• Aucun résultat trouvé

Quantitative recurrence in two-dimensional extended processes

N/A
N/A
Protected

Academic year: 2022

Partager "Quantitative recurrence in two-dimensional extended processes"

Copied!
20
0
0

Texte intégral

(1)

www.imstat.org/aihp 2009, Vol. 45, No. 4, 1065–1084

DOI: 10.1214/08-AIHP195

© Association des Publications de l’Institut Henri Poincaré, 2009

Quantitative recurrence in two-dimensional extended processes

Françoise Pène

1

and Benoît Saussol

aUniversité Européenne de Bretagne, Université de Brest, laboratoire de Mathématiques, CNRS UMR 6205, France.

E-mail: francoise.pene@univ-brest.fr; benoit.saussol@univ-brest.fr Received 21 January 2008; revised 22 September 2008; accepted 17 October 2008

Abstract. Under some mild condition, a random walk in the plane is recurrent. In particular each trajectory is dense, and a natural question is how much time one needs to approach a given small neighbourhood of the origin. We address this question in the case of some extended dynamical systems similar to planar random walks, includingZ2-extension of mixing subshifts of finite type. We define a pointwise recurrence rate and relate it to the dimension of the process, and establish a result of convergence in distribution of the rescaled return time near the origin.

Résumé. Sous certaines conditions, une marche aléatoire dans le plan est récurrente. En particulier, chaque trajectoire est dense, et il est naturel d’estimer le temps nécessaire pour revenir dans un petit voisinage de l’origine. Nous nous intéressons à cette question dans le cas de systèmes dynamiques étendus similaires à des marches aléatoires planaires, notamment celui desZ2-extension de sous-shifts de type fini mélangeants. Nous déterminons une vitesse de convergence ponctuelle que nous relions à la dimension du processus et nous établissons un résultat de convergence en loi du temps de retour à l’origine, correctement normalisé.

MSC:37B20; 37A50; 60Fxx

Keywords:Return time; Random walk; Subshift of finite type; Recurrence; Local limit theorem

1. Introduction 1.1. Motivation

Let us consider a recurrent random walk or a recurrent dynamical system (with finite orσ-finite invariant measure).

Given an initial condition, say x, we thus know that the process will return back ε close to its starting point x. A basic question iswhen? For finite measure preserving dynamical systems this question has some deep relations to the Hausdorff dimension of the invariant measure. Namely, ifτε(x)represents this time, in many situations

τε(x)≈ 1 εdim

for typical pointsx, where dim is the Hausdorff dimension of the underlying invariant measure. This has been proved for interval maps [15] and rapidly mixing systems [1,14]. Another type of result is the exponential distribution of rescaled return times and the lognormal fluctuations of the return times [4,9].

In this paper we are dealing with systems where the underlying natural measure is indeed infinite. This causes the return time to be non-integrable, in contrast with the finite measure case. However, the systems we are thinking about have in common the property that, in some sense, the behaviours at small scale and at large scale are independent.

The large scale dynamics being some kind of recurrent random walk, and the small scale dynamics a finite measure

1Supported in part by the ANR project TEMI (Théorie Ergodique en mesure infinie).

(2)

Table 1

Recurrence forZk-extensions.Bεdenotes the ball of radiusε,νis a Gibbs measure on the base anddis the Hausdorff dimension ofν

Dimension Z0-extension Z1-extension Z2-extension

Scale lim

ε→0 logτε

logε=d lim ε→0

logτε

logε =d lim ε→0

loglogτε

logε =d

Local law ν(ν(Bεε> t )e−t lim

ε→0ν(ν(Bε)τε> t )1+βt1 ν(ν(Bε)logτε> t )1+βt1

Lognormal fluctuations εdτε εdτε εdlogτε

preserving system. We provide results in three different cases: two probabilistic models (with some hypothesis of independence) and an infinite measure dynamical system (a Z2-extension of a finite measure dynamical system).

