• Aucun résultat trouvé

Limit laws of transient excited random walks on integers

N/A
N/A
Protected

Academic year: 2022

Partager "Limit laws of transient excited random walks on integers"

Copied!
26
0
0

Texte intégral

(1)

www.imstat.org/aihp 2011, Vol. 47, No. 2, 575–600

DOI:10.1214/10-AIHP376

© Association des Publications de l’Institut Henri Poincaré, 2011

Limit laws of transient excited random walks on integers

Elena Kosygina

a,1

and Thomas Mountford

b,2

aDepartment of Mathematics, Baruch College, Box B6-230, One Bernard Baruch Way, New York, NY 10010, USA.

E-mail:elena.kosygina@baruch.cuny.edu

bEcole Polytechnique Fédérale, de Lausanne, Département de mathématiques, 1015 Lausanne, Switzerland. E-mail:thomas.mountford@epfl.ch Received 30 August 2009; revised 2 May 2010; accepted 10 May 2010

Abstract. We consider excited random walks (ERWs) onZwith a bounded number of i.i.d. cookies per site without the non- negativity assumption on the drifts induced by the cookies. Kosygina and Zerner [15] have shown that when the total expected drift per site,δ, is larger than 1 then ERW is transient to the right and, moreover, forδ >4 under the averaged measure it obeys the Central Limit Theorem. We show that whenδ(2,4]the limiting behavior of an appropriately centered and scaled excited random walk under the averaged measure is described by a strictly stable law with parameterδ/2. Our method also extends the results obtained by Basdevant and Singh [2] forδ(1,2]under the non-negativity assumption to the setting which allows both positive and negative cookies.

Résumé. On considère des marches aléatoires excitées surZavec un nombre borné de cookies i.i.d. à chaque site, ceci sans l’hypothèse de positivité. Auparavent, Kosygina et Zerner [15] ont établi que si la dérive totale moyenne par site,δ, est strictement superieur à 1, alors la marche est transiente (vers la droite) et, de plus, pourδ >4 il y a un théorème central limite pour la position de la marche. Ici, on démontre que pourδ(2,4]cette position, convenablement centrée et réduite, converge vers une loi stable de paramètreδ/2. L’approche permet également d’étendre les résultats de Basdevant et Singh [2] pourδ(1,2]à notre cadre plus général.

MSC:Primary: 60K37; 60F05; 60J80; secondary: 60J60

Keywords:Excited random walk; Limit theorem; Stable law; Branching process; Diffusion approximation

1. Introduction and main results

Excited random walk (ERW) on Zd was introduced by Benjamini and Wilson in [3]. They proposed to modify the nearest neighbor simple symmetric random walk by giving it a positive drift (“excitation”) in the first coordinate direction upon reaching a previously unvisited site. If the site had been visited before, then the walk made unbiased jumps to one of its nearest neighbor sites. See [4,13,17] and references therein for further results about this particular model.

Zerner [19,20] generalized excited random walks by allowing to modify the transition probabilities at each site not just once but any number of times and, moreover, choosing them according to some probability distribution. He obtained the criteria for recurrence and transience and the law of large numbers for i.i.d. environments onZdand strips and also for general stationary ergodic environments onZ. It turned out that this generalized model had interesting behavior even ford=1, and this case was further studied in [1,2,18].

1Supported in part by NSF Grants DMS-08-25081 and DMS-06-35607.

2Supported in part by the C.N.R.S. and by the SNSF Grant 200020-115964.

(2)

Results obtained in all these works rely on the assumption that projections of all possible drifts on some fixed direction are non-negative. In fact, the branching processes framework introduced in [14] for random walks in random environment (d=1) and employed in [1,2] for excited random walks, does not depend on the positivity assumption, and it seems natural to use this approach for extending the analysis to environments which allow both positive and negative drifts. This was done in [15], where the authors discussed recurrence and transience, laws of large numbers, positive speed, and the averaged central limit theorem for multi-excited random walks onZin i.i.d. environments with bounded number of “excitations” per site. We postpone further discussion of known results ford=1 and turn to a precise description of the model considered in this paper.

Given an arbitrary positive integerMlet

M:=

ωz(i)

i∈N

z∈Z|ωz(i)∈ [0,1], fori∈ {1,2, . . . , M}andωz(i)=1/2,fori > M, z∈Z .

