• Aucun résultat trouvé

Zero Krengel entropy does not kill Poisson entropy

N/A
N/A
Protected

Academic year: 2022

Partager "Zero Krengel entropy does not kill Poisson entropy"

Copied!
9
0
0

Texte intégral

(1)

www.imstat.org/aihp 2012, Vol. 48, No. 2, 368–376

DOI:10.1214/10-AIHP393

© Association des Publications de l’Institut Henri Poincaré, 2012

Zero Krengel entropy does not kill Poisson entropy

Élise Janvresse and Thierry de la Rue

Laboratoire de Mathématiques Raphaël Salem, UMR 6085 CNRS, Université de Rouen, Avenue de l’Université, B.P. 12, F76801 Saint-Étienne-du-Rouvray Cedex, France. E-mail:Elise.Janvresse@univ-rouen.fr;Thierry.de-la-Rue@univ-rouen.fr

Received 28 June 2010; revised 30 September 2010; accepted 7 October 2010

Abstract. We prove that the notions of Krengel entropy and Poisson entropy for infinite-measure-preserving transformations do not always coincide: We construct a conservative infinite-measure-preserving transformation with zero Krengel entropy (the induced transformation on a set of measure 1 is the Von Neumann–Kakutani odometer), but whose associated Poisson suspension has positive entropy.

Résumé. Nous prouvons que les notions d’entropie de Krengel et d’entropie de Poisson pour les transformations préservant une mesure infinie ne coïncident pas toujours : nous construisons une transformation conservative préservant une mesure infinie qui a une entropie de Krengel nulle (la transformation induite sur un ensemble de mesure 1 est l’odomètre de Von Neumann–Kakutani), mais dont la suspension de Poisson a une entropie strictement positive.

MSC:37A05; 37A35; 37A40; 28D20

Keywords:Krengel entropy; Poisson suspension; Infinite-measure-preserving transformation;d-distance.¯

1. Introduction

1.1. Entropy for infinite-measure-preserving transformations

There exist several notions of entropy for infinite transformations, which elegantly generalize Kolmogorov’s entropy of a probability-preserving transformation. Krengel [4] comes down to the finite-measure case by considering the entropy of the induced transformation on a set of finite measure: TheKrengel entropyof a conservative measure- preserving transformation(X,B, μ, T )is defined as

hKr(X,B, μ, T ):= sup

A∈F+

μ(A)h(A,BA, μA, TA),

whereF+is the collection of sets inBwith finite positive measure,μAis the normalized probability measure onA obtained by restrictingμtoBA, andTA:AAis the induced map onA. Recall that this map is defined by

TA(x):=TrA(x)(x),

whererA(x):=min{k≥1: Tk(x)A} is thefirst-return-time map associated toA. As soon as T is not purely periodic, Krengel proved that

hKr(X,B, μ, T )=μ(A)h(A,BA, μA, TA),

whereAis any finite-measuresweep-outset (i.e., a set such that

n0TnA=X).

(2)

TheParry entropyof an infinite-measure-preserving transformationT has been defined in [7] as the supremum of the conditional entropy ofCwith respect toT1C, for allσ-finite sub-σ-algebrasCsuch thatT1CC.

Recall now that to each infinite-measure-preserving transformationT we can associate a probability-preserving transformationTcalled itsPoisson suspension, and which can be described as follows (we refer to [8] for details):

We consider a Poisson process on X with intensityμ, which we can consider as a random collection of particles.

These particles are distributed overXin such a way that, denoting byNBthe random number of particles in any finite- measure setB, for any finite collection of pairwise disjoint, finite-measure setsB1, . . . , BnB, the random variables NB1, . . . , NBnare independent, and follow Poisson distributions with respective parametersμ(B1), . . . , μ(Bn). Then T is defined on the canonical space of this Poisson process, and it consists in moving individually each of these particle according to the transformationT onX. ThePoisson entropyof an infinite-measure-preserving transformation was defined by Roy [8] as the Kolmogorov entropy of its Poisson suspension.