The first case treated in Section 2 is a toy model designed to give the hint of the general case. Then, in Section 3 we consider the case of planar random walks. Finally, in Section 4 we give a complete analysis of the quantitative behaviour of return times in the case ofZ2-extensions of subshifts of finite type. The recurrence forZ2-extensions of subshifts of finite type essentially comes from Guivarc’h and Hardy’s local limit theorem [8]. It was proven afterwards by Conze [6] and Schmidt [16] that it is also a consequence of the central limit theorem.

1.2. Description of the main result:Z2-extensions of subshifts of finite type

We emphasize that this dimension 2 is at the threshold between recurrent and non-recurrent processes, since in higher dimension these processes are not recurrent (except if degenerate). It makes sense to show how our results behave with respect to the dimension. For completeness, we call the non-extended system itself aZ0-extension. In this non- extended case, the type of results we have in mind (see Table 1) have already been established respectively by Ornstein and Weiss [13], Hirata [9] and Collet, Galves and Schmitt [4]. The case ofZ2-extension is completely done in Sec- tion 4. The case ofZ1-extension can be partly derived following the technique used in the present paper.2The essential difference is that the local limit theorem has the one-dimensional scaling in 1

n, instead of 1n in the two-dimensional case. The following table summarizes the different results as the dimension changes. The first line of results cor- responds to Theorem 8, the second to Theorem 9 and the third to Corollary 10. We refer to Section 4 for precise statements.

2. A toy model in dimension two

We present a toy model designed to posses a lot of independence. It has the advantage of giving the right formulas with elementary proofs.

2.1. Description of the model and statement of the results

Let us consider two sequences of independent identically distributed random variables(Xn)n1and(Yn)n0indepen- dent one from the other such that:

• the random variableX1is uniformly distributed on{(1,0), (−1,0), (0,1), (0,−1)};

• the random variableY0is uniformly distributed on[0,1]2. Let us notice thatSn:=n

k=1Xk(with the conventionS0:=0) is the symmetric random walk onZ2. We study a kind of random walkMnonR2given byMn=Sn+Yn.

2ForZ1-extensions, the method easily gives the three results presented in the table. However, the result on the local law is not sharp. The determi- nation of the precise law will be the object of a forthcoming paper.

(3)

Another representation of our model could be the following. LetS=Rdand consider the systemZ2×S. Attached to each sitei∈Z2of the lattice, there is a local system which lives onSandσnis a i.i.d. sequence ofS-valued random variables with some densityρ, independent of theXn’s. Then we look at the random walk(Sn, σn), thinking atσnas a spin.

We want to study the asymptotic behaviour, as ε goes to zero, of the return time in the open ballB(M0, ε)of radiusεcentered atM0(for the euclidean metric). Let

τε:=min

m≥1: |MmM0|< ε . We will prove the following:

Theorem 1. Almost surely,log loglogετε converges to the dimension2of the Lebesgue measure onR2asεgoes to zero.

Theorem 2. For allt≥0we have:

εlim0P λ

B(M0, ε)

logτεt

= 1 1+π/t.

2.2. Proof of the pointwise convergence of the recurrence rate to the dimension First, let us defineR1:=min{m≥1: Sm=0}. According to [7], we know that we have:

P(R1> s)∼ π

logs assgoes to infinity. (1)

We then define for anyp≥0 thepth return timeRpin[0,1)2by induction:

Rp+1:=inf{m > Rp: Sm=0}.

Observe thatRp is thepth return time at the origin of the random walkSn on the lattice, thus the delays between successive return timesRpRp−1, settingR0=0, are independent and identically distributed. Consequently:

P(RpRp1> s)=P(R1> s). (2)

The proof of Theorem 1 follows from these two lemmas.

Lemma 3. Almost surely,log loglognRn→1asn→ ∞.

Proof. It suffices to prove that for any 0< α <1, almost surely, en1αRn≤2nen1+α providedn is sufficiently large. By independence and Eq. (2) we have

P

logRnn1α

≤P

∀p≤n,log(RpRp1)n1α

=P

logR1n1αn . According to the asymptotic formula (1), fornsufficiently large

P

logR1n1αn

1− π

2n1α n

≤e−πnα/2.