An element ofM is called a cookie environment. For eachz∈Z, the sequence{ωz(i)}i∈N can be thought of as a pile of cookies at sitez, andωz(i)is referred to as “theith cookie atz”. The numberωz(i)is equal to the transition probability fromztoz+1 of a nearest-neighbor random walk upon theith visit toz. Ifωz(i) >1/2 (resp.,ωz(i) <

1/2) the corresponding cookie will be called positive (resp. negative), ωz(i)=1/2 will correspond to a “placebo”

cookie or, equivalently, the absence of an effectiveith cookie at sitez.

LetPbe a probability measure onM, which satisfies the following two conditions:

(A1) Independence: the sequencez(·))z∈Zis i.i.d. underP.

(A2) Non-degeneracy:

E M

i=1

ω0(i)

>0 and E M

i=1

1−ω0(i)

>0.

Notice that we do not make any independence assumptions on the cookies at the same site.

It will be convenient to define our ERW model using a coin-toss construction. Let(,F)be some measurable space equipped with a family of probability measuresPx,ω, x∈Z, ωM, such that for each choice ofx∈Zand ωM we have±1-valued random variablesBi(z), z∈Z, i≥1,which are independent underPx,ωwith distribution given by

Px,ω

Bi(z)=1

=ωz(i) and Px,ω

Bi(z)= −1

=1−ωz(i). (1.1)

LetX0 be a random variable on(,F, Px,ω)such thatPx,ω(X0=x)=1. Then an ERW starting at x∈Zin the environmentω,X:= {Xn}n0, can be defined on the probability space(,F, Px,ω)by the relation

Xn+1:=Xn+B#(X{rn)n|X

r=Xn}, n≥0. (1.2)

Informally speaking, upon each visit to a site the walker eats a cookie and makes one step to the right or to the left with probabilities prescribed by this cookie. Sinceωz(i)=1/2 for alli > M, the walker will make unbiased steps fromzstarting from the(M+1)th visit toz.

Events{Bi(z)=1},i∈N,z∈Z, will be referred to as “successes” and events{Bi(z)= −1}will be called “failures”.

The consumption of a cookieωz(i)induces a drift of size 2ωz(i)−1 with respect toPx,ω. Summing up over all cookies at one site and taking the expectation with respect toPgives the parameter

δ:=E

i1

0(i)−1

=E M

i=1

0(i)−1

, (1.3)

which we call theaverage total drift per site. It plays a key role in the classification of the asymptotic behavior of the walk.

(3)

We notice that there is an obvious symmetry between positive and negative cookies: if the environmentz)z∈Zis replaced byz)z∈Zwhereωz(i)=1−ωz(i), for alli∈N, z∈Z, thenX:= {Xn}n0, the ERW corresponding to the new environment, satisfies

X= −D X, (1.4)

where=Ddenotes the equality in distribution. Thus, it is sufficient to consider, say, only non-negativeδ(this, of course, allows both negative and positive cookies), and we shall always assume this to be the case.

Define the averagedmeasurePx by settingPx(·)=E(Px,ω(·)). Below we summarize known results about this model.

Theorem 1.1 ([15]). Assume(A1)and(A2).

(i) Ifδ∈ [0,1]thenXis recurrent,i.e.forP-a.a.ωit returnsP0,ω-a.s.infinitely many times to its starting point.If δ >1thenXis transient to the right,i.e.forP-a.a.ω,Xn→ ∞asn→ ∞P0,ω-a.s.

(ii) There is a deterministicv∈ [0,1]such thatXsatisfies forP-a.a.ωthe strong law of large numbers,

nlim→∞

Xn

n =v, P0,ω-a.s. (1.5)

Moreover,v=0forδ∈ [0,2]andv >0forδ >2.

(iii) Ifδ >4then the sequence Btn:=X[t n]− [t n]v

n , t≥0,

converges weakly underP0to a non-degenerate Brownian motion with respect to the Skorohod topology on the space of cádlág functions.

This theorem does not discuss the rate of growth of the ERW when it is transient but has zero linear speed (1< δ≤ 2). It also leaves open the question about fluctuations whenδ≤4.

The rate of growth of the transient cookie walk with zero linear speed was studied in [2] for the case of determin- istic spatially homogeneous non-negative cookie environments. For further discussion we need some notation for the limiting stable distributions that appear below. Givenα(0,2]andb >0, denote byZα,ba random variable (on some probability space) whose characteristic function is determined by the relation

logEeiuZα,b=

b|u|α

1−i|uu|tanπα

2

, ifα =1,

b|u|

1+2iπ|uu|log|u|

, ifα=1. (1.6)

Observe thatZ2,b is a centered normal random variable with variance 2b. The weak convergence with respect toP0 will be denoted by⇒.

