• Aucun résultat trouvé

Shape transition under excess self-intersections for transient random walk

N/A
N/A
Protected

Academic year: 2022

Partager "Shape transition under excess self-intersections for transient random walk"

Copied!
29
0
0

Texte intégral

(1)

www.imstat.org/aihp 2010, Vol. 46, No. 1, 250–278

DOI:10.1214/09-AIHP318

© Association des Publications de l’Institut Henri Poincaré, 2010

Shape transition under excess self-intersections for transient random walk

Amine Asselah

Laboratoire d’Analyse et de Mathématiques Appliquées, Université Paris-Est, UMR CNRS 8050, France. E-mail:amine.asselah@univ-paris12.fr Received 3 May 2008; revised 10 February 2009; accepted 13 March 2009

Abstract. We reveal a shape transition for a transient simple random walk forced to realize an excessq-norm of the local times, as the parameterqcrosses the valueqc(d)=dd2. Also, as an application of our approach, we establish a central limit theorem for theq-norm of the local times in dimension 4 or more.

Résumé. Nous décrivons un phénomène de transition de forme d’une marche aléatoire transiente forcée à réaliser une grande valeur de la norme-qdu temps local, lorsque le paramètreqtraverse la valeur critiqueqc(d)=dd2. Comme application de notre approche, nous établissons un théorème de la limite centrale pour la norme-qdu temps local en dimension 4 et plus.

MSC:60K35; 82C22; 60J25

Keywords:Self-intersection local times; Large deviations; Random walk

1. Introduction

We consider a simple random walk{S(n), n∈N}onZd, starting at the origin. For any setA, we denote by1A the indicator ofA, and consider the local times of the walk{ln(z), z∈Zd}given by

ln(z)=1{S(0)=z}+ · · · +1{S(n1)=z}. (1.1) For a realq >1, we form the sum of theqth power of the local times

lnqq=

z∈Zd

ln(z)q. (1.2)

Whenq is integer, lnqq can be written in terms of the q-fold self-intersection local times of a random walk. For instance, whenq=2

ln22=n+2

0i<j <n

1{S(i)=S(j )}.

Forq positive real, we still calllnqqtheq-fold self-intersection local times.

In dimension three and more, Becker and König [6] have shown that there are positive constants, sayκ(q, d), such that almost surely

nlim→∞

lnqq

n =κ(q, d). (1.3)

(2)

Here, we are concerned with estimating the deviations oflnqq away from its mean. That is, ifP denotes the law of the walk started at 0, we give estimates for

P

lnqqE lnqq

ξ n

(1.4) forξ positive, andngoing to infinity.

There is a rich literature concerning the two-fold self-intersection local times. The reason being thatln2 is a natural object in quantum-field theory (see [1,14,22], for instance), as well as in the statistical physics of polymers (see [8,9,13], for instance). Howeverlnqforq∈R\Nhas no such direct links with physics. It comes up naturally in studying large and moderate deviations forrandom walk in random sceneries(see [5] and [15]).

Now, in the large deviations results for the two-fold self-intersection of a transient random walk (see [2,3,5,11]) two strategies have a distinguished role:

• Strategy A: the walk visits of the order of (ξ n)1/q-times, finitely many sites in a ball of bounded radius. For a transient walk, the number of visits of a bounded domain is bounded by a geometric variable. Thus, strategy A costs of the order of exp(−O((nξ )1/d)), where we use the notationyn=O(xn)for two positive sequences{xn, yn, n∈N}, to mean that there isK >0 such that 0≤ynKxn.

• Strategy B: the walk visits of the order of ξ1/(q1)-times, about n/ξ1/(q1) sites. Presumably, the walk stays, a timen, in a ball of volume n/ξ1/(q1). The cost of staying a time n within a ball of radiusrn

n is about exp(−O(n/rn2)), so that strategy B costs of the order of exp(−O(n12/dξ2/(d(q1)))).

Whenq=2, [2,5] have shown that strategy A is adopted ind≥5, whereas [3] (see also Chapter 8.4 of [11]) suggests that strategy B is adopted ind=3.