Relations between these notions of entropy are studied in [3]: On large classes of transformations (e.g., quasi-finite transformations, rank-one transformations), it is proved that Poisson entropy is equal to Krengel entropy and to Parry entropy. Moreover, in any case, Parry entropy is dominated by both Krengel and Poisson entropy.

It was asked in [3] whether, for any conservative measure-preserving transformation, these three definitions always coincide. The purpose of the present paper is to show that the answer is negative, by constructing a counterexample.

Theorem 1.1. There exists a conservative infinite-measure-preserving transformation with zero Krengel entropy (hence zero Parry entropy),but whose associated Poisson suspension has positive entropy.

2. Construction

2.1. Von Neumann–Kakutani odometer

The transformationT is constructed as a tower over the Von Neumann–Kakutani odometerS. Let us recall the con- struction of the latter by cutting and stacking (see Fig.1). We start with the intervalA:= [0,1]. The first step consists in cuttingAinto two sub-intervalsA1andA\A1of measure 1/2, and stackingA\A1overA1. We get a tower of height 2 which we call Tower 1. After stepn,Ahas been cut into 2nsub-intervals which are stacked to get Towern.

This means that at this step each point ofA(except those lying on the top of the tower) is mapped bySto the point of Alying above it. We construct Towern+1 by cutting Towerninto two equal parts. We callAn+1the left half of the top interval of Towernand we stack the right part of the tower overAn+1, thus dividing by 2 the measure of the set whereSis not yet defined. Repeating this procedure defines the Von Neumann–Kakutani odometer which preserves the Lebesgue measure onA. It is well known that the odometerSis ergodic and has zero entropy.

2.2. Construction ofT

T is constructed onR+ in such a way that the induced transformationTA coincides with the odometer previously defined.T is completely defined (up to isomorphism) by giving for each pointxAthe first return timerA(x)toA.

We fix an increasing sequence of integers(Mn)n0, withMn→ ∞. For any n≥1, we choose a large enough integerkn(to be precised later). We define the first return time toAso that its restriction toAnis uniformly distributed on{Mn, Mn+1, . . . , Mn+kn−1}for anyn.

We will see in Section4that by choosingknlarge enough, the entropy ofTis positive.

Fig. 1. First steps in the construction of the Von Neumann–Kakutani odometer by cutting and stacking.

(3)

3. Poisson approximation lemma

The purpose of this section is to prove the key lemma. This lemma roughly states that whenkn is large enough, it is almost impossible in the Poisson suspension to keep track individually of the particles when they leaveAn if we only have access to the number of particles inA. For this, we compare two processes: The first one is simply an i.i.d.

sequence of Poisson random variables, whereas the second one modelizes particles leavingAnand coming back toA.

The comparison between the two processes uses the notion ofd-distance, of which we recall some properties.¯ 3.1. Thed¯-distance

Thed-distance between two stationary processes has been introduced by Ornstein for the proof of the isomorphism¯ theorem of Bernoulli shifts. We refer to [6] or [9] for the properties of this distance which we use later and which we recall here.

Letξ andζ be two stationary processes taking values in a countable alphabetB. For any integersp < q, we denote byξ|qpthe finite sequencep, ξp+1, . . . , ξq).

ForL≥1, letJL(ξ, ζ )be the set of all joinings ofξ|L1 andζ|L1, that is probability distributions onBL×BLwhose marginals are the distributions ofξ|L1 andζ|L1. We first defined¯Lfor anyL≥1, by

d¯L(ξ, ζ ):= min

λJL(ξ,ζ )Eλ

dL ξ|L1, ζ|L1

,

wheredLis the Hamming distance between sequences of lengthL:

dL(x1· · ·xL, z1· · ·zL):= 1 L

L

1

1xi =zi.

Then, thed¯-distance betweenξ andζ is defined by d(ξ, ζ )¯ :=sup

L1

d¯L(ξ, ζ ).

