borel set X ⊂ S1 of full measure such that for all j≥ 1 the decimal expansion of each x ∈ X contains the sequence of its first j digits infinitely many times.

Example 4.1.4. The Gauss transformation. Let φ: [0, 1]→ [0, 1] be given by

φ(x) = (1

x

¹1 x º

if x6= 0,

0 if x = 0.

(3)

This map has important connections with the theory of continuous fractions. We claim that φ preserves the Borel probability measure µ given by

µ(A) = 1 log 2

Z

A

1 1 + xdx

for every Borel set A⊆ [0, 1]. Using the Exercise 4.1.1, it suffices to prove that φ preserves the measures of finite unions of intervals, and hence it suffices to check that µ¡

φ−1([a, b])¢

= µ([a, b]) for every inter- val [a, b]⊆ [0, 1]. Since µ is absolutely continuous with respect to the Lebesgue measure, it suffices to check that µ¡

φ−1([0, b])¢

= µ([0, b]) for every b∈ [0, 1]. But indeed we have

φ−1([0, b]) = [ n=1

· 1 b + n,1

n

¸ ,

and so

µ¡

φ−1([0, b])¢

= X n=1

µ µ· 1

b + n,1 n

¸¶

= X n=1

1 log 2

Z 1/n

1/(b+n)

1 1 + tdt =

X n=1

1

log 2 log 1 + 1n 1 +b+n1

= 1

log 2 X m=1

µ

log n + 1

b + n + 1 − log n b + n

= 1

log 2log(b + 1)

= 1

log 2 Z b

0

1

1 + tdt = µ([0, b]).

Example 4.1.5. Assume we have given a probability measure on the set{0, . . . , N − 1}, that is a sequence of non-negative numbers p0, . . . , pN−1≥ 0 so that p0+· · ·+pN−1= 1. The associated Bernoulli (or product) measure on ΩN and Ω+N is the probability Borel measure µ defined on the cylinders by

µ(Cam11...a...mrr) = pa1· · · par.

It is clearly (full) left shift-invariant. A slightly more involved construction yields another kind of invariant measure. A stochastic matrix is a matrix S ∈ MN,N(R+), with row and column indexes running from 0 to N − 1, such that PN−1

i=0 sij = 1 for j = 0, . . . , N − 1. If S is a stochastic matrix, it is possible to prove that there exists a vector p = (p0, . . . , pN−1)∈ RN with non-negative coordinates such that Sp = p and p0+· · · + pN−1 = 1; furthermore, p is unique if S is transitive (in the sense of Definition 1.7.9). Then the Markov measure µS,p on ΩN is defined on cylinders by

µS,p(Cam11...a...mrr) = sa1a2· · · sar−1arpar. Again, it is easy to check that it is shift-invariant.

Exercise 4.1.2. Let f : X → X be an endomorphism of a measure space (X, A, µ). Given A ∈ A with non-zero measure, put ˜A = {x ∈ A | fj(x) ∈ A for infinitely many j ≥ 1}, and define ˜f : ˜A → A by set- ting ˜f (x) = fj(x), where j is the least positive index such that fj(x) ∈ A ( ˜f is sometimes called the first-return map associated to f ). Prove that ˜f ( ˜A)⊆ ˜A, that µ¡A˜\ ˜f ( ˜A)¢

= 0, and that µ|A˜is ˜f -invariant.

Exercise 4.1.3. (Hopf) Let X be a locally compact separable metric space and f : X→ X a Borel-measurable map. If µ is a locally finite f -invariant Borel measure, prove that µ-almost every x∈ X is recurrent or has empty ω-limit set. (Hint: since µ is locally finite, Lusin’s theorem yields a sequence{Xi} of open sets with finite measure whose union has full measure. For every Xi let ˜Xi be defined as in the previous exercise.

Applying Theorem 4.1.2 to ˜f : ˜Xi → ˜Xi deduce that almost every point of ˜Xi is recurrent. Conclude that almost every point of Xi is recurrent or has ω-limit set disjoint from Xj.)

(4)

Definition 4.1.2: Let µ be a Borel measure on a topological space X. The support of µ is the closed set supp µ = X\[

{A | A open in X, µ(A) = 0}.

