• Aucun résultat trouvé

An alternative to the coupling of Berkes-Liu-Wu for strong approximations

N/A
N/A
Protected

Academic year: 2021

Partager "An alternative to the coupling of Berkes-Liu-Wu for strong approximations"

Copied!
21
0
0

Texte intégral

(1)

HAL Id: hal-01565627

https://hal.archives-ouvertes.fr/hal-01565627

Preprint submitted on 20 Jul 2017

HAL

is a multi-disciplinary open access archive for the deposit and dissemination of sci- entific research documents, whether they are pub- lished or not. The documents may come from teaching and research institutions in France or

L’archive ouverte pluridisciplinaire

HAL, est

destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires

An alternative to the coupling of Berkes-Liu-Wu for strong approximations

Christophe Cuny, Jérôme Dedecker, Florence Merlevède

To cite this version:

Christophe Cuny, Jérôme Dedecker, Florence Merlevède. An alternative to the coupling of Berkes-

Liu-Wu for strong approximations. 2017. �hal-01565627�

(2)

An alternative to the coupling of Berkes-Liu-Wu for strong approximations

Christophe Cuny

a

, J´ erˆ ome Dedecker

b

and Florence Merlev` ede

c

July 20, 2017

Abstract

In this paper we propose an alternative to the coupling of Berkes, Liu and Wu [1] to obtain strong approximations for partial sums of dependent sequences. The main tool is a new Rosen- thal type inequality expressed in terms of the coupling coefficients. These coefficients are well suited to some classes of Markov chains or dynamical systems, but they also give new results for smooth functions of linear processes.

a Universit´e de la Nouvelle-Cal´edonie, Institut de Sciences Exactes et Appliqu´ees.

Email: christophe.cuny@univ-nc.nc

b Universit´e Paris Descartes, Sorbonne Paris Cit´e, Laboratoire MAP5 (UMR 8145).

Email: jerome.dedecker@parisdescartes.fr

c Universit´e Paris-Est, LAMA (UMR 8050), UPEM, CNRS, UPEC.

Email: florence.merlevede@u-pem.fr

1 Introduction

Let (εi)iZbe a sequence of independent and identically distributed (iid) random variables, and let (Xn)nZ be a strictly stationary sequence such that

Xn =f(. . . , ε0, . . . , εn1, εn), (1) for some real-valued measurable functionf.

In 2005, Wu [9] introduced the so-calledphysical dependence measuredefined in terms of the following coupling: letε0 be distributed asε0and independent of (εi)iZ, and let

n =f(. . . , ε1, ε0, ε1, . . . , εn1, εn). (2) The physical dependence coefficents in Lp (assuming that kX0kpp = E(|X0|p) < ∞) are then given by

δp(n) =kXn−X˜nkp.

As pointed out by Wu, the coefficientδp(n) can be computed for a large variety of examples, including iterated random functions and functions of linear processes. As we shall see, it is particularly easy to compute whenXn is a smooth function of a linear process.

LetSn=Pn

k=1Xk. In a recent paper, Berkes, Liu and Wu [1] use the coupling defined above to prove the following strong approximation result: under an appropriate polynomial decay of the coefficientsδp(n), the sequence n1E (Sn−nE(X1))2

converges to σ2 as n → ∞ and, if

(3)

σ2 > 0, one can redefine (Xn)n1 without changing its distribution on a (richer) probability space on which there exist iid random variables (Ni)i1 with common distribution N(0, σ2), such that, Sn−nE(X1)−

Xn i=1

Ni

=o

n1/p P-a.s.

The proof (from which we extract Proposition 13, Section 4) is based on a approximation by m-dependent sequences combined with an application of a deep result by Sakhanenko [6].

This is a very important result, because it gives a full extension of the Komlos, Major and Tusnady (KMT) strong approximation [5] for partial sums of iid random variables in Lp (for which δp(n) = 0 if n ≥ 1). Most of the previous results in the dependent context were limitated to the raten1/4, because they were based on the Skorokhod representation theorem for martingales. As an exception, let us mention the paper [7], where the rateO(logn) is reached for bounded observables of geometrically ergodic Markov chains.