It can be shown thatd(ξ, ζ )¯ is also the minimum ofλ(ξ0 =ζ0)whenλranges over all stationary joinings ofξ andζ. The two key properties of thed-distance that we shall use are: On the one hand, the fact that entropy of processes¯ close ind¯-distance can be compared (Lemma 3.1). On the other hand, a practical tool to estimate the d-distance¯ between processes using conditional distributions on the past: If, for all large enoughn,

b∈B

P

ξ0=b|ξ|1n

−P

ζ0=b|ζ|1n< ε (1)

for all pastξ|1noutside a set of measureεand all pastζ|1noutside a set of measureε, thend(ξ, ζ )¯ ≤3ε. Moreover, the same conclusion holds if we replace in (1) the conditional distributions with respect to the past by conditional distributions with respect to finerσ-algebras.

The fact that entropy is a continuous function of processes taking values in a given finite alphabet, when these processes are topologized with thed¯-distance, is well known (see, e.g., [9], p. 100). This is however no longer true if we consider processes taking countably many values. In this case, entropy is only a lower semicontinuous function of the process.

Lemma 3.1. Letξ be a stationary process taking values in a countable alphabetBwhich has finite entropy.For any ε >0,there existsδ >0such that any stationary process ζ taking values in the same alphabetBwithd(ξ, ζ ) < δ¯ satisfiesh(ζ ) > h(ξ )ε.

(4)

Proof. Without loss of generality we can assumeB=Z+. For any integer >0 we define the process)taking values in the finite alphabet{0,1, . . . , }by

)x= ξx ifξx< , otherwise.

Theσ-algebra generated by the process)increases to theσ-algebra generated byξ. Hence we can choose large enough so thath(ξ) > h(ξ )ε/2. By continuity of the entropy in the case of a finite alphabet, we can find δ >0 such that any process taking values in{0,1, . . . , }atd¯-distance at mostδ from)has entropy at least h(ξ )ε. Now, ifd(ξ, ζ ) < δ, then¯ d((ξ¯ ∧), (ζ)) < δ(where)is defined fromζ in a similar way), hence

h(ζ )h(ζ)h(ξ )ε.

3.2. Comparison between connected and disconnected processes

LetAbe a finite alphabet,PB andPW be two probability measures onAandλA×A be a joining ofPB andPW. Letδbe some fixed positive real number. We define two processesξ(thedisconnected process) andζ (theconnected process) on(NA)Z, and estimate thed-distance between them.¯

The processξ is constructed from two independent sequences of i.i.d. random variables distributed according to the Poisson distribution of parameterδ/2, which can be interpreted as numbers of black and white particles lying on each site ofZ. Then to each black (resp., white) particle we randomly and independently associate a label picked inA according toPB(resp.,PW). For anyx∈Zand any labelsa, b∈A,ξxB(a)(resp.,ξxW(b)) is the total number of black (resp., white) particles labelled bya(resp.,b) at positionx. In other words, the processξassociates in an i.i.d. way to each sitex∈Za finite sequenceξx=xB(a), ξxW(b)))a,b∈Aof independent random variables respectively distributed according to the Poisson distribution of parameterPB(a)δ/2 andPW(b)δ/2. The processξ is called the disconnected process because its two componentsξBandξWare independent.

LetkandM be two integers. The processζ is also constructed from black and white particles onZ, but which are no longer independent. This is why we call it the connected process. The number of black particles at each site is given by a sequence of i.i.d. random variables distributed according to the Poisson distribution of parameterδ/2.

For each black particle at positionx∈Z, we first pick a random integerj, uniformly in{M, M+1, . . . , M+k−1}

and independently of all other particles. Then we link the black particle to a white particle that we put at position x +j. For each such couple of black and white particles, a couple of labels is picked in A×Awith probability λA×A: The first label is associated to the black particle and the second label to the white one. Then, for anyx∈Z, ζx=xB(a), ζxW(b)))a,b∈A denote the number of black particle labelled by a and the number of white particles labelled bybat positionx.