Theorem 1.2 ([2]). Letωz(i)=pi ∈ [1/2,1),i∈Nfor allz∈Z,wherepi=1/2fori > M,andδ be as in(1.3), that isδ=M

i=1(2pi−1).

(i) Ifδ(1,2)then there is a positive constantbsuch that asn→ ∞ Xn

nδ/2(Zδ/2,b)δ/2.

(ii) Ifδ=2then(Xnlogn)/nconverges in probability to some constantc >0.

The above results also hold ifXnis replaced bysupinXi orinfinXi.

The proof of Theorem1.2used the non-negativity of cookies, though this assumption does not seem to be essential for most parts of the proof. It is certainly possible that the approach presented in [2] could yield the same results without the non-negativity assumption.

(4)

The functional central limit theorem for ERWs withδ∈ [0,1)in stationary ergodic non-negative cookie envi- ronments was obtained in [7]. The limiting process is shown to be Brownian motion perturbed at extrema (see, for example, [5,6]).

The main results of this paper deal with the case whenδ(2,4], though they apply also toδ(1,2]. Moreover, our approach provides an alternative proof of Theorem1.2for general cookie environments that satisfy conditions (A1) and (A2) (see Remark9.2).

We establish the following theorem.

Theorem 1.3. LetTn=inf{j≥0|Xj=n}andvbe the speed of the ERW(see(1.5)).The following statements hold under the averaged measureP0.

(i) Ifδ(2,4)then there is a constantb >0such that asn→ ∞ Tnv1n

n2/δZδ/2,b and (1.7)

Xnvn

n2/δ ⇒ −v1+2/δZδ/2,b. (1.8)

(ii) Ifδ=4then there is a constantb >0such that asn→ ∞ Tnv1n

nlognZ2,b and (1.9)

Xnvn

nlogn⇒ −v3/2Z2,b. (1.10)

Moreover, (1.8)and(1.10)hold ifXnis replaced bysupinXi orinfinXi.

The paper is organized as follows. In Section2 we recall the branching processes framework and formulate two statements (Theorems2.1and2.2), from which we later infer Theorem1.3. Section3explains the idea of the proof of Theorem2.2and studies properties of the approximating diffusion process. In Section4we determine sufficient conditions for the validity of Theorem2.2. Section5contains the main technical lemma (Lemma5.3). It is followed by three sections, where we use the results of Section5 to verify the sufficient conditions of Section4 and prove Theorem2.1. The proof of Theorem1.3is given in Section9. TheAppendixcontains proofs of several technical results.

2. Reduction to branching processes

Suppose that the random walk{Xn}n0starts at 0. Sinceδ≥0, Lemma 5 of [15] implies thatP0(Tn<)=1 for all n∈N. At first, we recall the framework used in [1,2,15]. The main ideas go back at least to [16] and [14].

Forn∈Nandkndefine Dn,k=

Tn1

j=0

1{Xj=k,Xj+1=k1},

the number of jumps fromktok−1 before timeTn. Then Tn=n+2

kn

Dn,k=n+2

0kn

Dn,k+2

k<0

Dn,k. (2.1)

The last sum is bounded above by the total time spent byXnbelow 0. Whenδ >1, i.e.Xnis transient to the right, the time spent below 0 isP0-a.s. finite, and, therefore, for anyα >0

nlim→∞

k<0Dn,k

nα =0, P0-a.s. (2.2)

(5)

This will allow us to conclude that for transient ERWs fluctuations ofTnare determined by those of

0knDn,k, once we have shown that the latter are of ordern2/δ.

We now consider the “reversed” process(Dn,n, Dn,n1, . . . , Dn,0). Obviously,Dn,n=0 for everyn∈N. Moreover, givenDn,n, Dn,n1, . . . , Dn,k+1, we can write

Dn,k=

Dn,k+1+1

j=1

# of jumps fromktok−1 between the(j−1)th

andjth jump fromktok+1 before timeTn

, k=0,1, . . . , n−1.

Here we used the observation that the number of jumps fromktok+1 before timeTnis equal toDn,k+1+1 for all kn−1. The expression “between the 0th and the 1st jump” above should be understood as “prior to the 1st jump”.