In this paper, we follow the main steps of the proof of Berkes, Liu and Wu [1], but we use a different coupling. Let (εi)iZ be an independent copy of (εi)iZ and let

Xn=f(. . . , ε2, ε1, ε0, ε1, . . . , εn1, εn). Our coupling coefficient inLp is then defined as

˜δp(n) =kXn−Xnkp

Asδp, this coefficient can be computed for a large class of examples (see Section 3).

Notice thatE(Xn1, . . . , εn) =E( ˜Xn1, . . . , εn) a.s., and thatkXn−E(Xn1, . . . , εn)kp

˜δp(n), from which we easily deduce that δp(n)≤2˜δp(n). Moreover, it seems natural to think that, in many situations, the coefficientδp(n) should be much smaller thanδp(n), becauseXn

differs from ˜Xnby changing only the coordinate at point 0, while all the coordinates before time 0 have been changed inXn. For Markov chains, however, the two coefficients should be of the same order (we shall give in Section 2 an alternative definition of ˜δp(n), which is more adapted to the Markovian setting).

Sinceδp(n)≤2˜δp(n), a reasonable question is then: what could be the interest to deal with

˜δp(n)? The answer is simple : the couplingXn is often easier to handle thanXn (becauseXnis by definition independent of the pastσ-algebraF0 =σ(εi, i≤0)) and we can develop specific tools involving the coefficent ˜δp(n). In this paper, we shall prove and use a new Rosenthal-type inequality (see Section 5.1) expressed in terms of the coefficients ˜δ2 and ˜δp. As a consequence, the conditions that we impose on ˜δp(n) are weaker than the corresponding conditions onδp(n) in the paper by Berkes, Liu and Wu. The two results are not comparable, but we shall obtain better conditions in all the cases where δp(n) and ˜δp(n) are exactly of the same order (for instance in the case of Markov chains).

Let us present a simple example where our conditions are less restrictive than those of Berkes, Liu and Wu. Assume that

Xn=g X i=0

aiεni

!

where (ai)i0∈ℓ1, and (εi)iZis a sequence of iid random variables inLp. Heregis a continuous function such that

|g(x)−g(y)| ≤c(|x−y|)

wherecis a non-decreasing concave function andc(0) = 0 (cis then a concave majorant of the modulus of continuity ofg). In that case, using Lemma 5.1 in [4], it is easy to see that

δp(n)≤ kc(|an0−ε0)|)kp≤c(2kε0kp|an|),

(4)

and δ˜p(n)≤

c

X i=n

aini−εni)

!

p

≤c

X i=n

aini−εni) p

≤c

Cp0kp vu utX

i=n

a2i

, where we have used Burkholder’s inequality for the last upper bound (the positive constantCp

depends only on p). As expected, we see that the upper bound for δp(n) is smaller than the upper bound forδp(n).

Let us consider now the case where |ai| =O(ia) for some a > 1, and c(x)≤ C|x|β in a neighborhood of 0 for someβ∈(0,1]. In that case, the conditions of Berkes, Liu and Wu onδp

hold provided

a > 2

β forp∈(2,4], and a >τ(p) + 1

β forp >4, where

τ(p) := (p−2)p

p2+ 20p+ 4 +p2−4

8p ,

while our condition on ˜δp are satisfied as soon as a > κ(p)

β +1 2, where, forp >2,

κ(p) := (p−2)p

p2+ 12p+ 4 +p2+ 4p−4

8p . (3)

As one can see, our condition on ais always less restrictive: for p∈(2,4] it suffices to notice thatκis increasing andκ(4)<1.4. Forp >4, it suffice to notice thatκ(p)< τ(p) + 0.5. Note that, sinceκ(p)→1/2 asp→2, in the case whereβ= 1 (Lipschitz observables), we only need a >1 and a moment of orderb >2 forε0to get a strong approximation of ordero(n1/(2+ǫ)) for someǫ >0.