Lemma 3.2. For anyε >0,ifkis large enough,d(ξ, ζ ) < ε.¯

Proof (Simple case). We first prove the lemma in the case whereAis reduced to a singleton. Since all particles have the same label, we just forget it and simply count the number of black and white particles on each site.

We now prove that ifkis large enough, for anynM+k,

j,∈N

P

ζ0B=j, ζ0W=|ζ|1n

−eδ/2(δ/2)j j!

eδ/2(δ/2) !

< ε,

with probability 1−εonζ|1n. Sinceζ0Bis Poisson distributed with parameterδ/2, and independent fromζ|1n, and sinceζ0W is independent fromζ0B conditionally to the past, it is enough to prove that ifk is large enough, for any nM+k,

∈N

P

ζ0W=|ζ|1n

−eδ/2(δ/2) !

< ε, with probability 1−εonζ|1n.

(5)

Fig. 2. The enriched past.

In fact we rather condition with respect to an enriched past (see Fig.2): Formally, theenriched pastis theσ-algebra generated by the number of black and white particles on each sitex∈ {−n, . . . ,−1}, and the knowledge of all links existing between a black particle and a white particle both lying in{−n, . . . ,−1}.

Conditionally to the enriched past, we can distinguish two kinds of black particles between−n and−1: Those which are linked to a white particle lying on the left of 0, and those, calledfree particles, whose white particle’s position is unknown. Observe that only free particles may have some influence onζ0W. Hence, black particles lying on the left of site−(M+k−1)have no influence onζ0W since they are not free. Black particles lying on the right of site

Mare free but nevertheless have no influence onζ0Wsince a black particle is linked to a white particle at distance at leastM.

So it remains to study the influence onζ0W of free black particles at sites between−(M+k−1)and−M. For 1≤jklet us denote byFjthe number of free particles at site−(M+k)+j. Any black particle at site−(M+k)+j has probabilityj/ kto be free. Therefore,Fj follows the Poisson distribution with parameter δj2k.

Fix 1≤jk. Assume there is a free particle at site(M+k)+j. Since there arej possible positions for its white particle, the latter has probability 1/j to lie at site 0. Hence, conditionally to our enriched past, the number of white particles at site 0 can be written as

k

j=1 Fj

=1

Bj,

where(Bj)are independent Bernoulli random variables with respective parameter 1/j. The law of such a sum of independent Bernoulli variables is close to a Poisson distribution of parameterδ/2 as soon as the sum of the parameters is close toδ/2 and all parameters are small enough (see [2], Theorem 23.2, p. 312). Therefore we can choose a large enough integerJ, andε1small enough so that, ifZ is a sum of independent Bernoulli random variables, each with parameter less than 1/Jand such that the sum of parameters is withinε1ofδ/2, then

0

P(Z=)−exp(−δ/2)(δ/2) !

< ε.

We have now to avoid bad configurations, that is configurations of the enriched past which have free particles close to−(M+k−1)giving rise to Bernoulli with large parameters, and configurations such that the sum of the parameters is not close enough toδ/2.

(6)

Control of the parameters’ size.We compute the probability that no free particle lie between−(M+k)+1 and

(M+k)+J:

P(Fj=0, j=1, . . . , J )= J

j=1

exp

δj 2k

=exp

δJ (J+1) 4k

.

Under this condition, ζ0W is (conditionally to the enriched past) the sum of independent Bernoulli variables with parameters smaller than 1/J. Ifkis large enough, this happens with probability larger than 1−ε/2.

Control of the parameters’ sum. Since the Fj are independent and Poisson distributed with parameter δj2k, the expected value of the sumS:=k

j=1j1Fjof the parameters isδ/2, and its variance is varS=

k

j=1

1 j2

δj 2k = δ

2k k

j=1

1 j. Hence, ifkis large enough,

P

|Sδ/2|< ε1

>1−ε/2.