To summarize in words our main finding, assumed ≥3, fixξ >0 and look at typical paths realizing {lnqqE[lnqq] ≥ξ n}. As we increaseq, we step on a value,qc(d), above which our large deviation event is realized by strategy A, and below which it is realized by strategy B. The critical valueqc(d)=dd2 is obtained as we equal the costs of strategies A and B.

Note thatqc(d)is a well known number: ifqis integer, thenqindependent simple random walks, onZd, intersect infinitely often if and only ifq < qc(d)(see, for instance, [18], Proposition 7.1 and [17], Section 4.1).

Let us now describe, in mathematical terms, this shape transition. The first theorem deals with the sub-critical regimeq < qc(d).

Theorem 1.1. Assume dimensiond≥3.Then,forq anddsuch that1< q <dd2,there are constantsc±1(q, d) >0 such that forξ≥1,andnlarge enough

exp

c1(q, d)ξ(2/d)(1/(q1))n12/d

P

lnqqE lnqq

ξ n

≤exp

c+1(q, d)ξ2/d(1/(q1))n12/d

. (1.5)

Moreover,in this regime the sites visited more than some large constant do not contribute to realizing the excess self-intersection.In other words,

lim sup

A→∞ lim sup

n→∞

1

n12/d logP

z∈Zd

1{ln(z)>A}ln(z)qξ n

= −∞. (1.6)

Our second theorem deals with thesuper-critical regimeq > qc(d).

Theorem 1.2. Assume dimensiond≥3.Forqandd such thatq >dd2,there are constantsc±2(q, d) >0such that forξ≥1,andnlarge enough

exp

c2(q, d)(ξ n)1/q

P

lnqqE lnqq

ξ n

≤exp

c+2(q, d)(ξ n)1/q

. (1.7)

(3)

Moreover,the sites visited much less thann1/q do not contribute to realizing the excess self-intersection.In other words,

lim sup

ε0

lim sup

n→∞

1 n1/qlogP

z∈Zd

1{ln(z)<εn1/q}ln(z)qE lnqq

+ξ n

= −∞. (1.8)

Remark 1.3. In Theorems1.1and1.2,we could takeξ to grow withn.The only(necessary)bound onξncomes from the boundlnqnwhich imposes thatξnnq1.The proofs are written with generalξn≥1.

The next result deals with the contribution of some level sets of the local times to deviation on a much larger scale than the mean, and can be obtained by the same approach yielding Theorem1.2. We include it in this form since it can be of independent interest, while showing the possibilities offered by our approach. Also, it generalizes Lemma 1.8 of [5].

Lemma 1.4. Assumed≥3andqqc(d).Choosea, b >0such that1< a <1+b(q−1).Then,for anyε >0,and nlarge enough

P

z∈Zd

1{ln(z)<nb}ln(z)qna

≤enζ (q,a,b)ε withζ (q, a, b)=b+ 1

qc(d)(aqb). (1.9)

Remark 1.5. Our approach is not suited to studying smallξn for reasons explained later in Remark1.7.However, when1> ξnnδ,for some positiveδsmall enough,our approach yields a constantc1such that forq < qc(d)

P

lnqqE lnqq

ξnn

≤exp

c1ξn2/d(q/(q1))n12/d

. (1.10)

Whenq > qc(d),we have a constantc2such that P

lnqqE lnqq

ξnn

≤exp

c2ξn1/q+2/dn1/q

. (1.11)

We believe that the powers ofξnin(1.10)and(1.11)are not optimal.However, (1.10)and(1.11)are useful in deriving a central limit theorem stated in Theorem1.9.

Our initial goal was to improve the main result of [3], which states that in dimension 3, there isχ >0 andε >0 such that forξ >0, andnlarge

P

z∈Z3

1{ln(z)>log(n)χ}l2n(z) > nξ

≤exp

n1/3log(n)ε

. (1.12)

Note that (1.6) improves (1.12). One reason to studylnq forq >2, is that the upper bound (1.5) forq >2, yields (1.6) at once. More precisely, forq < qc(d), chooseqwithq < q< qc(d), and for anyA >0, the obvious inequality

z∈Zd

1{ln(z)>A}lqn(z)lnqq

Aqq, (1.13)

implies that P

z∈Zd

1{ln(z)>A}lnq(z)

P

lnqqAqq .