In addition to this example of functions of linear processes, we shall apply our main results to some classes of Markov chains or dynamical systems. The Markov chains we shall consider are not (or have no reasons to be) irreducible, and some kind of regularity on the observables is required (as in the previous example). We shall express these regularity conditions in terms of the modulus of continuity (orLp-modulus of continuity) of the observables. These examples of Markov chains are different from the examples we considered in the previous paper [3], where we used the coefficient ˜δ1. On the one hand, the coupling coefficient ˜δ1(n) can be computed for a larger class of examples, but on the other hand we need to impose a moment condition related topand to the decay rate of ˜δ1(n) to get the strong approximation with rateo(n1/p).

In all the paper, we shall use the notationan≪bn, which means that there exists a positive constantC not depending onnsuch thatan≤Cbn, for all positive integersn.

2 Main results

Before giving our first main result, let us give the appropriate definition of the coefficient ˜δp

when (Xn)nZ is a stationary sequence such that

Xn:=f(εn, εn+1, . . .) (4)

for some measurable real-valued functionf. This representation will play an important role in the application to certain non-invertible dynamical systems.

(5)

Recall that (εi)i0 is a sequence of iid random variables, and that (εi)i0is an independent copy of (εi)i0. Define then ˜X1,n :=f(ε1, . . . , εn, εn+1, εn+2, . . .). Then, for everyp≥1, the coefficient ˜δp(n) is defined by:

˜δp(n) :=kX1−X˜1,nkp. (5) Recall also that, for anyp >2, the fonctionκ(p) has been defined in (3).

Theorem 1 Let (Xn)nZ be a stationary sequence defined by either (1) or (4), and assume thatX0 has a moment of orderp >2. Assume in addition that there exists a positive constant csuch that for anyn≥1,

δ˜p(n)≤cnγ, (6)

for someγ > κ(p). LetSn =Pn

k=1Xk. Thenn1E (Sn−nE(Y1))2

→σ2 asn→ ∞ and one can redefine(Yn)n0 without changing its distribution on a (richer) probability space on which there exist iid random variables(Ni)i1 with common distribution N(0, σ2), such that,

Sn−nE(X1)− Xn i=1

Ni

=o

n1/p

P-a.s.

Remark 2 Concerning the function κ, note that κ(p)<(p+ 4)/4 and that the function p→ (p+ 4)/4 is an asymptot of κasp→ ∞.

As quoted in the introduction, we shall now consider the case where the variables Xn are functions of random iterates. In that case the representation (1) is not necessarily appropriate (nor even easy to establish), and we need to define an appropriate coefficientδp similar to ˜δp.

Let (εi)i1 be iid random variables with values in a measurable spaceGand common dis- tributionµ. Let W0be a random variable with values in a measurable spaceX, independent of (εi)i1and letF be a measurable function fromG×X toX. For anyn≥1, define

Wn=F(εn, Wn1), (7)

and assume that (Wn)n1has a stationary distributionν. Let nowhbe a measurable function fromG×X to Rand define, for anyn≥1,

Xn=h(εn, Wn1). (8)

Then (Xn)n1 is a stationary sequence with stationary distribution, sayπ. Let (Gi)iZ be the non-decreasing filtration defined as follows: for anyi <0,Gi={∅,Ω},G0=σ(W0) and for any i≥1,Gi=σ(εi, . . . , ε1, W0). It follows that for anyn≥1,Xn isGn-measurable.