Putting things together, we have proved that with probability larger than 1−εon the enriched past, the conditional distribution of the numberζ0W of white particles at site 0 satisfies

0

P

ζ0W=|enriched past

−exp(−δ/2)(δ/2) !

< ε.

This proves Lemma3.2whenAis reduced to a singleton.

Proof (General case). We consider the family of independent processesζa,b=B,a,b, ζW,a,b),(a, b)∈A×A, which counts the number of black and whiteζ particles atx belonging to a pair of black and white particles respec- tively labelled byaandb. Thenζa,bis a simple-caseζ process, for which the expected number of black particles per site isδλA×A(a, b)/2. From the proof in the simple case, we know that as soon askis large enough, thed¯-distance betweenζa,b andξa,b is smaller thanε/|A|2, whereξa,b=B,a,b, ξW,a,b)is composed of two i.i.d. sequences of Poisson random variables of parameterδλA×A(a, b)/2. We can recoverζxB(a)andζxW(b)by

ζxB(a)=

b∈A

ζxB,a,b and ζxW(b)=

a∈A

ζxW,a,b.

On the other hand,

b∈AξB,a,b(resp.,

a∈AξW,a,b) has the same distribution asξB(a)(resp.,ξW(b)): It is an i.i.d.

sequence of Poisson random variables of parameterPB(a)δ/2 (resp.,PW(b)δ/2). Summing overa andb, it follows

thatd(ζ, ξ ) < ε.¯

4. Positive Poisson entropy

We denote byξ()the stationary process living in the Poisson suspension of our transformationT, defined by ξx():=number of particles inAat timex.

The purpose of this section is to show that the entropy of the processξ()is positive as soon as thekn’s are chosen large enough. This will be proved by showing that thed¯-distance betweenξ()and an i.i.d. sequenceξ(0)of random Poisson variables with parameter 1 can be made as small as we want. By Lemma3.1, this will be enough to conclude.

(7)

Fig. 3. First steps in the construction of the return time toA.

Our strategy is the following: As one goes along in the construction of the return time toA, we define a sequence (T(n))of infinite-measure-preserving transformations acting onR+. Then we consider the processξ(n)living in the Poisson suspension ofT(n):

ξx(n):=number of particles inAat timexfor the Poisson suspension overT(n).

The transformationsT(n)will be constructed so that the processξ(n)converge ind¯-distance toξ.

The transformationT(0)is constructed by stacking infinitely many pairwise disjoints intervals of length 1,Abeing one of them, into a doubly infinite tower, each interval being mapped onto the one just above. Therefore, as mentioned previously,ξ(0)is an i.i.d. sequence of Poisson variables with parameter 1. At stepnof the construction, we define the return time toAon the subsetAn. This return time toAwill be the same for all transformationsT(m),mn, and for the final transformationT. By choosing the return time adequately, we will make sure that

d¯

ξ(n1), ξ(n)

<2nε, so that for alln

d¯

ξ(0), ξ(n)

< ε. (2)

Let us describe the first step. Recall that Tower 1 is of height 2, with basisA1(see Fig.1). We cutA1intok1equal subintervals, and we define onA1the return time toAto beM1+j−1 on thejth subinterval. We insert Tower 1 into a doubly infinite tower of intervals of length 1/2 and add spacers betweenA1and its image by the odometerS: We insertM1+j−2 spacers of width 1/(2k1)between thejth subinterval ofA1and its image byS. The transformation T(1)is defined by mapping each point to the point right above it (see Fig.3).

The processξ(1)counts the number of particles inAat timex for the Poisson suspension overT(1): We interpret points inA1as black particles and points inA\A1as white particles. For the suspension overT(0), black and white particles are independent, whereas for the suspension overT(1), they are linked through the return time toAof the point corresponding to the black particle. Hence, a direct application of Lemma3.2in the simple case with no alphabet gives thatd(ξ¯ (0), ξ(1)) < ε/2 whenk1is large enough.