(4)

ForAandnlarge enough,Aqq≥2E[lnqq]. Thus, if we setβ=2dqqq1>0, then from (1.5), we have a constant c1(d, q)such that

P

z∈Zd

1{ln(z)>A}lnq(z)

≤exp

c1 d, q

ξ(2/d)(1/(q1))Aβn12/d

. (1.14)

Thus, in order to improve (1.12) ind=3, we were left with studyingq-fold self-intersections with 2< q <3=qc(3).

In most works on two-fold self-intersection, a starting point, which we trace back to the work of Westwater [21]

and Le Gall [19], is a decomposition ofln22in terms ofintersection local timesof two independent random walks starting at the origin. However, such a decomposition is restricted toq-fold self-intersection local times withq∈N:

Whenq =2 and d =3 (in the sub-critical regime) Le Gall’s decomposition is a first step in obtaining, in [11], a moderate and large deviations principles. Whenq=3 andd≥4 (in thesuper-critical regime), [15] uses a type of Le Gall’s decomposition to obtain moderate and large deviations estimates.

Here, our starting point is an approximate decompositionobtained by slicing lnqq over level sets of the local times, for any realq >1. This is based on the following simple inequality. Let{bn, n∈N}be a subdivision of[1,∞), and letl1andl2be positive integers (which we think of as the local times of a given site in each half time-period).

Then, forq >1, we have the upper bound (l1+l2)ql1q+l2q+2q

i=0

biq+121{bimax(l1,l2)<bi+1}l1×l2, (1.15)

as well as the obvious lower bound:(l1+l2)ql1q+l2q. The desirable feature of (1.15) is that on its right-hand side, theqth power ofl1andl2comes without penalty, whereas the terml1×l2yields an intersection local times. Thus, (1.15) leads to the following result which plays here the role of Le Gall’s decomposition of [19].

Proposition 1.6. For any integersnandl,with2l< n,let{ni, i=1, . . . ,2l}be positive integers summing up ton.

Let{l·(i), i=1, . . . ,2l}be the local times of2l independent random walks starting at0.If{bi, i∈N}is a subdivision of[1, n],then,

Sq(l)lnqqSq(l)+ l j=1

Ij, whereS(l)q law=

2l

i=1

l(i)n

i q

q, (1.16)

and,forj=1, . . . , l,andmk=n(k1)2lj+1+ · · · +nk2lj fork=1, . . . ,2j Ij

law=

2j1

k=1

i

2qbiq+11

z:bilm(2k)

2k(z)<bi+1

l(2km 1)

2k1 (z)+

z:bil(2km2k−11)(z)<bi+1

l(2k)m

2k(z)

. (1.17)

Remark 1.7. We first note some natural limitations in using the approximate decomposition (1.16).When we deal with{lnqqE[lnqq] ≥ξnn}for smallξn,we need to bound the difference betweenE[lnqq]and the expectation of the upper bound in(1.16).When,we takelsuch that2ln1δ0,then this difference turns out to be of order smaller thann1δ0/2,allowing us to write

lnqqE lnqq

ξnn

Sq(l)E Sq(l)

ξn 2n

l

j=1

IjE[Ij] ≥ξn

2nn1δ0/2

. (1.18)

(1.18)requires thatξnnδ0/2.

Proposition1.6is our initial step in the proof of Theorems1.1and1.2, and leads to a central limit theorem (CLT) forlnqqin dimension 4 or more, as well as a characterization of the variance oflnqq.

(5)

Before stating results concerning the typical behavior of lnqq, we give a heuristic discussion of the proof of Theorem1.1assuming Proposition1.6. More precisely, we wish to sketch the reasons why theapproximate decompo- sition(1.16), reduces large deviations for theq-norm of the local times, to large deviations for a sum of independent geometric variables.