LetW0 andW0 be two random variables with lawν, and such that W0 is independent of (W0,(εi)i1). For anyn≥1, let

Xn=h(εn, Wn1) withWn=F(εn, Wn1). We then define the coefficients (δp(n))n0 as follows

δp(0) :=kX1kp and δp(n) := sup

knkXk−Xkkp, n≥1. (9) It is not difficult to see that, for any positive integern,

δp(n)≤2kXn−Xnkp. (10) For such functions of random iterates, the following counterpart of Theorem 1 holds:

(6)

Theorem 3 Let(Xn)n1be a stationary sequence defined by (8)and assume that its stationary distribution π has a moment of order p > 2. Assume in addition that there exists a positive constantc such that for anyn≥1,

δp(n)≤cnγ, (11)

for someγ > κ(p). LetSn =Pn

k=1Xk. Thenn1E (Sn−nE(X1))2

→σ2 asn→ ∞and one can redefine(Xn)n1 without changing its distribution on a (richer) probability space on which there exist iid random variables(Ni)i1 with common distribution N(0, σ2), such that,

Sn−nE(X1)− Xn i=1

Ni

=o

n1/p P-a.s.

3 Applications

3.1 Applications to contracting iterated random functions.

We use the notations from the second part of Section 2, with a Markov chain (Wn)n1defined by the recursive equation (7) and a sequence (Xn)n1 defined by (8). Assume thatX is equipped with a metric d and that it is endowed with the corresponding Borelσ-algebra. Let us fix a

“base point”x0∈X. For everyx∈X, writeχ(x) := 1 +d(x0, x).

Let us assume that there existsC >0,ρ∈(0,1) andα≥1, such that Z

G

χ(F(g, x0))α

µ(dg)<∞, (12)

and

E d(Wn,x, Wn,y)α

≤Cρn(d(x, y))α, (13)

whereWn,xis the chain defined by (7) starting fromW0=x.

The next lemma is a combination of Theorem 2 and Lemma 1 of Shao and Wu [8].

Lemma 4 Assume that (12) and (13) hold for some α ≥ 1, C > 0 and ρ ∈ (0,1). Then the Markov chain (Wn)nN admits a stationary distribution ν such that R

X(χ(x))αν(dx)<∞. Moreover, there existC(α)>0 andρ(α)∈(0,1), such that for every n≥1,

Z Z

E (d(Wn,x, Wn,y))α

ν(dx)ν(dy)≤C(α)(ρ(α))n.

From now, the sequence (Xn)n1 is defined by (8), where the chain (Wn)n0 is strictly stationary, with stationary distributionν.

We shall say that a functionh : G×X→Rsatisfies the assumptionHs,t, for somes, t≥0 if there exist non-negative functionsη and ˜η and a non-decreasing functionβ : [0,1]→[0,+∞) such that, for every (g, x)∈G×X,

h(g, x)≤η(g)(χ(x))s, (14)

and for everyu∈(0,1], sup

yX:d(x,y)u

h(g, x)−h(g, y)≤β(u)˜η(g)(χ(x))t. (15)

Lemma 5 Let p≥1. Assume that (12)and (13)hold for someα≥1. Assume that (14)and (15)hold for some 0≤s < α/p,0≤t≤α/pand someη,η˜such that R

G(η(g))pµ(dg)<∞ and R

G(˜η(g))pµ(dg)<∞. Then, there exist0< ω1, ω2<1 andC >0, such that

δp(n)≤C(β(ω1n) +ω2n) ∀n≥1. (16)

(7)

Proof. Letε >0, and letun(x, y) :=h(εn+1, Wn,x)−h(εn+1, Wn,y). Writing,

upn(x, y) =upn(x, y)1{d(Wn,x,Wn,y)ρεn}+upn(x, y)1{d(Wn,x,Wn,y)>ρεn}, (17) we obtain the upper bound

ZZ

E(upn(x, y))ν(dx)ν(dy)≤In+IIn. Clearly,

In≤(β(ρεn))p Z

G

(˜η(g))pµ(dg) Z

X

(χ(x))ptν(dx)

. Moreover, using H¨older’s inequality and Lemma 4, we have

IIn

≤ 2p1 ρε(αsp)n

Z

G

ηp(g)µ(dg) ZZ

E

sp(Wn,x) +χsp(Wn,y)) (d(Wn,x, Wn,y))αsp

ν(dx)ν(dy)

≤ C

ρε(αsp)n ZZ

E(d(Wn,x, Wn,y))αν(dx)ν(dy)

1sp/α

≤C˜

(ρ(α))1/α ρε

sp)n

.