Lemma 4.1.3: Let µ be a Borel measure on a topological space X. Then:

(i) if X is second countable then µ(X\ supp µ) = 0;

(ii) any set of full measure is dense in supp µ;

(iii) if f : X → X is continuous and µ is f-invariant then supp µ is f-invariant;

(iv) if X is a separable metric space, f : X→ X is continuous, and µ is an f-invariant probability measure, then supp µ is contained in the closure of recurrent points, and hence supp µ⊆ NW (f).

Proof : (i) Indeed, we can write X\ supp µ as countable union of measure zero open sets.

(ii) Assume that C is not dense in supp µ; this means that there exist x∈ supp µ and an open neighbour- hood A of x disjoint from C. Since A∩ supp µ 6= 0, we must have µ(A) > 0; therefore µ(X \ C) ≥ µ(A) > 0, and so C is not of full measure.

(iii) Since f is continous and µ is f -invariant, we clearly have f−1(X\ supp µ) ⊆ X \ supp µ. There- fore f−1(supp µ)⊇ supp µ, and hence supp µ is f-invariant.

(iv) Theorem 4.1.2 says that the set of recurrent points has full measure; hence it is dense in supp µ, and we are done.

4.2 Existence of invariant measures

A natural question now is whether invariant measures exist. We have a positive answer in the case of continuous self-maps of compact metric spaces. To prove it let us recall a couple of definitions and theorems.

Definition 4.2.1: Let A be a σ-algebra on a set X. We shall denote by MA(X) the set of probability measures onA. If X is a topological space and A is the σ-algebra of Borel subsets, we shall write M(X) instead ofMA(X). If f : X→ X is a A-measurable map, we shall denote by MfA(X) the set of f -invariant probability measures onA, and by Mf(X) the set of f -invariant Borel probability measures.

The Borel measures on compact Hausdorff spaces have a nice relationship with the dual of the space of continuous functions:

Theorem 4.2.1: (Riesz representation theorem) Let X be a compact Hausdorff space. Then:

(i) for every bounded linear functional T on C0(X) there exists a unique pair of mutually singular finite Borel measures µ and ν such that T (ϕ) =R

Xϕ dµR

Xϕ dν for all ϕ∈ C0(X).

(ii) a bounded linear functional T on C0(X) is positive (that is, T (ϕ) ≥ 0 if ϕ ≥ 0) if and only if there exists a unique finite Borel measure µ such that T (ϕ) =R

Xϕ dµ for all ϕ∈ C0(X).

As a consequence, if X is a compact Hausdorff space we can realize M(X) as the subset of the unit ball of the dual of C0(X) consisting of the positive functionals. Now, the dual of a normed vector space is endowed with the weak-∗ topology, i.e., the topology of pointwise convergence (a sequence {Tj} of functionals on the normed space V converges to a functional T if and only if Tj(v)→ T (v) for all v ∈ V ), and we have the famous

Theorem 4.2.2: (Banach-Alaoglu) The unit ball (with respect to the dual norm) of the dual of a normed vector space is weak-∗ compact.

This is actually a consequence of Tychonoff’s theorem on the product of compact spaces. Indeed, if B is the unit ball in the normed vector space V , we can realize the unit ball of the dual as a closed (with respect to the product topology) subset of [−1, 1]B, which is compact by Tychonoff’s theorem. Since the topology of pointwise convergence coincides with the product topology, we are done.

Since the set of positive linear functionals on C0(X) is clearly closed with respect to the weak-∗ topology, as a corollary we have

(5)

Corollary 4.2.3: Let X be a compact Hausdorff space. ThenM(X) is compact in the weak-∗ topology.

Exercise 4.2.1. Let X be a compact metric space, and take a countable subset{gj}j∈N⊂ C0(X) dense in the unit ball of C0(X). Given µ, ν∈ M(X) put

d(µ, ν) = X j=0

1 2j

¯¯¯¯Z

X

gj Z

X

gj¯¯

¯¯.

Prove that d is a metric onM(X) inducing the weak-∗ topology.