Fix anωM and denote byFm(k)the number of “failures” in the sequenceB(k)(see (1.1) withzreplaced byk) before themth “success”. Then, givenDn,k+1,

Dn,k=FD(k)

n,k+1+1.

Since the sequences B(k), k∈Z, are i.i.d. underP0, we have that Fm(k)=DFm(nk1) and can conclude that the distribution of (Dn,n, Dn,n1, . . . , Dn,0)coincides with that of(V0, V1, . . . , Vn), where V = {Vk}k0 is a Markov chain defined by

V0=0, Vk+1=FV(k)

k+1, k≥0.

Forx≥0 we shall denote by[x]the integer part ofxand byPxV the measure associated to the processV, which starts with[x]individuals in the 0th generation. Observe thatV is a branching process with the following properties:

(i) V has exactly 1 immigrant in each generation (the immigration occurs before the reproduction) and, therefore, does not get absorbed at 0.

(ii) The number of offspring of themth individual in generation k is given by the number of failures between the(m−1)th andmth success in the sequenceB(k). In particular, ifVkMthen the offspring distribution of each individual after theMth one is Geom(1/2)(i.e., geometric on{0} ∪Nwith parameter 1/2).

Therefore (here and throughout taking any sum fromktofork > to be zero) we can write

Vk+1=

M(Vk+1)

m=1

ζm(k)+

VkM+1

m=1

ξm(k), k≥0, (2.3)

where{ξm(k);k≥0, m≥1}are i.i.d. Geom(1/2)random variables, vectors1(k), ζ2(k), . . . , ζM(k)),k≥0, are i.i.d. under PxV and independent of{ξm(k);k≥0, m≥1}. For eachk≥0 the random variables{ζm(k)}Mm=1are neither independent nor identically distributed, but, given that for somej < M

j

m=1

ζm(k)M,

that is all cookies at sitekhave been eaten before thejth jump fromkto(k+1), we are left with{ζm(k)}Mm=j+1that are independent Geom(1/2) random variables. Define

σ0V =inf{j >0|Vj=0}, SV=

σ0V1

j=0

Vj. (2.4)

(6)

Detailed information about the tails ofσ0V andSV will enable us to use the renewal structure and characterize the behavior of

0knDn,k, and, therefore, ofTn asn→ ∞for transient ERWs. We shall show in Section9that the following two statements imply Theorem1.3.

Theorem 2.1. Letδ >0.Then

nlim→∞nδP0V

σ0V> n

=C1(0,). (2.5)

Theorem 2.2. Letδ >0.Then

nlim→∞nδ/2P0V

SV > n

=C2(0,). (2.6)

Remark 2.3. In fact, a weaker result than (2.5) is sufficient for our purpose: there is a constant B such that nδP0V0V > n)B for alln∈N(see condition(A)in Lemma4.1).We also would like to point out that the lim- its in(2.5)and(2.6)exist for every starting pointx∈N∪ {0}withC1andC2depending onx.The proofs simply repeat those forx=0.

For the model described in Theorem1.2,the convergence(2.5)starting fromx∈Nis shown in[2],Proposition3.1, (forδ(1,2)),and(2.6)forδ(1,2]is the content of[2],Proposition4.1.Theorem2.1can also be derived from the construction in[15] (see Lemma17)and[12].We use a different approach and obtain both results directly without using the Laplace transform and Tauberian theorems.

We close this section by introducing some additional notation. Forx≥0 we set

τxV =inf{j >0|Vjx}, (2.7)

σxV=inf{j >0|Vjx}. (2.8)

We shall drop the superscript whenever there is no possibility of confusion.

When the random walkXis transient to the right,PyV0V<)=1 for everyy≥0. This implies thatPyVxV <

)=1 for everyx∈ [0, y).

Let us remark that when we later deal with a continuous process on[0,∞)we shall simply use the first hitting time ofxto record the entrance time in[x,)(or[0, x]), given that the process starts outside of the mentioned interval.

We hope that denoting the hitting time ofx for such processes also byτxwill not result in ambiguity.

3. The approximating diffusion process and its properties

The bottom-line of our approach is that the main features of branching processV killed upon reaching 0 are reasonably well described by a simple diffusion process.