The first inequality follows then from the Borel–Cantelli lemma.

Moreover, according to formulas (2) and (1), we have

n1P(log(RnRn1) > n1+α)) <+∞. Hence, by the Borel–Cantelli lemma, we know that almost surely, for allnsufficiently large, we haveRnRn1≤en1+α. From this

we get the second inequality.

(4)

LetTε:=min

≥1:|YRY0|< ε . Lemma 4. Almost surely,logλ(B(YlogTε

0,ε))→1asε→0.

Proof. Since(Y)is an i.i.d. sequence independent of theXks, the random variableTε has a geometric distribution with parameterλε:=λ(B(Y0, ε))ε2. For anyα >0 we have the simple decomposition

P logTε

−logλε−1 > α

=P

Tε> λε1α +P

Tε< λε1+α . The first term is directly handled by Markov inequality:

P

Tε> λε1α

λαε,

while the second term may be computed using the geometric distribution:

P

Tε< λε1+α

=1−(1λε)λ−1+αε

=1−exp λε1+α

log(1−λε)

≤ − λε1+α

log(1−λε)

=O λαε

.

Let us defineεn:=n1/α. According to the Borel–Cantelli lemma, logTlogλεn

εn converges almost surely to the constant 1.

The conclusion follows from the facts thatn)n≥1is a decreasing sequence of real numbers satisfying limn→+∞εn= 0 and limn→+∞ εn

εn+1 =1, andTεis monotone inε.

Proof of Theorem 1. Observe that wheneverB(M0, ε)is contained in(0,1)2, we haveτε=RTε. The theorem follows from Lemmas 3 and 4 since

log logτε

−logε =log logRTε

logTε

logTε

−logλε

logλε

logε →1×1×2

almost surely asε→0.

2.3. Proof of the convergence in distribution of the rescaled return time Proof of Theorem 2. Observe thatlogRlogτε

converges almost surely to 1. Lett >0. By independence ofTεand theRn

we have

Fε(t ):=P λ

B(Y0, ε)

logRTεt

=

n1

P(Tε=n)P

logRnt λε

.

SinceTεhas a geometric law with parameterλε,Fε(t )is equal toGλε(t )with:

Gδ(t ):=

n1

δ(1δ)n1P

logRnt δ

.

First, we notice that the independence of the times between the successive returns gives for anyu >0 that P(Rnu)≤P

k=max1,...,nRkRk1u

=P(R1u)n.

(5)

Letα <1. Using the inequality above and the equivalence relation (1) we get that for anyδ >0 sufficiently small, Gδ(t )

n≥1

δ(1δ)n1

1−απδ t

n

= 1

1+απ/t +O(δ).

This implies that lim supε0Fε(t )11/t.

FixA >0 and keeping the same notations observe that we haveFε(t )Hλε(t )with:

Hδ(t ):=

1nA/δ

δ(1δ)n1P

logRnt δ

.

Note that the independence gives in addition that for anyu >0 P(Rnu)≥P

k=max1,...,nRkRk1u/n

=P(R1u/n)n.

Letα >1. Using the inequality above and the equivalence relation (1) we get that for sufficiently smallδ >0 Hδ(t )

1nA/δ

δ(1δ)n−1

1−α π

t /δ−logn n

1nA/δ

δ(1δ)n−1

1−α2πδ t

n

.

Evaluating the limit whenδ→0 of the geometric sum and then lettingA→ ∞we end up with lim infδ0Hδ(t )

1

1+π/t, which gives the result.

3. Random walk on the plane

3.1. Almost sure convergence

We now consider a true random walk onR2,Sn=X1+ · · · +Xnwhere theXi’s are i.i.d. random variables distributed with a lawμof zero mean, with (invertible) covariance matrixΣ2and characteristic functionμ(t )ˆ =

eit·xdμ(x).

Letτεbe the first return time of the walk in theε-neighbourhood of the origin:

τε:=min

n≥1:|Sn|< ε .