The desired bound follows by takingεsmall enough.

Proposition 6 Let 1 < p < α. Assume that there exist C > 0 and ρ ∈ (0,1) such that (12) and (13) hold. Let 0 ≤ s < α/p and 0 ≤ t ≤ p/α. Let h satisty Hs,t. Assume that R

G(˜η(g))pµ(dg)<∞,R

G(η(g))pµ(dg)<∞ and that β(2n) =O(nγ), with γ > κ(p). Then, the conclusion of Theorem 3 holds.

Proof. Starting from Theorem 3 and Lemma 5, it is enough to prove that our assumption implies that, for any a ∈(0,1), β(an) = O(nγ). Too see this, we note that there exists an integerℓ ≥1 such that a ≤ 1/2. Letn ≥2ℓ. Notice that n/(2ℓ)≤ [n/ℓ]≤n/ℓ. Since β is non-decreasing, we have

β(an)≤β(2[n/ℓ])≤C([n/ℓ])γ ≤C(2ℓ/n)γ,

and the result follows.

3.2 Applications to dilating endomorphisms of the torus

LetA be anm×mmatrix with integral entries. Then,A induces a transformationθA of the m-dimensional torusTm:=Rm/Zm preserving the Haar measureλ.

Assume that A is dilating, i.e. that all its eigenvalues have modulus strictly greater than one. Let Γ be a system of representative of Zm/AZm. Then, θA admits a Perron-Frobenius operatorPA given by

PAf(x) = 1 N

X

γΓ

f(A1x+A1γ), (18) for every continuous functionf onTm, where N=|detA|= #Γ.

SincePA is markovian, there exists a Markov chain with state spaceTm admittingλas sta- tionary distribution. This Markov chain may be realized as follows: letW0be a random variable taking values inTm and (εi)i1 be iid variables uniformly distributed on Γ and independent of

(8)

W0. For everyn≥1, defineWn:=A1Wn1+A1εn. Denote by (Wn,x)n0the Markov chain starting atx∈Tm.

Let h be some measurable function from Tm to R, and let Xn = h(Wn) where W0 has distributionλ. Let alsoXn,x=h(Wn,x).

For everyp≥1 and everyf ∈Lp(λ) theLp-modulus of continuity off is given by ωp,f(δ) := sup

|x|≤δkf(·+x)−fkp ∀0≤δ≤1, where| · |stands for the euclidean norm.

Lemma 7 Let p≥1andh∈Lp(λ). The following upper bound holds:

ZZ

E(|Xn,x−Xn,y|p)λ(dx)λ(y) 1/p

≤2m/pωp,h ∆ An([0,1]m) ,

where∆ (An([0,1]m))stands for the diameter ofAn([0,1]m). Consequently (using (10)), the coefficientsδp(n)of the stationary sequence(Xn)nZ satisfy

δp(n)≤2m/p+1ωp,h ∆ An([0,1]m) .

Proof. We start by some preliminary considerations. Iterating the recursive equationWn,x= A1Wn1,x+A1εn, we get that

Wn,x=Anx+ Xn i=1

Aiεni+1.

Note that the random variableWn,x has the same distribution asY1,x, where Y1,x is the first iteration of the Markov chain starting at x with transition PAn = PAn. As explained at the beginning of this section, this may be realized as

Y1,x =Anx+Anξ1,

whereξ1is uniformly distributed over Γn (a system of representative ofZm/AnZm). LetZ1,x= h(Y1,x) It follows that

ZZ

E(|Xn,x−Xn,y|p)λ(dx)λ(y) = ZZ

E(|Z1,x−Z1,y|p)λ(dx)λ(y).

From this last equality, we see that it suffices to prove Lemma 7 forn= 1, the general case then follows by consideringAn rather thanA.

We refer to [2] for the results that we need about tiling. There exists a unique compact set K⊂Rm, such that

K=∪γΓ(A1K+A1γ) (19) and an integerq≥1 such that

X

nZm

1K+n=q λ-almost everywhere.