Consider a choice oflsuch that 2lis close tonin (1.16), andnin/2l. Then,Sq(l)is a sum of aboutnindependent terms, each one bounded by its time-span,n/2l, to the powerq. Recall that the probability of deviating from the mean, for a sum ofnindependent and essentially bounded variables, is of order exp(−O(n))(see Lemma2.4for a precise statement). We can therefore neglect the contribution ofS(l)q to the excessq-norm, though the mean ofS(l)q is close toE[lnqq], as easily seen from Lemma1.8below. Now, for a fixedj in{1, . . . , l}, letm=n/2j, and note thatIj

is a finite sum of independent terms distributed asbq1lm(D˜(b)), whereD˜(b)is the set{z: l˜m(z)b}, withl˜man independent copy oflm, andbspans a subdivision of[1, n]. Since we consider transient random walks,lm(D˜(b))is bounded by a geometric variable (when fixingD˜(b), as shown in Lemma 1.2 of [4]). At this point, one normalizes lm(D˜(b)). IfP0andP˜0are the law of the two independent copies of the (transient) walk started at 0, define

X= lm(D˜(b))

E0[lm(D˜(b))], so that for someκ >0, P˜0P0(X > t)≤eκt. (1.19) Now, it is well known that for anym,E0[lm(D˜(b))] ≤C| ˜D(b)|2/d, (see LemmaA.2). Thus, an estimate of the large deviation probability requires an estimate on the volume of level sets of the local times. Now, in obtaining a bound on the volume ofD˜(b), assume for simplicity that we only have two types ofb: that is, we distinguishoften visited sites, say sites visitednx-times withxclose to 1/q, whose level sets are part of what we calltop levels, and say the once-visited sites, whose level sets are part of what we callbottom levels. Thebottom levelsare the easiest to treat (see Section3.2and Lemma3.1). Indeed, we use essentially that| ˜D(b)| ≤nforb∼1, so that we expect (when we restrictIj only tobottom levelsand usingXk=Xin law)

P

IjE[Ij] ≥

P 2j

k=1

XkE[Xk] ≥ n2/d

∼exp

−O

n12/d .

Thetop levels, treated in Section3.3, require a type ofbootstrap argument, using that ifD˜(b)has a large volume, it implies that theq-norm of the local timel˜mis large. The bootstrap argument yields Corollary2.2. It allows us to normalize properly our geometric-like random variablesbq1lm(D˜(b)).

We turn now to the typical behavior of the q-norm. Chen has provided in [10] asymptotics for the variance of ln22ind≥3. He shows that (i) ind=3, var(ln22)λ3nlog(n), and (ii) in d≥4, var(ln22)λdn, whereλd are constants expressed in terms of the Green’s function of the walk. Following ideas of Jain and Pruitt [16], and of Le Gall and Rosen [20], Chen obtains a CLT in dimension 3 or more for ln22. Finally, Becker and König in [6]

have shown that forqinteger, (i) ind=3, var(lnqq)n3/2, (ii) ind=4, var(lnqq)nlog(n), and (iii) ind≥5, var(lnqq)cdn. Our result deals with the general case (q >1 real), where no representation oflnqq is possible in terms of multiple time-intersections. We transform Lindeberg’s condition into alarge deviationevent forlTn22on the scale of time of the CLT, that isTn≈√

n.

We start with an estimate for the expectation oflnqq, of the same type as Theorem 1 of Dvoretzky and Erdös [12]

for the range of a transient random walk. Thus, ifγdis the probability of never returning to its original position, it is shown in [12] that for positive constantscd, whenRnis the set of visited sites before timen,

E

|Rn|

dcdψd(n), withψd(n)=

⎧⎨

n1/2 for d=3, log(n) for d=4,

1 for d≥5,

(1.20)

Jain and Pruitt [16] obtain the asymptotics var(|Rn|)alog(n)n for somea >0 ind =3, and var(|Rn|)cdn in d >3, for some positive constantscd. The corresponding CLT (ind≥3) was shown by Jain and Pruitt [16] for the simple random walk, and by Le Gall and Rosen [20] for stable random walks. Note that the limiting law is Gaussian, ind≥3 but fails to be so ind=2, as shown by Le Gall in [18].