LetΩ= {∀n≥1, Sn=0}. We notice that outsideΩthe return timeτε is obviously bounded by the first time n≥1 for whichSn=0. Therefore, we will only consider the asymptotic behaviour ofτεonΩ.

Theorem 5. Assume additionally that the distributionμsatisfies the Cramer condition lim sup

|t|→∞

μ(t )ˆ <1.

Then onΩwe havelimε0log logτε

logε =2almost surely.

We remark that a kind of Cramer’s condition on the law is necessary, since there exist some planar recurrent random walks for which the statement of the theorem is false (the return time being even larger than expected). We discovered after the completion of the proof of this theorem that its statement is contained in Theorem 2 of Cheliotis’s recent paper [5]. For completeness we describe the strategy of our original proof. A key point is a uniform version of the local limit theorem. Indeed we need an estimation of the typeP(|Sn|< ε)n2, with some uniformity inε(for some c >0). We will follow the classical proof of the local limit theorem (see Theorem 10.17 of [3]) to get the following (see Section 3.3 for its proof):

(6)

Lemma 6. There existsa >0,ε0>0,an integerN,a sequenceκn→0such that,for anyn > N,anyε(0, ε0)and for anyA∈R2we have

P

SnA+ [−ε, ε]2

γ ε2 n exp

Σ2A·A 2n

ε2κn

n +exp(−an) ε2 , withγ= 2

π detΣ2.

Then, this information on the probability of return is strong enough to estimate the first return time to the ε- neighbourhood of the origin.

Proof of Theorem 5. For any α >12, usingεn=1/logαn, we get thatP(|Sn|< εn)is summable. By the Borel–

Cantelli lemma, for almost everyxΩ, for allnlarge enough, we haveτεn(x) > nand thus:

lim inf

n→∞

log logτεn(x)

−logεn ≥lim inf

n→∞

log logn log logαn=1

α,

which implies by monotonicity and the fact thatαis arbitrary that lim infε0log logτε(x)

logε ≥2.

Letα < 12. To control the lim sup, we will take n=nε = exp(ε1/α). Letγ< γ andm=(logε)4. We use a similar decomposition to that of Dvoretsky and Erdös in [7]. LetAk= {|Sk|< εand∀p=k+1, . . . , n,|SpSk|>

2ε}. TheAk’s are disjoint, hence by independence and invariance, and according to Lemma 6, forεsmall enough, we have:

1≥ n k=m

P(Ak)n k=m

P

|Sk|< ε

P> nk)n k=m

P

|Sk|< ε

P> n)n k=m

γε2

k P> n).

Hence we havePε > n)21logn1/α2, if n is large enough. Let εp =p2/(α2). By the Borel–Cantelli lemma we have logτεpεp1/αalmost surely, hence lim supp→∞log logτlogεεp

pα1. By monotonicity and the fact thatα

is arbitrary we get the result.

3.2. Limit distribution of return times in neighbourhoods

We introduce a slight modification of the model. Letε >0. LetM0εbe uniformly distributed on the ballB(0, ε)and independent3of the sequence(Sn)introduced in the Section 3.1.

We define the random walkMnε=M0ε+Sn. We are interested in the limiting distribution of the first return time in the ballB(0, ε):

˜ τε=inf

n≥1: Mnε < ε .

Theorem 7. Under the hypotheses of Theorem5,the random variableε2logτ˜εconverges in distribution to a random variableY with distributionP(Y > t )= P)

1+2

det(Σ−2)P)t (t >0).

Proof. To simplify notations, we omit the indicesε.

Upper bound:FixR >0 and an integerK >0. Let Γ =

i=1, . . . , K, Mi=M0and|Mi| ≤R .

3Using if necessary a larger probability space, letUbe a random variable uniformly distributed in the ballB(0,1)and independent of the sequence (Sn). Then one can takeM0ε=εU.