We are now able to prove the existence of invariant Borel measures:

Theorem 4.2.4: (Krylov-Bogolubov) Let f : X → X be a continuous self-map of a compact metric space X.

Then there exists an f -invariant probability Borel measure.

Proof : Let f:M(X) → M(X) be defined by (fµ)(A) = µ¡

f−1(A)¢

for any Borel subset A ⊆ X. In particular we have

∀ϕ ∈ C0(X)

Z

X

◦ f) dµ = Z

X

ϕ d(fµ),

and thus f is continuous in the weak-∗ topology. Our aim is to find a fixed point of f. Take µ0∈ M(X), and let µn∈ M(X) be defined by

µn= 1 n + 1

Xn m=0

(f)mµ0.

Since, by Corollary 4.2.3 and Exercise 4.2.1, M(X) is a compact metric space, we can extract a subse- quencenk} converging to µ ∈ M(X). Now,

fµnk= 1 nk+ 1

nk

X

m=0

(f)m+1µ0= 1 nk+ 1

nk

X

m=0

(f)mµ0 1

nk+ 1µ0+ 1

nk+ 1(f)nk+1µ0. The last two terms converge to 0 when k goes to infinity; hence

fµ = lim

k→∞fµnk = lim

k→+∞

1 nk+ 1

nk

X

m=0

(f)mµ0= lim

k→∞µnk= µ, and we are done.

Exercise 4.2.2. Let X be a compact Hausdorff space, and f : X → X a continuous map. Prove that a bounded positive linear functional T on C0(X) is of the form T (ϕ) =R

Xϕ dµ for a f -invariant finite Borel measure µ∈ Mf(X) if and only if T (ϕ◦ f) = T (ϕ) for all ϕ ∈ C0(X).

Exercise 4.2.3. Let f : [0, 1]→ [0, 1] be given by

f (x) =

½1

2x if 0 < x≤ 1, 1 if x = 0.

Prove that f is Borel measurable and that there are no f -invariant probability Borel measures on [0, 1].

(6)

4.3 Ergodic measures

In general, there may exist several probability measures invariant under the action of a given continuous map (remember the Example 4.1.1). We would like to single out measures related to the whole dynamics of the map, and not only to part of it.

Definition 4.3.1: Let f : X → X be a continuous map of a topological space. An f-invariant Borel measure µ is ergodic if every completely f -invariant Borel set A ⊆ X is either of zero or of full measure, that is, either µ(A) = 0 or µ(X\ A) = 0.

There is a particular case when this happen:

Definition 4.3.2: We shall say that a continuous map f : X → X of a topological space is uniquely ergodic if there exists one and only one f -invariant Borel probability measure.

The unique invariant measure of a uniquely ergodic map is ergodic. To prove it we need a definition:

Definition 4.3.3: Let (X,A, µ) be a measure space, and A ∈ A with 0 < µ(A) < ∞. Then the conditional measure associated to A is the probability measure µA defined by

∀B ∈ A µA(B) = µ(B∩ A)

µ(A) . We shall sometimes write µ(B| A) instead of µA(B).

Proposition 4.3.1: Let f : X → X a uniquely ergodic continuous map. Then the unique f-invariant Borel probability measure µ is ergodic.

Proof : If µ is not ergodic we can find a completely f -invariant Borel set A so that 0 < µ(A), µ(X\ A) < 1.

Then µA and µX\A are f -invariant Borel probability measures, and they are distinct because µA(A) = 1 whereas µX\A(A) = 0.

Lemma 4.3.2: Let µ be a finite f -invariant measure for a continuous self-map f of a topological space X.

Then µ is ergodic if and only if every ϕ∈ L1(X, µ) which is f -invariant it is constant µ-almost everywhere.

Proof : If µ is not ergodic, then there is completely f -invariant Borel set A with µ(A), µ(X\ A) 6= 0. Then the characteristic function χAof A is an element of L1(X, µ) which is f -invariant and not constant µ-almost everywhere.