Moreover, for everyγ, γ ∈Γ withγ 6=γ, λ (A1K+A1γ)∩(A1K+A1γ)

= 0. Using that

1K = X

nZm

1(K([0,1]mn)= X

nZm

1((K+n)[0,1]m)n λ-almost everywhere,

(9)

we then infer that for everyZm-periodic locally integrable function gonRm, Z

K

g dλ= Z

Rm

X

nZm

1((K+n)[0,1]m)n

g dλ= X

nZm

Z

(K+n)[0,1]m

g dλ=q Z

Tm

g dλ . (20) Leth∈Lp(λ) (we identifyhwith aZm-periodic function onRm). We have

ZZ

E(|X1,x−X1,y|p)λ(dx)λ(y) = Z Z

E(|h(A1x+A1ε1)−h(A1y+A1ε1)|p)λ(dx)λ(dy)

= Z

Tm

Z

Tmx

E(|h(A1x+A1ε1)−h(A1(x+y) +A1ε1)|p)λ(dy)

λ(dx)

= Z

[1,1]m

Z

(Tmy)Tm

E(|h(A1x+A1ε1)−h(A1(x+y) +A1ε1)|p)λ(dx)

! λ(dy).

Set

ψy(x) :=E(|h(A1x+A1ε1)−h(A1(x+y) +A1ε1)|p)

= 1 N

X

γΓ

|h(A1x+A1γ)−h(A1(x+y) +A1γ)|p.

Notice thatψy isZm-periodic. Hence, using (20) and (19), we have Z

(Tmy)Tm

ψy(x)λ(dx)≤ Z

Tm

ψy(x)λ(dx) = 1 q

Z

K

ψy(x)λ(dx)

= 1 q

Z

γ∈Γ(A−1K+A−1γ)|h(x)−h(x+A1y)|pλ(dx) = 1 q

Z

K|h(x)−h(x+A1y)|pλ(dx)

= Z

Tm|h(x)−h(x+A1y)|pλ(dx),

and the result follows.

We shall now explain how to obtain the strong approximation result with rate o(n1/p) for the partial sums of the process (h◦θAn)nN for h∈ Lp(λ). Let (εn)n0 be a sequence of iid variables uniformly distributed on Γ. We define a probabilityν on Tm by setting, for every f ∈C([0,1]),

Z

Tm

f dν :=E

f

X

k0

Ak1εk

.

By construction,ν isPA-invariant. SinceAis dilating, the onlyPA-invariant probability onTm isλ.

DefineZ0:=P

k0Ak1εk and for everyn≥1, (with equality inTm) Zn:=AnZ0=X

k0

Ank1εk =X

kn

Ank1εk=X

k0

Ak1εk+n.

Notice that for anyh∈Lp(λ) the processes (h◦θAn)n0(under λ) and (Yn)nN:= (h(Zn))n0

(underP) have the same distribution.

Let ˜δp(n) be the coefficients associated with (Yn)n0 as in (5). The computations done in the proof of Lemma 7 yield to the following bound

(10)

δ˜p(n)≤2m/p+1ωp,h ∆ An([0,1]m)

. (21)

As a consequence of Lemma 7 and of (21), Theorem 1 (applied to (h(Zn))n0) or Theorem 3 (applied to (h(Wn))n0), lead to the following proposition:

Proposition 8 Letp >2and letκ(p)be defined in (3). Leth∈Lp(λ)be such thatωp,h(2n)≤ O(nγ) for some γ > κ(p). Assume that, with the above notations, Sn = X1+· · ·+Xn = h(W1) +· · ·+h(Wn), orSn=h(Z1) +· · ·+h(Zn). Thenn1E (Sn−nλ(h))2

→σ2 asn→ ∞ and for every (fixed) x∈[0,1], one can redefine (Sn)n1 without changing its distribution on a (richer) probability space on which there exist iid random variables (Ni)i1 with common distribution N(0, σ2), such that,

Sn−nλ(h)− Xn i=1

Ni

=o

n1/p P-a.s.