(7)

Using the same decomposition as in the proof of Theorem 5, we have P(Γ )=

n k=0

P Γ

|Mk|< ε−2νand∀=k+1, . . . , n,|M| ≥ε−2ν ,

where nis equal to exp(t

ε2)for somet >0. Let ν=ε2 andA=21Z2B(0, ε−3ν). Notice that the sets Qa:=(aν/

2, a+ν/

2)2withaAare pairwise disjoints and contained inB(0, ε−2ν). Letm=(logε)4. Hence we have

P(Γ )≥P Γ

=1, . . . , n|M| ≥ε +

n k=m

a∈A

P Γ

MkQaand∀=k+1, . . . , n, M/B(0, ε−2ν)

≥P Γ

=1, . . . , n, |M| ≥ε +

n k=m

a∈A

P Γ

MkQaand∀=k+1, . . . , n, SSk/B(a, εν)

≥P

Γ ∩ { ˜τε> n} +

n k=m

a∈A

P

Γ ∩ {MkQa} P

=1, . . . , n, S/B(a, εν)

by independence (sincem > K wheneverεis sufficiently small). Note that P

Γ ∩ {MkQa}

=

{∀i,xi=x0,|xi|≤R}P(SkK∈ −xK+Qa)dP(M0,...,MK)(x0, . . . , xK).

According to Lemma 6 we get

km P

Γ ∩ {MkQa}

≥P(Γ )γν2 2k

for some fixedγ< γ, providedεis sufficiently small. Hence we have n

k=m

a∈A

P

Γ ∩ {MkQa} P

=1, . . . , n, S/B(a, εν)

n k=m

a∈A

P(Γ )γν2 2k P

=1, . . . , n, M/B(0, ε)|M0Qa

n k=m

a∈A

P(Γ )γπε2

k P(M0Qa)P

=1, . . . , n, M/B(0, ε)|M0Qa

n k=m

P(Γ )γπε2 k

P

=1, . . . , n, M/B(0, ε)

−P

ε−4ν≤ |M0| ≤ε

≥P(Γ )γπε2log(n)

1+o(1)

P˜ε> n)−o(1).

Therefore sinceε2logn=t+o(1)we get P(Γ )≥P

Γ ∩ { ˜τε> n}

+γπtP(Γ )P˜ε> n)+o(1)

≥P

{ ˜τε> n}

−P

∃1≤iK,|Mi|> R

+γπtP(Γ )P˜ε> n)+o(1).

(8)

Hence we have lim sup

ε0

P

˜ τε>exp

t ε2

≤P(Γ )+P(∃1≤iK,|Mi|> R) 1+γπtP(Γ ) . TakingR→ ∞first thenK→ ∞we obtain

lim sup

ε0

P

˜ τε>exp

t ε2

≤ P) 1+γπtP). This holds for allγ< γ, which gives the upper bound.

Lower bound: We only provide a sketch proof since the arguments are very similar. Let m= (logε)4, n= exp(t /ε2),ν=ε2. We have:

P(Γ )≤P

Γ ∩ { ˜τε> nlogn} +

m k=1

pk+

nlognn k=m

pk+

nlogn k=nlognn

pk,

wherepk=P ∩ {|Mk|< ε+2νand∀=k+1, . . . , n,|M| ≥ε+2ν}).

By Theorem 5 we have m

k=1

pk≤P

Γ ∩ {τm}

≤P

Γ \Ω +o(1)

since almost sure convergence implies convergence in probability.

Next, takingA=21Z2B(0, ε+3ν),Q¯a:= [aν/

2, a+ν/

2]2and using the same arguments as for the upper bound we get

nlognn k=m

pk

nlognn k=m

a∈A

P Γ

Mk∈ ¯Qaand∀=k+1, . . . , n, M/B(0, ε−2ν)

≤P(Γ )γπε2log(nlognn)P˜ε> n)+o(1), where nowγ> γ.

Finally,

nlogn k=nlognn

pk

nlogn k=nlognn

P Γ

|Mk| ≤ε+2ν

nlogn k=nlognn

P(Γ )γ2 k =o(1).

Putting these estimates together we end up with P(Γ )≤P

Γ ∩ { ˜τε> n}

+o(1)+P

Γ \Ω

+P(Γ )γπtP˜ε> n).