Conversely, assume µ ergodic, and let f ∈ L1(X, µ) be f -invariant. Then for every c ∈ R the set Ac ={x ∈ X | f(x) ≤ c} is completely f-invariant; therefore either µ(Ac) = 0 or µ(Ac) = µ(X), and, since X =S

c∈QAc, we must have µ(Ac) = µ(X) for at least some c∈ R. Analogously, since ∅ =T

c∈QAc, we must have µ(Ac) = 0 for at least some c∈ R. Let

c0= inf{c ∈ R | µ(Ac) = µ(X)} > −∞.

Then

µ ({x ∈ X | f(x) < c0}) = µ

[

c<c0 c∈Q

Ac

 = 0,

and thus

µ({x ∈ X | f(x) = c0}) = µ(Ac)− µ ({x ∈ X | f(x) < c0}) = µ(X), that is f≡ c0µ-almost everywhere.

The arguments used in the previous section also yields the existence of ergodic measures. To see this, we need the following

Lemma 4.3.3: Let f : X→ X be a continuous map of a topological space X. If µ ∈ Mf(X) is not ergodic then there exist µ1, µ2∈ Mf(X) and 0 < t < 1 so that µ16= µ2 and µ = tµ1+ (1− t)µ2.

Proof : Let A⊂ X be a completely f-invariant Borel set such that 0 < µ(A) < 1, and set µ1= µA, µ2= µX\A and t = µ(A). Then µ16= µ2and µ = tµ1+ (1− t)µ2.

(7)

Definition 4.3.4: Let V be a vector space, and K ⊂ V a closed convex subset. An extreme point of K is a point v ∈ K such that if we can write v = tv1+ (1− t)v2 with v1, v2 ∈ K and 0 < t < 1 then necessarily v1= v2= v.

So extreme points ofMf(X) are ergodic measures (the converse is true too; see Exercise 4.3.1). Thus to prove the existence of ergodic measures it suffices to prove the existence of extreme points:

Theorem 4.3.4: Let f : X→ X be a continuous self-map of a compact metric space X. Then there exists an ergodic f -invariant probability Borel measure.

Proof : Letj}j∈Nbe a countable dense set in C0(X), and defineMj by settingM0=Mf(X) and Mj+1=

½

µ∈ Mj

¯¯¯¯Z

X

ϕjdµ = max

ν∈Mj

Z

X

ϕj

¾ . Since ν 7→R

Xϕjdν is continuous (and recalling Theorem 4.2.4) every Mj is a not empty convex compact set; hence their intersectionG is not empty. To conclude the proof it suffices to show that every µ ∈ G is an extreme point ofMf(X).

Take µ∈ G, and write µ = tµ1+ (1− t)µ2with 0 < t < 1 and µ1, µ2∈ Mf(X). Then

∀ϕ ∈ C0(X)

Z

X

ϕ dµ = t Z

X

ϕ dµ1+ (1− t) Z

X

ϕ dµ2. Since µ∈ M1, this implies

t Z

X

ϕ01+ (1− t) Z

X

ϕ02= Z

X

ϕ0≥ max

½Z

X

ϕ01, Z

X

ϕ02

¾ , which is possible if and only if R

Xϕ0dµ =R

Xϕ01 =R

Xϕ02; in particular, µ1, µ2 ∈ M1. Repeating this argument, by induction we see that

Z

X

ϕjdµ = Z

X

ϕj1= Z

X

ϕj2 for every j ∈ N. But {ϕj} is dense in C0(X); hence R

Xϕ dµ = R

Xϕ dµ1 = R

Xϕ dµ2 for all ϕ∈ C0(X).

The uniqueness statement in Theorem 4.2.1.(ii) then implies µ1 = µ2 = µ, and so µ is an extreme point ofMf(X), as desired.

Actually, it is possible to use convex analysis (more precisely, Choquet’s theorem on extreme points) to prove a sort of decomposition of any invariant measure as integral or ergodic ones:

Theorem 4.3.5: Every invariant Borel probability measure µ for a continuous self-map f of a compact metric space X can be decomposed into an integral of ergodic invariant Borel probability measures in the following sense: there is a partition (modulo null sets) of X into completely f -invariant subsets{Xα}α∈A, where A is a measure space endowed with a probability measure dα (actually, A can be taken a Lebesgue space), so that every Xα carries an ergodic f -invariant measure µα and we have

∀ϕ ∈ C0(X)

Z

X

ϕ dµ = Z

A

Z

Xα

ϕ dµαdα.