Remark 9 Alternatively, one can also apply Theorem 2 in [3], by using the upper bound on δ1(n)given in Lemma 7. For instance, ifhis bounded and such thatP

n>0np2ω1,h(2n)<∞, then the conclusion of Proposition 8 holds. If m = 1 (for instance for the transformation θ(x) = 2x−[2x]), this implies that, for BV-observables, the strong approximation holds with the rateo(n1/p)for anyp >2.

3.3 Applications to dilating piecewise affine maps

LetKbe a countable set, with|K| ≥2, and (Ik)kK be a collection of disjoint open subintervals of [0,1] such that S

kKIk = [0,1] (it is possible to have several accumulation points). Notice thatP

kKλ(Ik) = 1, whereλstands for the Lebesgue measure on [0,1].

LetT : [0,1]→[0,1] be a map such that T|Ik is affine and onto (0,1), so thatT|Ik extends in a trivial way to an affine map from Ik onto [0,1], that we still denote by T|Ik. The values of T on [0,1]\S

kKIk will be irrelevant in the sequel. For every k ∈ K, denote by sk the inverse ofT|Ik from (0,1) onto Ik. There exist realsαk andβk, such that for every x∈(0,1), sk(x) =αkx+βk (henceT|Ik(u) = (u−βk)/αk). Then,|αk|=λ(Ik).

Such a mapT admits a Perron-Frobenius operatorP defined by P f(x) := X

kK

k|f(αkx+βk), for every continuous functionf on [0,1].

Since P is Markovian and leavesλinvariant, there exists a Markov chain with state space [0,1] admittingλas stationary distribution. Since|K| ≥ 2 then 0<|αk|<1 for everyk∈K and one may easily prove thatλis the onlyP-invariant measure on [0,1].

The above Markov chain may be realized as follows. Let W0 be a random variable taking values in [0,1]. Let (εi)i1be iid random variables independent ofW0, taking values inK, such thatP(ε1 =k) =|αk|for everyk∈K. For everyn≥1, set Wn :=sεn(Wn1) and denote by (Wn,x)n0 the Markov chain starting from x ∈[0,1]. Notice that for every n ≥1 and every x∈[0,1],

Wn,x=sεn◦ · · · ◦sε1(0) +αεn. . . αε1x:=Bn+Anx .

Let h be some measurable function from [0,1] to R, and let Xn = h(Wn) where W0 has distributionλ. Let alsoXn,x=h(Wn,x).

For everyf ∈C([0,1]) define

ω,f(δ) = sup

x,y[0,1],|xy|≤δ|f(x)−f(y)|, ∀δ∈[0,1].

(11)

Define also

δ(n) := sup

x,y[0,1]

E|Xn,x−Xn,y|.

Lemma 10 Let h∈C([0,1]), and letα¯ := maxkKk|. For every integer n≥1, we have sup

x,y[0,1]|Xn,x−Xn,y| ≤2ω,h(¯αn).

In particular for everyn≥1,δp(n)≤2ω,h(¯αn)for any p≥1, andδ(n)≤2ω,h(¯αn).

Proof. For everyx∈[0,1], we have

|Xn,x−Xn,y|=|h(Bn+Anx)−h(Zn+Any)|

≤ |h(Bn+Anx)−h(Bn)|+|h(Bn+Any)−h(Bn)|,

and the result follows.

Remark 11 When K ={1, . . . , r} and|α1|=· · ·=|αr|, similar computations as those done in Section 3.2 allow to control(δp(n))nN thanks to theLp-modulus of continuity. Notice that, actually, the case whereα1=· · ·=αris included in section 3.2 (takingm= 1).