This gives lim inf

ε0 P

˜ τε>exp

t ε2

≥P(Γ )−P \Ω)+o(1) 1+γπtP(Γ ) .

And the conclusion follows as for the upper bound.

(9)

3.3. Proof of Lemma6

Proof. FixA∈R2. Letε >0. Letn≥2. For any 0< a < b, letha,b(t )be the piecewise affine function equal to 1 on [−a, a], to 0 outside[−b, b]and with slopeb1a in between. LetHa,b(x1, x2)=ha,b(x1)ha,b(x2). Let us notice that we haveHεε/n,ε≤1[−ε,ε]2Hε,ε+ε/nand thus

E

Hεε/n,ε(SnA)

≤P

|SnA|ε

≤E

Hε,ε+ε/n(SnA) .

For anyHL1(R2)we define its Fourier transformHˆ and its inverse Fourier transformHˇ by:

u∈R2 H (u)ˆ = 1 2π

R2H (x)exp(−ixu)dx and H (u)ˇ = ˆH (u).

With this definition, we immediately get that ˆHa,bL = 21πHa,bL1 = (a+2πb)2 and after an easy computation ˆHL1(a+b+b−a4 )2.

Let us notice that we have for allx∈R2,Ha,b(x)= ˇˆHa,b(x). Hence with the use of Fubini’s theorem we have E

Ha,b(SnA)

=EHˇˆa,b(SnA)

=E 1

R2

Hˆa,b(u)exp

iu(SnA) du

= 1 2π

R2

Hˆa,b(u)E exp

iu(SnA) du

= 1 2π

R2

Hˆa,b(u) E

exp(iuX1)n

eiuAdu.

Denote the characteristic function ofX1byφ (u)=E[exp(iuX1)]. Because of the hypothesis on the distribution ofX1, we know that, ifuis non-null, then|φ (u)|<1. Let us recall that we have:

φ (u)=1−1

2Σ2u·u+g(u)Σ2u·u, with lim

u0g(u)=0.

Let us fix someβ >0 such that:

uB(0, β) φ (u)

1−1

2Σ2u·u ≤1

4Σ2u·u and such that:

usup2

φ (u) <1−1 4αβ2,

whereαis the smallest modulus of the eigenvalues ofΣ2. Now we take(a, b)=(ε, ε+εn)or(a, b)=εn, ε).

Hence we have:

u2>βn1/6

Hˆa,b(u) φ (u)n

eiuAdu

≤ ˆHa,bL1

1−αβ2n1/3 4

n

3ε+4n ε

2

exp

αβ2n2/3 4

. (3)

(10)

Hence, it remains to estimate the following quantity:

B(0,βn−1/6)

Hˆa,b(u) φ (u)n

eiuAdu=1 n

B(0,βn1/3)

Hˆa,b

v/n

φ v/

nn

eivA/ndv.

We will compare this quantity to 1n

R2Hˆa,b(0)exp(−12Σ2v·v)eivA/ndv.First we have:

1 n

B(0,βn1/3)

Hˆa,b v/

n

eivA/n

φ v/

nn

−exp

−1 2Σ2v·v

dv

≤1 n

B(0,βn1/3)

Hˆa,b

v/n nexp

n−1 4n Σ2v·v

1 ng1

v

n

Σ2v·v

dv

≤ ˆHa,bL

n sup

wB(0,βn1/6)

g1(w)

R2exp

−1

8Σ2v·v

Σ2v·v dv

≤4ε2

n sup

wB(0,βn1/6)

g1(w)

R2exp

−1

8Σ2v·v

Σ2v·v dv≤ ε2

n, (4)

with limu0g1(u)=0 and limn→+∞κn=0.

Second we have:

1 n

B(0,βn1/3)

exp

−1

2Σ2v·v Hˆa,b

v/n

− ˆHa,b(0) dv

≤1

n sup

wB(0,βn1/6)

∇ ˆHa,b(w)

B(0,βn1/3)

|v|2

nexp

−1 2Σ2v·v

dv

≤1 n

2ελ(B(0,2ε)) 2π

R2

|v|2

nexp

−1 2Σ2v·v

dv≤ε2

n, (5)

with limn→+∞κn=0.