Two different ergodic measures are related to completely different parts of the dynamics:

Proposition 4.3.6: Two distinct ergodic measures for a continuous self-map f of a topological space X are mutually singular.

Proof : Let µ and µ1 be two f -invariant ergodic measures, and assume that µ1 is not absolutely continuous with respect to µ. Then there is a Borel set A such that µ(A) = 0 but µ1(A) > 0. Let

A =˜ \

n≥0

[

j≥n

f−j(A)

be the set of points whose orbit intersects A infinitely many times. Then ˜A is a completely f -invariant subset such that µ( ˜A) = 0 and µ1( ˜A) > 0 (compare Proposition 4.1.1). By ergodicity, it follows that X\ ˜A has full µ-measure while ˜A has full µ1-measure, and hence µ and µ1 are mutually singular.

If, on the other hand, µ1<< µ, the Radon-Nikodym derivative dµ1/dµ is f -invariant, and thus µ-almost everywhere constant (by Lemma 4.3.2), which implies µ1= µ.

(8)

Remark 4.3.1. Notice that the last argument in the previous proof shows that if µ1 is f -invariant, µ is ergodic and µ1<< µ then µ1= µ.

Exercise 4.3.1. Prove that a measure f∈ Mf(X) is ergodic if and only if it is an extreme point ofMf(X).

There is a stronger notion of ergodicity:

Definition 4.3.5: An f -invariant measure µ is mixing if

n→∞lim µ¡

f−n(A)∩ B¢

= µ(A)µ(B) (4.3.1)

for every pair of measurable sets A and B.

Lemma 4.3.7: Any mixing measure is ergodic.

Proof : Indeed, if A⊆ X is completely f-invariant we have µ(A)µ(X\ A) = lim

n→∞µ¡

f−n(A)∩ X \ A¢

= µ¡

A∩ (X \ A)¢

= 0, and hence µ(A) = 0 or µ(X\ A) = 0.

Remark 4.3.2. In the next section we shall show that an f -invariant probability measure µ is ergodic if and only if

n→∞lim 1 n

n−1X

m=0

µ¡

f−m(A)∩ B¢

= µ(A)µ(B) for every pair of measurable sets A and B.

Ergodicity is the statistical equivalent of topological transitivity, unique ergodicity of minimality, and mixing of topological mixing. We shall give a better explanation of this in the next section, but meanwhile we can prove the following:

Proposition 4.3.8: Let f : X → X be a homeomorphism of a locally compact Hausdorff space with a countable basis of open sets and no isolated points. Assume that there exists an ergodic measure µ with supp µ = X. Then f is topologically transitive.

Proof : Since supp µ = X, every open set has positive measure. But, since µ is ergodic, there cannot be two not empty disjoint completely f -invariant open subsets of X and, by Corollary 1.4.5, f is topologically transitive.

Remark 4.3.3. If f is a homeomorphism, it is easy to check that the support of an f -invariant measure is completely f -invariant. Then if an ergodic measure µ gives no mass to isolated points the previous argument shows that f is topologically transitive when restricted to the support of µ. In the next section we shall show that this is true even when f is not a homeomorphism.

Remark 4.3.4. Not every f -invariant measure for a topologically transitive map is ergodic. For instance, if µ1and µ2are Bernoulli measures on Ω2, they are ergodic for the shift (they are mixing; see Example 4.3.5);

on the other hand, µ = (µ1+ µ2)/2 is still shift-invariant and of full support, but it is not ergodic, because of Proposition 4.3.6.

Proposition 4.3.9: Let f : X→ X be a uniquely ergodic continuous self-map of a compact metric space X, and µ the unique element ofMf(X). Then f is minimal on supp µ.