Suppose the return time toAhas already been defined on allAi, 1≤in−1. Consider Towern. The return time to Ahas already been defined on all rungs but the roof and An (which is the rung of level 2n1). LetRn1

be the maximum value of the already defined return time toAonA1∪ · · · ∪An1. We consider the finite alphabet A:= {1, . . . , Rn1}2n11. For each pointy inAn, lethB(y)∈Abe the sequence of the return times toAwhen we climb the first half of Towernbefore reachingy, in other wordshB(y)is the sequence of the last 2n1−1 return

(8)

times toA. LethW(y)∈Abe the sequence of the return times toAwhen we climb the second half of Towernstarting fromSy. We denote byr(y)the return time toA, which is to be defined at this step. We wantr(y)to be independent ofhB(y)andhW(y). To this end, we consider the finite partition ofAngenerated byhB andhW. Each atom of this partition is cut intokn equal pieces, and we definerto beMn+j −1 on thejth piece of each atom. Here is how we define the transformation T(n): we insert Tower ninto a doubly infinite tower of intervals of length 1/2n and insert as many spacers as we need between the rungs of Towernto achieve the already defined return time toA. The transformationT(n)maps each point to the point right above it.

Let us turn to the estimation ofd(ξ¯ (n1), ξ(n))forn≥2. We want to apply Lemma3.2: Black particles are points inAn and white particles are points inSAn. To each pointy inAn, we associate the labelhB(y)∈A, and to each pointSyinSAn, we attach the labelhW(y)∈A. Let the processξ (resp.,ζ) count the numbers of black and white particles together with their label in the suspension overT(n1)(resp.,T(n)). These processes are exactly of the form studied in Lemma3.2.

Now, observe that we can recoverξ(n1)andξ(n)fromξ andζ: ξ0(n1)=

h∈A

ξ0B(h)+ξhB

1(h)+ · · · +ξhB

1+···+h2n−11(h)+

h∈A

ξ0W(h)+ξWh

1(h)+ · · · +ξW(h

1+···+h2n11)(h) and

ξ0(n)=

h∈A

ζ0B(h)+ζhB

1(h)+ · · · +ζhB

1+···+h2n11(h)+

h∈A

ζ0W(h)+ζWh

1(h)+ · · · +ζW(h

1+···+h2n11)(h).

It follows that, if ξ andζ coincide on {−Rn1(2n1−1), . . . , Rn1(2n1−1)}, then ξ0(n1)=ξ0(n). Let λ be a stationary joining ofξ andζ achieving thed-distance: We have¯ d(ξ, ζ )¯ =λ(ξ0 =ζ0). Then

d¯

ξ(n1), ξ(n)

λ

ξ0(n1) =ξ0(n)

≤2nRn1λ(ξ0 =ζ0)=2nRn1d(ξ, ζ ).¯

By Lemma 3.2, d(ξ, ζ )¯ can be made arbitrarily small by choosing kn large enough. Hence we can assure that d(ξ¯ (n1), ξ(n)) <2nε.

Finally, note that sinceMnis increasing, the return time toAon the roof of Towern(the union ofAi,i > n) will be larger thanMn+1. Hence, for anyL >0, ifnis large enough so thatMn+1> L, the distribution ofξ(n)|L01coincides with the distribution ofξ()|L01. Therefore,

d¯L

ξ(), ξ(0)

= ¯dL

ξ(n), ξ(0)

≤ ¯d

ξ(n), ξ(0)

< ε, which implies

d¯

ξ(), ξ(0)

ε.

5. Comments and open questions

In view of previously known results on the subject, some comments on the infinite-measure-preserving transformation T constructed in Section2may be made.

First, although its construction is derived from the standard cutting-and-stacking procedure used to build the most elementary rank-one system (the Von Neumann–Kakutani odometer), the transformationT is not even of finite rank.

Indeed, Proposition 10.1 in [3] shows that for finite-rank systems, both the Poisson and Krengel entropy vanish.