Third we have:

1 n

R2\B(0,βn1/3)

exp

−1

2Σ2v·v Hˆa,b(0) dv≤ε2

n, (6)

with limn→+∞κn=0.

Hence, sinceHˆa,b(0)=21πHa,bL1,to estimateE(Ha,b(SnA))we are led to study the quantity Ha,bL1

n4π2

R2exp

−ivA/√ n

exp

−1 2Σ2v·v

dv=Ha,bL1 n4π2

2πexp(−Σ2A·A/(2n))

√detΣ2 .

To conclude we notice that|Ha,bL1−4ε2| ≤10εn2.

4. Case of Euclidean extensions of subshifts of finite type

4.1. Description of theZ2-extensions of a mixing subshift

Let us fix a finite setAcalled alphabet. Let us consider a matrixMindexed byA×Awith 0–1 entries. We suppose that there exists a positive integern0 such that each component ofMn0 is non-zero. We define the set of allowed sequencesΣas follows:

Σ:=

ω:=n)n∈Z:∀n∈Z, M(ωn, ωn+1)=1 .

(11)

Fig. 1. Dynamics of theZ2-extensionFof the shift.

We endowΣwith the metricd given by d

ω, ω

:=em,

wheremis the greatest integer such thatωi=ωiwhenever|i|< m. We define the shiftθ:ΣΣbyθ ((ωn)n∈Z)= n+1)n∈Z. For any functionf:Σ→Rwe denote bySnf=n1

=0fθits ergodic sum. Let us consider an Hölder continuous functionϕ:Σ→Z2. We define theZ2-extensionF of the shiftθby (see Fig. 1)

F:Σ×Z2Σ×Z2, (x, m)

θ x, m+ϕ(x) .

We want to know the time needed for a typical orbit starting at(x, m)Σ×Z2to returnε-close to the initial point after iterations of the mapF. By translation invariance we can assume that the orbit starts in the cellm=0. More precisely, let

τε(x)=min

n≥1: Fn(x,0)∈B(x, ε)× {0}

. Observe thatFn(x, m)=nx, m+Snϕ(x)), thus

τε(x)=min

n≥1: Snϕ(x)=0 andd θnx, x

< ε .

Letν be the Gibbs measure associated to some Hölder continuous potentialh, and denote byσh2the asymptotic variance ofhunder the measureν. Recall thatσh2vanishes if and only ifhis cohomologous to a constant, and in this caseνis the unique measure of maximal entropy.

We know that there exists a positive integerm0such that the functionϕis constant on eachm0-cylinder.

Let us denote byσϕ2the asymptotic covariance matrix ofϕ:

σϕ2= lim

n→+∞Covν

1

nSnϕ

.

We suppose that

Σϕdν=0.

We add the following hypothesis of non-arithmeticity onϕ: We suppose that, for anyu∈ [−π;π]2\ {(0,0)}, the only solutions(λ, g), withλ∈Candg:Σ→Cmeasurable with|g| =1, of the functional equation

gσ g=λeiu·ϕ

is the trivial oneλ=1 andg=const.

(12)

Note that this condition implies thatσϕ2is invertible. If this non-arithmeticity condition was not satisfied, then the range ofSnϕ would be essentially contained in a sub-lattice and we could make a change of variable (and eventually of dimension) to reduce our study to the corresponding twistedZa-extension (whereais the rank ofσϕ2). Moreover, we emphasize that this non-arithmeticity condition is equivalent to the fact that all the circle extensionsTudefined by Tu(x, t )=(θ (x), t+u·ϕ(x))are weakly-mixing foru∈ [−π;π]2\ {(0,0)}.

In this context, we prove the following results:

Theorem 8. The sequence of random variables log logτlogεε converges almost surely asε→0to the Hausdorff dimen- siondof the measureν.