The parameters of such diffusion processes can be easily computed at the heuristic level. ForVkM, (2.3) implies that

Vk+1Vk= M

m=1

ζm(k)M+1+

VkM+1

m=1

ξm(k)−1

. (3.1)

By conditioning on the number of successes in the firstM tosses it is easy to compute (see Lemma 3.3 in [1] or Lemma 17 in [15] for details) that for allx≥0

ExV M

m=1

ζm(k)M+1

=1−δ. (3.2)

The termM

m=1ζm(k)M+1 is independent ofVkM+1

m=1 m(k)−1). WhenVk is large, the latter is approximately normal with mean 0 and variance essentially equal to 2Vk.

(7)

Therefore, the relevant diffusion should be given by the following stochastic differential equation:

dYt=(1δ)dt+

2YtdBt, Y0=y >0, t∈ 0, τ0Y

, (3.3)

where forx≥0 we set

τxY=inf{t≥0|Yt =x}. (3.4)

Throughout the rest of the paper, unless stated otherwise, we shall assume that δ >0. Observe that τ0Y <∞ a.s., since 2Yt is a squared Bessel process of dimension 2(1−δ) <2 (for a proof, seta=0 and letb→ ∞in part (ii) of Lemma3.2).

The above heuristics are justified by the next lemma.

Lemma 3.1. LetY = {Yt}t0be the solution of (3.3).Fix an arbitraryε >0.Fory(ε,)letV0= [ny],and define Ytε,n=V[nt]∧σV

εn

n , t∈ [0,∞),

whereσxV is given by(2.8).Then the sequence of processesYε,n= {Ytε,n}t0converges in distribution asn→ ∞with respect to the Skorohod topology on the space of càdlàg functions to the stopped diffusionYε= {YtτY

ε}t0,Y0=y. Proof. We simply apply the (much more general) results of [10]. We first note that our convergence result considers the processes up to the first entry into(−∞, ε]forε >0 fixed. So we can choose to modify the rules of evolution for V whenVkεn: we consider the process(Vkn,ε)k0where, with the existing notation,

V0n,ε= [ny], Vkn,ε+1= M

m=1

ζm(k)+

Vkn,ε(εn)M+1

m=1

ξm(k), k≥0. (3.5)

Then (given the regularity of points for the limit process) it will suffice to show the convergence of processes Ytε,n=V[n,εnt]

n , t∈ [0,∞),

to the solution of the stochastic integral equation dYt=(1δ)dt+

2(Ytε)dBt, Y0=y >0, t∈ [0,∞). (3.6)

We can now apply Theorem 4.1 of Chapter 7 of [10] withXn(t)=Ytε,n. The needed uniqueness of the martingale problem corresponding to operator

Gf =(xε)f+(1δ)f (3.7)

follows from [10], Chapter 5, Section 3 (Theorems 3.6 and 3.7 imply the distributional uniqueness for solutions of the corresponding stochastic integral equation, and Proposition 3.1 shows that this implies the uniqueness for the

martingale problem).

We shall see in a moment that this diffusion has the desired behavior of the extinction time and of the total area under the path before the extinction (see Lemmas3.3and3.5). Unfortunately, these properties in conjunction with Lemma3.1do not automatically imply Theorems2.1and2.2, and work needs to be done to “transfer” these results to the corresponding quantities of the processV. Nevertheless, Lemma3.1is very helpful whenV stays large as we shall see later.

In the rest of this section we state and prove several facts aboutY. When we need to specify that the processY starts atyat time 0 we shall writeYy. Again, whenever there is no ambiguity about which process is being considered we shall drop the superscript inτxY defined in (3.4).

(8)

Lemma 3.2. Fixy >0.

(i) (Scaling)LetY = {Yt}t0,whereYt=Yytyy.ThenY=DY1. (ii) (Hitting probabilities)Let0≤a < y < b.Then

PyYa< τb)=bδyδ bδaδ.

Proof. Part (i) can be easily checked by Itô’s formula applied toYt or seen from scaling properties of the generator.

The proof of part (ii) is standard once we notice that the process(Yty)δ stopped upon reaching the boundary of[a, b]

is a martingale. We omit the details.

Lemma 3.3. LetY be the diffusion process defined by(3.3).Then

xlim→∞xδP1Y0> x)=C3(0,).

Proof. For everyε >0 and for allx >1/εwe have by Lemma3.2 xδP1Y0> x)xδP1Y0> x|τεx< τ0)P1Yεx< τ0)

xδPεxY0> x)(εx)δ=εδP1Y

τ0> ε1

>0.