Proof : Take x∈ supp µ, and let Λ ⊆ supp µ be the closure of the orbit of x. Since Λ is clearly f-invariant, Theorem 4.2.4 yields an f|Λ-invariant Borel measure ν. Define then a Borel measure ˜ν on X by set- ting ˜ν(A) = ν(A∩ Λ) for every Borel set A. Since Λ is f-invariant we have

˜ ν¡

f−1(A)¢

= ν¡

f−1(A)∩ Λ) = ν¡

(f|Λ)−1(A∩ Λ)¢

= ν(A∩ Λ) = ˜ν(A),

and hence ˜ν is f -invariant. By unique ergodicity, ˜ν = µ; but since supp ˜ν⊆ Λ, this implies Λ = supp µ, and we are done.

(9)

Exercise 4.3.2. Let f : X → X be a continuous map of a topological space X. Assume that there exists a mixing measure µ. Prove that f is topologically mixing on supp µ.

We shall now describe some examples of ergodic and uniquely ergodic maps. We begin with some generalities on topological groups.

Definition 4.3.6: A Borel measure on a topological group G invariant under all left (respectively, right) translations will be said left (respectively, right) invariant.

We have the following important (and deep)

Theorem 4.3.10: (Haar) Let G be a locally compact topological group. Then there exists a unique (up to multiplicative constants) locally finite left invariant Borel measure on G.

Definition 4.3.7: Let G be a compact topological group. The unique (by the previous theorem) Borel left invariant probability measure on G is called the Haar measure of G.

Example 4.3.1. The usual Lebesgue measure on the torusTn (the one induced by the Lebesgue measure ofRn), suitably normalized, is the Haar measure ofTn.

Lemma 4.3.11: The Haar measure µ of a compact group G is also right invariant, and invariant under continuous homomorphisms.

Proof : Given x∈ G let ν = (Rx)µ, where Rxis the right translation by x. Then ν is left invariant, because for all y∈ G and Borel sets A we have

ν¡

L−1y (A)¢

= µ¡

R−1x L−1y (A)¢

= µ¡

L−1y R−1x (A)¢

= µ¡

R−1x (A)¢

= ν(A),

where Ly is the left translation by y. Since we clearly have ν(G) = 1, the uniqueness of the Haar measure yields ν = µ, and thus µ is right invariant.

Analogously, if H: G→ G is a continuous homomorphism, setting ν = Hµ the same argument shows that ν is left-invariant with ν(G) = 1, and hence ν = µ, as desired.

Exercise 4.3.3. Prove that the support of the Haar measure of a compact group G is the whole group G.

Theorem 4.3.12: Let G be a compact abelian group, µ its Haar measure, and Lx: G→ G a left translation.

Then the following properties are equivalent:

(i) Lx is topologically transitive;

(ii) Lx is minimal;

(iii) µ is Lx-ergodic;

(iv) Lx is uniquely ergodic;

(v) the orbit of x is dense.

Proof : (i)⇒ (ii) Lemma 1.4.2.

(ii)⇒ (v) Obvious.

(v)⇒ (iv) Take ν ∈ MLx(G); it suffices to prove that ν is the Haar measure of G. Given y∈ G, choose a sequence{nj} ⊂ N such that Lnxj−1(x) = njx converges to y. We shall prove that

∀ϕ ∈ C0(G) lim

j→∞

Z

G

ϕ◦ Lnjxdν = Z

G

ϕ◦ Lydν. (4.3.2)

By uniform continuity of ϕ, for every ε > 0 there is a neighbourhood U of the identity element e such that

|ϕ(y1)− ϕ(y2)| ≤ ε for every y2∈ y1+ U . Now,

|(ϕ ◦ Lnjx)(z)− (ϕ ◦ Ly)(z)| = |ϕ(z + njx)− ϕ(z + y)|;

if j is large enough we have njx− y ∈ U, and thus kϕ ◦ Lnjx− ϕ ◦ Lyk≤ ε, and (4.3.2) follows. But then Z

G

ϕ d(Ly)ν = Z

G

ϕ◦ Lydν = lim

j→∞

Z

G

ϕ◦ Lnjxdν = lim

j→∞

Z

G

ϕ d(Lnxj)ν = Z

G

ϕ dν.

(10)

But this means that ν is left invariant, and hence it is the Haar measure of G, as claimed.

(iv)⇒ (iii) Proposition 4.3.1.