Theorem 9. The sequence of random variables ν(Bε(·))logτε(·)converges in distribution asε→0to a random variable with distribution functiont1+βtβt1(0;+∞)(t ),withβ:= 1

2π detσϕ2.

Corollary 10. If the measure ν is not the measure of maximal entropy, then the sequence of random variables

log logτ ε+dlogε

logε converges in distribution asε→0to a centered Gaussian random variable of varianceh2.

In the case whereνis the measure of maximal entropy,then the sequence of random variablesεdlogτεconverges in distribution to a finite mixture of the law found in the previous theorem,that is,there exists some probability vector α=n)and positive constantsβnsuch that the sequence of random variablesεdlogτεconverges in distribution to a random variable with distribution function

nαn1+βnβt

nt1(0;+∞)(t ).

Example 11. We provide an example where the function ϕ(x) only depends on the first coordinate x0, that is, ϕ(x)=ϕ(x0).On the shiftΣ= {I, E, N, W, S}Zthe functionϕgiven byϕ(I )=(0,0),ϕ(E)=(1,0),ϕ(N )=(0,1), ϕ(W )=(−1,0)andϕ(S)=(0,−1)fulfills the hypothesis.

The remaining part of the section is devoted to the proof of these results. In Section 4.2 we recall some preliminary results and prove a uniform conditional local limit theorem. In Section 4.3 we prove Theorem 8 and in Section 4.4 we prove Theorem 9 and Corollary 10.

4.2. Spectral analysis of the Perron–Frobenius operator and local limit theorem

In order to exploit the spectral properties of the Perron–Frobenius operator we quotient out the “past.” We define:

Σˆ :=

ω:=n)n∈N: ∀n∈N, M(ωn, ωn+1)=1 , dˆ

n)n0, ωn

n0

:=e−ˆr(ω,ω)

withr((ωˆ n)n0, (ωn)n0)=inf{m≥0:ωm=ωm}and θˆ

n)n0

=n+1)n0.

Let us define the canonical projectionΠ:Σ→ ˆΣbyπ((ωn)n∈Z)=n)n0. Letνˆbe the image probability measure (onΣ) ofˆ νbyΠ. There exists a functionψ:Σˆ →Z2such thatψΠ=ϕθm0.

Let us denote byP:L2(ν)ˆ →L2(ν)ˆ the Perron–Frobenius operator such that:

f, gL2(ν)ˆ

ˆ Σ

Pf (x)g(x)dν(x)ˆ =

ˆ Σ

f (x)g◦ ˆθ (x)dν(x).ˆ

Letη(0,1). Let us denote byB the set of boundedη-Hölder continuous function g:Σˆ →Cendowed with the usual Hölder norm:

gB:= g+sup

x=y

|g(y)g(x)| d(x, y)ˆ η .

Références

Documents relatifs

Consider the line on which all the sparse squares sit in the previous layout. We call it the base line. In order to build the squares properly, we set a specific prefix before the

Layer TO This layer is the core of this proof, and it is where 3D actually comes to play: in the thick aperiodic yz planes we will compute the Π10 oracle by encoding an

Our first result states that a strongly aperiodic effectively closed subshift (and in particular a strongly aperiodic SFT) forces the group to have a decidable word problem in the

Under suitable assumptions on the point process and the transition probabilities, we obtain in this paper almost sure recurrence or transience results, namely Theorem 1 and Theorem

ρ-mixing case (Proof of Application 1 of Section 1). 4-5] for additive functionals. Exten- sion to general MRW is straightforward. Then Application 1 of Section 1 follows from

In some cases (conditions C 2a and C 2b below), the comparison with Bessel processes suffices to conclude.. It cannot converge to −∞. We give two concrete examples when this

The proposed ELMM (5) is a natural extension of the LMM (1) in order to model spectral variability. The ELMM is based on the definition of a spectral variability function that

Although the question of the recurrence of linearly edge reinforced random walks (ERRW) on infinite graphs has known important breakthroughs in the recent years (cf. notably [MR09]),