This implies that for eachε >0 lim inf

x→∞ xδP1Y0> x)εδP1Y

τ0> ε1

>0.

Taking the lim supε0in the right-hand side we get lim inf

x→∞ xδP1Y0> x)≥lim sup

ε0

εδP1Y

τ0> ε1

=lim sup

x→∞ xδP1Y0> x).

This would immediately imply the existence of a finite non-zero limit if we could show that lim sup

x→∞ xδP1Y0> x) <. This is the content of the next lemma.

Lemma 3.4. LetY be the diffusion process defined by(3.3).Then lim sup

x→∞ xδP1Y0> x) <.

The proof is very similar to the proof of the discrete version (see (A) in Lemma4.1and its proof in Section6) and,

thus, is omitted.

The final result of this section can be viewed as the “continuous counterpart” of Theorem2.2. It concerns the area under the path ofY.

Lemma 3.5. LetY be the diffusion process defined by(3.3).Then

ylim→∞yδP1Y τ0

0

Ytdt > y2

=C4(0,).

(9)

Proof. The proof uses scaling and follows the same steps as the proof of Lemma3.3. For everyε >0 andy >1/ε we have

yδP1Y τ0

0

Ytdt > y2

yδP1Y τ0

0

Ytdt > y2|τεy< τ0

P1Yεy< τ0)

yδPεyY τ0

0

Ytdt > y2

(εy)δ=εδPεyY

τ0/(εy)

0

Yεysds > y ε

=εδPεyY

τ0/(εy) 0

Yεys

εy ds > ε2

=εδP1Y τ0

0

Ysds > ε2

>0.

This calculation, in fact, just shows that yδP1Y

τ0 0

Ytdt > y2

is a non-decreasing positive function ofy. Therefore, we only need to prove that it is bounded asy→ ∞. But for y >1

yδP1Y τ0

0

Ytdt > y2

=P1Y τ0

0

Ytdt > y2τy< τ0

+yδP1Y τ0

0

Ytdt > y2, τy> τ0

≤1+yδP1Y0> y, τy> τ0)≤1+yδP1Y0> y).

An application of Lemma3.4finishes the proof.

4. Conditions which imply Theorem2.2

We have shown that the diffusion processY has the desired asymptotic behavior of the area under the path up to the exit timeτ0Y. In this section we give sufficient conditions under which we can “transfer” this result to the processV and obtain Theorem2.2.

Lemma 4.1. Suppose that:

(A) There is a constantBsuch thatnδP0V0> n)Bfor alln∈N. (B) For everyε >0

nlim→∞PεnV σ

01

i=0

Vi> n2

=P1Y τ0

0

Ytdt > ε2

.

(C) limn→∞nδP0Vn< σ0)=C5. Then

nlim→∞nδP0V σ

01

i=0

Vi> n2

=C4C5,

whereC4is the constant from Lemma3.5.

Proof. Fix anε(0,1)and split the path-space ofV into two parts, the eventHn,ε:= {τεnV < σ0V}and its complement, Hn,εc .

(10)

First, consider the behavior of the total progeny on the eventHn,εc . OnHn,εc , the processV stays belowεnuntil the timeσ0. Estimating eachVi from above byεnand using (A) we get for alln∈N

nδP0V σ

01

i=0

Vi> n2, Hn,εc

nδP0V0> n/ε)(2ε)δB.

Therefore, for alln∈N 0≤nδP0V

σ

01

i=0

Vi> n2

nδP0V σ

01

i=0

Vi> n2, Hn,ε

(2ε)δB.

Hence, we only need to deal with the total progeny on the eventHn,ε. The rough idea is that, onHn,ε, it is not unnatural for the total progeny to be of ordern2. This means that the decay of the probability that the total progeny is overn2 comes from the decay of the probability ofHn,ε, which is essentially given by condition (C). This would suffice if we could letε=1 but we needεto be small, thus, some scaling is necessary to proceed with the argument, and this brings into play condition (B) and the result of Lemma3.5.

To get a lower bound onFn:=nδP0V(σ01

i=0 Vi> n2, Hn,ε)we use monotonicity ofV with respect to the initial number of particles, conditions (B) and (C), and Lemma3.5:

εlim0lim inf

n→∞ Fn≥ lim

ε0 lim

n→∞nδP0V(Hn,ε)PεnV σ

01

i=0

Vi> n2

=C5lim

ε0εδP1Y τ0

0

Ytdt > ε2

=C4C5.