(iii)⇒ (i) If G is a finite group, then the Haar measure is the normalized counting measure. If x has period strictly less than the cardinality of G, then the orbit of x is a completely Lx-invariant subset with measure strictly greater than zero and strictly less than one, against the ergodicity assumption. So the orbit of x must be the whole of G, and Lx is topologically transitive.

If G has an isolated point, then all points are isolated, and thus G (being compact) must be finite. So we can assume that G has no isolated points, and in this case the assertion follows from Proposition 4.3.11, because it is easy to see that supp µ = G (exercise).

Corollary 4.3.13: (i) Irrational rotations of S1 are uniquely ergodic.

(ii) The Lebesgue measure of Tn is ergodic for a translation Tγ:Tn → Tn if and only if 1, γ1, . . . , γn are rationally independent if and only if Tγ is uniquely ergodic.

Proof : It follows from Propositions 1.4.1, 1.4.7 and 4.3.12.

Example 4.3.2. The Lebesgue measure µ of S1is ergodic for the maps Em: S1→ S1with|m| > 1. Indeed, let ϕ∈ L1(S1, µ) be f -invariant, and let

ϕ(x) =

+

X

k=−∞

ϕkexp(2πikx)

its expansion in Fourier series. Since ϕ◦ Em= ϕ almost everywhere, it follows that ϕmk= ϕk for all k∈ Z.

Butk| → 0 as |k| → +∞; hence necessarily ϕk = 0 for k6= 0, and thus ϕ is constant. The assertion then follows from Lemma 4.3.2.

We end this section with a list of examples of mixing measures, together with an (unproved) proposition useful in proving that a measure is mixing.

Definition 4.3.8: A collectionC of measurable sets in a measure space (X, A, µ) is dense if for every A ∈ A and any ε > 0 there is A0 ∈ C so that µ(A4A0) < ε, where A4A0 = (A\ A0)∪ (A0\ A) is the symmetric difference. More generally, a collection C ⊂ A is said sufficient if the family of finite disjoint unions of elements ofC is dense.

Proposition 4.3.14: Let f : X→ X be an endomorphism of a measure space (X, A, µ). Then:

(i) if (4.3.1) holds for all A and B belonging to a sufficient collection of measurable sets, then µ is mixing;

(ii) µ is mixing if and only if a given complete system Φ of functions in L2(X, µ) and any ϕ, ψ∈ Φ one has

n→+∞lim Z

X

◦ fn)ψ dµ = Z

X

ϕ dµ· Z

X

ψ dµ. (4.3.3)

Example 4.3.3. The Lebesgue measure of the torus is never mixing with respect to translations Tγ, while it is mixing for hyperbolic automorphisms ofT2.

Example 4.3.4. The Lebesgue measure of S1 is mixing with respect to the maps Em with|m| > 1.

Example 4.3.5. The Bernoulli measure is mixing with respect to the full shift both in ΩN and in Ω+N.

4.4 Birkhoff ’s theorem

The main result of this section is the following

Theorem 4.4.1: (Birkhoff) Let f : X→ X be an endomorphism of a probability space (X, A, µ). Then for every ϕ∈ L1(X, µ) the following limit

ϕf(x) = lim

n→+∞

1 n

n−1X

j=0

ϕ¡ fj(x)¢

Riferimenti

Documenti correlati

[r]

Siciak, Transfinite diameter and the analytic continuation of functions of two complex variables, in Studies in Mathematical Analysis and Re- lated Topics, pp.. 3

[r]

(b) No: the unique plane containing the first three points has equation 3x − 2z = −1, and the fourth point is not contained

Otherwise, you can compute a basis as in the

Solution proposed by Roberto Tauraso, Dipartimento di Matematica, Universit`a di Roma “Tor Vergata”, via della Ricerca Scientifica, 00133 Roma,

This is a contradiction because a positive integer is the sum of two squares if and only if all the prime factors congruent to 3 mod 4 have an even exponent in

Solution proposed by Roberto Tauraso, Dipartimento di Matematica, Universit`a di Roma “Tor Vergata”, via della Ricerca Scientifica, 00133 Roma, Italy.. The answer