Second, it was also proved in [3] that Krengel and Poisson entropies coincide forquasi-finite transformations, namely transformations for which there exists a sweep-out setAof measure 1 such that the return-time partition of Ahas finite entropy. There exist only few examples of transformation for which the non quasi-finiteness has been established: an unpublished example constructed by Ornstein has been mentioned by Krengel in [5], and the only published example which we are aware of is a rank-one system, given by Aaronson and Park in [1]. Our construction thus provides a new example of a non-quasi-finite transformation.

(9)

After having proved that the different notions of entropy for infinite-measure-preserving transformations do not always coincide, a natural question is to ask whether they are always ordered in the same way: Is it true that Poisson entropy always dominates Krengel entropy? Can we at least decide whether zero Poisson entropy implies zero Krengel entropy? And what about similar questions regarding the comparison between Parry entropy and Poisson entropy? It may be worth recalling here that the equality of Parry entropy and Krengel entropy in the quasi-finite case was proved by Parry in 1969 [7], but that the question whether they always coincide is, as far as we know, still open.

Acknowledgements

The construction of the transformationT in Section2has been inspired by a private communication of Benjamin Weiss concerning the Shannon–McMillan–Breiman theorem.

We are also much indebted to Emmanuel Roy for stimulating conversations on the subject.

References

[1] J. Aaronson and K. K. Park. Predictability, entropy and information of infinite transformations.Fund. Math.206(2009) 1–21.MR2576257 [2] P. Billingsley.Probability and Measure, 2nd edition. Wiley, New York, 1986.MR0830424

[3] É. Janvresse, T. Meyerovitch, E. Roy and T. de la Rue. Poisson suspensions and entropy for infinite transformations.Trans. Amer. Math. Soc.

362(2010) 3069–3094.MR2592946

[4] U. Krengel. Entropy of conservative transformations.Z. Wahrsch. Verw. Gebiete7(1967) 161–181.MR0218522

[5] U. Krengel. On certain analogous difficulties in the investigation of flows in a probability space and of transformations in an infinite measure space. InFunctional Analysis (Proc. Sympos., Monterey, CA, 1969)75–91. Academic Press, New York, 1969.MR0265558

[6] D. S. Ornstein.Ergodic Theory, Randomness, and Dynamical Systems. Yale Univ. Press, New Haven, CT, 1974.MR0447525 [7] W. Parry.Entropy and Generators in Ergodic Theory. W. A. Benjamin, New York, 1969.MR0262464

[8] E. Roy. Mesures de poisson, infinie divisibilité et propriétés ergodiques. Ph.D. thesis, 2005.

[9] P. C. Shields.The Ergodic Theory of Discrete Sample Paths. Graduate Studies in Mathematics13. Amer. Math. Soc. Providence, RI, 1996.

MR1400225

Références

Documents relatifs

Le tableaumontre que l’indice de réfraction ainsi que la densité massique des couches minces d’oxyde de titane dopées auZrO 2 ou auZnO augmente en fonction de l’augmentation de

We prove that the notions of Krengel entropy and Poisson entropy for infinite-measure-preserving transformations do not always coincide: We construct a

It will be pointed out here that this so-called (F, G)-geometry (which includes F -geometry as a special case) is identical to Zhang’s (2004) extension to the α-geometry, where the

Thé extent to which parents become involved in school activities is very much a function of thé level of teacher satisfaction and of thé school climate, of thé

In this part, we aim at proving the existence of a global in time weak solution for the lubrication model by passing to the limit in the viscous shallow water model with two

This SPR property implies that the geodesic flow admits a finite Bowen-Margulis measure, which is then the unique measure of maximal entropy, according to [OP04], see Theorem 3.15 ,

We obtain above estimates for these invariants for actions by diffeomorphisms of a compact manifold (with a Borel invariant measure) and, more generally, by

When the class consists of minimal zero entropy systems (denoted by M 0 ), we prove that if (X, T ) ⊥ M 0 , then (X, T ) has an invariant measure with full support, and if in