For an upper bound onFnwe shall need two more parameters,K(1,1/ε)andR >1. At the end, after taking the limits asn→ ∞and thenε→0 we shall letK→ ∞andR→1.

nδFn=P0V τ

εn1

i=0

Vi+

σ01

i=τεn

Vi> n2, Hn,ε

P0V σ

01

i=τεn

Vi> n2(1Kε),

τεn1

i=0

ViKεn2, Hn,ε

+P0V τ

εn1

i=0

Vi> Kεn2, Hn,ε

.

We bound the first term on the right-hand side by the following sum:

P0V σ01

i=τεn

Vi> n2(1Kε), VτεnRεn, Hn,ε

+P0V σ01

i=τεn

Vi> n2(1Kε), Vτεn> Rεn, Hn,ε

. Estimating these terms in an obvious way and putting everything back together we get

nδFnPRεnV σ01

i=0

Vi> n2(1Kε)

P0V(Hn,ε)

+P0V(Vτεn> Rεn, Hn,ε)+P0V τ

εn1

i=0

Vi> Kεn2, Hn,ε

=(I )+(II)+(III).

(11)

It only remains to multiply everything bynδand consider the upper limits.

Termnδ(I )gives the upper boundC4C5in the same way as we got a lower bound by sendingn→ ∞,ε→0, and thenR→1 and using easily verified continuity properties of the relevant distributions. ParameterKdisappears when we letε→0.

Term(II)is exponentially small innfor fixedεandR(see Lemma5.1), thusnδ(II)goes to zero asn→ ∞. Finally, sinceViεnfor alli < τεn, we get

nδP0V τ

εn1

i=0

Vi> Kεn2, Hn,ε

nδP0Vεn> Kn, Hn,ε)

nδP0V0> Kn, Hn,ε)nδP0V0> Kn)≤2δB

Kδ.

5. Main tools

The main result of this section is Lemma5.3, which is a discrete analog of Lemma3.2(ii).

We start with two technical lemmas. The first one will be used many times throughout the paper.

Lemma 5.1. There are constantsc1, c2>0andN∈Nsuch that for everyxNandy≥0, sup

0z<x

PzV(Vτx > x+y|τx< σ0)c1

ec2y2/x+ec2y

, (5.1)

sup

x<z<4x

PzV(Vσxτ4x < xy)c1ec2y2/x. (5.2)

This statement is a consequence of the fact that the offspring distribution ofV is essentially geometric. The proof is given in theAppendix.

Lemma 5.2. Fixa(1,2].Consider the processV with|V0an| ≤a2n/3and letγ=inf{k≥0|Vk/(an1, an+1)}. Then for all sufficiently largen

(i) PV dist

Vγ,

an1, an+1

a2(n1)/3

≤exp

an/4

; (ii)

PV

Vγan1

aδ aδ+1

an/4.

Part (i) is an immediate consequence of Lemma5.1. The proof of part (ii) is basic but technical and is given in the Appendix.

Lemma 5.3 (Main lemma). For eacha(1,2]there is an0∈Nsuch that if, m, u, x∈Nsatisfy0 < m < u and|xam| ≤a2m/3then

ha(m)−1

ha(u)−1 ≤PxVa> τau)h+a(m)−1 h+a(u)−1, where

h±a(i)= i

r=+1

aδaλr

, i > ,

andλis some small positive number not depending on.

Références

Documents relatifs

Theorem 2.1 asserts a functional central limit theorem for the random walk XúJ when it is confined to an infinite equivalence class, under the assumption that the density

BD-LETTER.LISP The book design specifications for the Letter Generic, Letter Screen, and Letter LGP2 document device types.. Various com- mands and environments used only by the

Annealed and quenched fluctuations of transient random walks in random environment..

Carleson for having suggested the theme of this paper and for valuable guidance... These choices are clearly

These include among others, the distribution of first and last hitting times of the random walk, where the first hitting time is given by the first time the random walk steps into

We now study the scaling limits of discrete looptrees associated with large conditioned Galton–Watson trees, which is a model similar to the one of Boltzmann dissections: With

We show that the random stable lamination can be coded by the continuous-time height function associated to the normalized excursion of a strictly stable spectrally positive

We define a partition of the set of integers k in the range [1, m−1] prime to m into two or three subsets, where one subset consists of those integers k which are &lt; m/2,