(6)

Lemma 1.8. Assume thatd≥3andq >1.There is a constantCd,such that 0≤κ(q, d)nE

lnqq

Cdψd(n), withκ(q, d)=γdE l(0)q

. (1.21)

Also,ifd=3,then,there is a constantc3such that var

lnqq

c3log(n)2n. (1.22)

Ifd≥4,then there are positive constantsv(q, d)andc(q, d),such that var(lnqq)

nv(q, d)

c(q, d)log(n)

n . (1.23)

Finally, we have the following central limit theorem.

Theorem 1.9. IfZis a standard normal variable,then lnqqnκ(q, d)

nv(q, d)

−→law Z. (1.24)

A challenging open question is to understand the strategy which realizes {lnqqE[lnqq] ≥ξ n}, right at the critical valueq=qc(d)=dd2.

The paper is organized as follows. The approximate decomposition oflnqq is given in Section2.1. The sub- critical regime is studied in Section3: The upper bound in (1.5) is proved in Section3.5, and the lower bound is given in Section3.4. The super-critical regime is studied in Section4. Theorem1.2is proved in Section4. The proof of (1.8) is given in Section4.1. The proof of Lemma1.4is given in Section4.2. Lemma1.8, as well as the CLT are proved in Section5. In theAppendix, we recall Lemma 5.1 of [3], and improve Lemma 5.3 of [3], used to control intersection local times-type quantities.

2. General considerations (q >1)

In this section, we deal with the general caseq >1. In Section2.1, we develop a approximation oflnqqas sums of two types of independent variables:

1. Intersection local times of independent walks.

2. Self-intersection local times, on a much shorter time-period.

In Section2.2, we treat the sums of self-intersection local times.

2.1. Approximate decomposition forlnqq

Before we prove Proposition1.6, we present a useful corollary which requires more notations.

For integersnandl, with 2l< n, we recall the “almost” dyadic decomposition ofnof Remark 2.1 of [5]. We divide ninto 2lintegersn(l)1 , . . . , n(l)

2l withn=n(l)1 + · · · +n(l)

2l and maxi

n(l)i

−min

i

n(l)i

≤1, n

2l −1≤n(l)in

2l +1 and n(lk1)=n(l)2k1+n(l)2k. (2.1) We run 2lindependent random walks starting at the origin. Theith walk runs for a time-period[0, n(l)i [, and we denote byli(l):Zd→Nits local times during time-period[0, n(l)i [. Also, we introduce, fork=1, . . . ,2l, the following sets

D(l)k,i=

z∈Zd: bilk(l)(z) < bi+1

. (2.2)

Now, for anyM >0, let{bi, i∈N}be a subdivision of[1, M], and denote byΘM(x)=x1{xM}.

(7)

Remark 2.1. We could restrict the sum overZdwhich enterslnqq in(1.16)over{z: ln(z)M}for any positiveM.

The proof of Proposition1.6yields,for any{bi, i∈N}subdivision of[1, M],

z∈Zd

ΘM

ln(z)q

2l

k=1

z∈Zd

ΘM

lk(l)(z)q

+ l j=1

Ij. (2.3)

The only difference with(1.16)is the subdivision which enters into the definition ofIj.The proof of Proposition1.6 is written in view of(2.3) (see the key step(2.12)).

As a corollary of (2.3), we obtain the following result.

Corollary 2.2. For anyM >0,let{bi, i∈N}be a subdivision of[1, M].For any integersnandL,with2L< n,and for any sequence of positive numbers{mn, εn, n∈N},we have

P ΘM(ln) q

qmn+εn

≤2L+1P 2L

j=1

ΘM lj(L) q

qmn

+ L h=1

2hP L

l=h

1{G(l)

1 ∩G2(l)}Ilεn

, (2.4)

where forlL,k=1, . . . ,2l,andi∈N G(l)k,i=Dk,i(l)mn+εn

bqi

, G1(l)=

2l

k=1

i

G2k,i(l) and G2(l)=

2l

k=1

i

G2k(l)1,i. (2.5)

Remark 2.3. The symbols,εnandmn,are suggestive of the fact that whenLis large enough,the sum of2Lindepen- dentq-fold self-intersections,that we calledSq(L),stays close to its mean,which is also close to the mean oflnqq. This is shown in Section2.2.So,mn stands for mean,andεnstands for excess.To estimate how small canεnbe,we now compute the expectation ofL

l=1Il.We use LemmaA.2in the worse case,that is in dimension3,to obtain for some constantsc1, c2andc3

E L

l=1

Il

= L

l=1

2l

i∈N

2q(bi+1)q1Cdψd n

2l

eκdbi

c1n

L l=1

2l/2

i∈N

(bi+1)q1eκdbi

c2 2Ln

i∈N

(bi+1)q1eκdbi. (2.6)

First,we need to choose a subdivision{bi, i∈N}such that the last sum in(2.6)is convergent.Secondly,the right-hand side of(2.4)is small ifL

l=hE[1{G1(l)G2(l)}Il] εn.From(2.6),we see thatεn can be chosen small if2Ln.

On the other hand,we see in Section2.2,thatL has to be large enough,for the probability of {Sq(L)mn}to be negligible,whenmn=E[lnqq] +nε. Remark2.5shows that a choice ofL such that2L> n1δ0 withqδ0<2d, fulfills both requirements.

Proof of Proposition1.6. The proof proceeds by induction onl≥1. It is however easy to see that proving the case l=1 requires the same arguments as going froml−1 tol. We focus on the first stepl=1, and omit the easy passage froml−1 tol.

For anyx∈ [0,1], andq≥1, we have

1+xq(1+x)q≤1+xq+2qx. (2.7)

(8)

Thus, for any nonnegative integersl1, l2with 0≤l1, l2M, we have from (2.7)

l1q+l2q(l1+l2)ql1q+lq2+2qMq2l1l2. (2.8) Now, for anyM >0, let{bi, i∈N}be a subdivision of[1, M], and recall thatΘM(x)=x1{xM}. For any nonnegative integersl1, l2

ΘM(l1+l2)q

ΘM(l1)q

+

ΘM(l2)q

+2q n i=1

biq+121

bi≤max(l1, l2) < bi+1

l1×l2. (2.9)

Indeed,l1+l2Mandl1, l2≥0, imply (i)l1M andl2M, and (ii) for somei0>0, max(l1, l2)∈ [bi0, bi0+1[. Then, from (2.8)

ΘM(l1+l2)q

ΘM(l1)q+ΘM(l2)q+2qbqi2

0+1l1×l2.

For any integern, we consider the local timeln, which we denote asl[0,n[(z)to emphasize the time period. For any integern1with 0< n1< n, setn2=nn1, and from the increments of our initial random walk, say{Yn, n∈N}, we build two independent random walks with local times

l](1,1)0,k](z)=1{Yn1=z} + · · · +1{Yn1+ · · · +Yn1k+1=z} and

l[(1,2)0,k[(z)=1{0=z} +1{−Yn1+1=z} + · · · +1{−Yn1− · · · −Yn1+k1=z}. (2.10) It is obvious that on{S(n1)=y},

ln(z)=l](1,1)0,n

1](yz)+l[(1,2)0,n

2[(yz). (2.11)

If we denotel¯(1)(z)=max(l](1,1)0,n

1](z), l[(1,2)0,n

2[(z)), and sum (2.11) overz∈Zd, we obtain

z

ΘM

ln(z)q

z

ΘM

l](1,1)0,n

1]

S(n1)zq

+

z

ΘM

l[(1,2)0,n

2[

S(n1)zq

+2q

z∈Zd

n i=1

bqi+121{b

i≤¯l(1)(S(n1)z)<bi+1}l(1,1)]0,n

1]

S(n1)z

×l[(1,2)0,n

2[

S(n1)z

z∈Zd

ΘM l(1,1)]0,n

1](z)q

+

z∈Zd

ΘM l[(1,2)0,n

2[(z)q

+2q

z∈Zd

n i=1

bqi+121{b

i≤¯l(1)(z)<bi+1}l](1,1)0,n

1](z)×l[(1,2)0,n

2[(z). (2.12)

Now, we rewrite (2.12) in a concise form as ΘM(ln) q

qΘM l(1,1)]0,n

1] q

q+ ΘM l[(1,2)0,n

2[) q

q+ ˜I1(n1, n2), (2.13)

where the term dealing with intersection times of independent strands is I˜1(n1, n2)=2q

z∈Zd

n i=1

biq+121{b

i≤¯l(1)(z)<bi+1}l(1,1)]0,n

1](z)×l[(1,2)0,n

2[(z). (2.14)

(9)

To prove the upper bound in (1.16) forl=1, it suffices to setM=n(so that this truncation disappears), and to show thatI˜1(n1, n2)I1given in (1.17). This latter inequality follows first by noting the obvious inclusion

z: bi≤max l](1,1)0,n

1](z), l(1,2)[0,n

2[(z)

< bi+1

z: bil](1,1)0,n

1](z) < bi+1

z: bil[(1,2)0,n

2[(z) < bi+1 , and secondly, using that

z∈Zd

1{b

i≤¯l(1)(z)<bi+1}l](1,1)0,n

1](z)×l(1,2)[0,n

2[(z)

z∈Zd

1

bil](1,1)0,n

1](z) < bi+1

bi+1l(1,2)[0,n

2[(z) +

z∈Zd

1

bil[(1,2)0,n

2[(z) < bi+1

bi+1l](1,1)0,n

1](z). (2.15)

As we iterate the approximate decompositionl-times, we obtain the upper bound in (1.16), or more generally the bound (2.3).

The lower bound in (1.16) is an obvious corollary of the inequality(l1+l2)ql1q+l2q, valid forq≥1 andl1, l2

nonnegative.

Proof of Corollary2.2. We write the caseM=n, that is the case with no truncation. The case with truncation is obtained as we replaceln(z)byΘM(ln(z))wherever it appears. Assume that we stop the induction in Proposition1.6 at some stepL (typically 2L=n1δ0 andδ0 small). For any sequence of positive numbersεn, mn, we have from (1.16),

P

lnqqmn+εn

P

Sq(L)mn +P

L

l=1

Ilεn

, whereSq(l)=

2l

k=1

l(l)

k q

q. (2.16)

We introduce, as in [3,5], a bootstrap control on the volume ofDk,i(l). ConsiderG(l)k,igiven in (2.5). On the complement ofG(l)=G1(l)G2(l), there isk0, i0such that|Dk(l)

0,i0|> (mn+εn)/biq

0 so that Sq(l)

zDk(l)

0,i0

l(l)k

0(z)q

mn+εn

bqi

0

bqi

0=mn+εn. (2.17)

WritingSq(0)= lnqq, we write a more suggestive relation

P

Sq(0)mn+εn

P

Sq(L)mn

+P L

l=1

1{G(l)}Ilεn

+

L l=1

P

Sq(l)mn+εn

. (2.18)

Starting the approximation withSq(l)withl < L, we obtain similarly

P

Sq(l)mn+εn

P

Sq(L)mn +P

L

j=l+1

1{G(j )}Ijεn

+ L j=l+1

P

Sq(j )mn+εn

. (2.19)

Assume now that forj > l, andj < Lwe have

P

Sq(j )mn+εn

≤2Lj+1P

Sq(L)mn

+ L h=j+1

2hj1P L

i=h

1{G(i)}Iiεn

. (2.20)

Références

Documents relatifs

In Section 5, we treat the problem of large and moderate deviations estimates for the RWRS, and prove Proposition 1.1.. Finally, the corresponding lower bounds (Proposition 1.3)

We obtain estimates for large and moderate deviations for the capacity of the range of a random walk on Z d , in dimension d ≥ 5, both in the upward and downward directions..

As we shall see the proposition is a consequence of the following more general result concerning the functional moderate deviation principle of an array of martingale

They suggest that conditioned on having a small boundary of the range, in dimension three a (large) fraction of the walk is localized in a ball of volume n/ε, whereas in dimension

We shall apply Shao’s method in [14] to a more general class of self-normalized large deviations, where the power function x 7→ x p used as a generator function for the normalizing

Large deviations principle for self-intersection local times for simple random walk in dimension d &gt; 4. Large deviations for the self-intersection times for simple random walk

We prove a level 3 large deviation principle, under almost every environment, with rate function related to a relative

Keywords: Random walk in random environment; Law of large numbers; Large deviations; Galton–Watson tree.. In particular, the tree survives