As in the previous subsection, let us also consider the process (h◦Tn)nN. Let (εn)nN

be iid random variables taking values in K such that P(ε1 =k) = |αk| for every k∈ K. For everyn∈N, setZn:=P

Nsεn◦ · · · ◦sεℓ+n(0) =P

N

Q1 j=0αεj+n

βεℓ+n. Then, (Zn)nNis identically distributed and the common law is invariant byP, so it is the Lebesgue measure on [0,1]. Moreover, one can see thatZn =TnZ0 for every n∈N. Hence, for every h∈C([0,1]), the processes (h◦Tn)nN (under λ) and (Yn)nN := (h(Zn))nN (under P) have the same distribution. As above the following upper bound clearly holds

˜δ(n)≤2ω,h(¯αn). (22)

Proposition 12 Letp >2,κ(p)be defined by (3), andh∈C([0,1]). LetSn=X1+· · ·+Xn= h(W1) +· · ·+h(Wn), orSn=h(Z1) +· · ·+h(Zn). Assume thatω,h(2n)≤O(nγ)for some γsuch that

γ > p−1

2 ifp∈(2,3], and γ > κ(p)ifp >3.

Then n1E (Sn−nλ(h))2

→σ2 as n→ ∞ and for every (fixed)x∈[0,1], one can redefine (Sn)n1 without changing its distribution on a (richer) probability space on which there exist iid random variables(Ni)i1 with common distribution N(0, σ2), such that,

Sn−nλ(h)− Xn i=1

Ni

=o

n1/p P-a.s.

Proof. It suffices to combine Lemma 10 (or (22)) with either our Theorem 3 (or our Theorem 1) forp >3 or Theorem 1 of [3] forp∈(2,3] (which applies also to processes as defined by (4)).

4 Proof of the results

As in [3], the proof is based on a general proposition that can be established by combining the arguments given in the paper by Berkes, Liu and Wu [1]. Let us now recall this proposition: it applies to a strictly stationary sequence (Xk)k1 of real-valued random variables inLp (p >2)

(12)

that can be well approximated by a sequence ofm-dependent random random variables, with the help of an auxiliary sequence of iid random variables (εi)i0. Let (Mk)k1 be a sequence of positive real numbers and define

ϕk(x) = (x∧Mk)∨(−Mk) andgk(x) =x−ϕk(x). (23) Then, define

Xk,jk(Xj)−Eϕk(Xj) and Wk,ℓ=

ℓ+3Xk−1

i=1+3k−1

Xk,i. (24)

Let now (mk)k1 be a non-decreasing sequence of positive integers such that mk =o(3k), as k→ ∞, and define

k,j =E ϕk(Xj)|εj, εj1, . . . , εjmk

−Eϕk(Xj) for anyj≥mk+ 1 and Wfk,ℓ =

ℓ+3Xk−1

i=1+3k−1

k,i. (25) Finally, setk0:= inf{k≥1 : mk ≤213k2} and define

νk =mk1

E(Wfk,m2 k) + 2E(fWk,mk(fWk,2mk−Wfk,mk)) . (26) Proposition 13 (Berkes, Liu and Wu [1]) Let p >2. Assume that we can find a sequence of positive reals (Mk)k1, a non-decreasing sequence of positive integers (mk)k1 such that mk=o(32k/pk1) ask→ ∞, in such a way that the following conditions are satisfied:

X

k1

3k(p1)/pE(|gk(X1)|)<∞, (27) there existsα≥1 such that

X

kk0

3αk/p

max

13k3k−1

Wk,ℓ−Wfk,ℓ

α

α

<∞, (28)

and there existsr∈]2,∞[ such that X

kk0

3k 3kr/pmk

E

1max3mk

fWk,ℓ

r

<∞. (29)

Assume in addition that

the series σ2= Var(X12) + 2X

i1

Cov(X1, Xi+1) converge, (30) and

3kk1/2−σ)2=o(32k/p(logk)1), ask→ ∞. (31) Then, one can redefine(Xn)n1 without changing its distribution on a (richer) probability space on which there exist iid random variables(Ni)i1with common distributionN(0, σ2), such that,

Sn−nE(X1)− Xn i=1

Ni

=o

n1/p

P-a.s. (32)

Theorem 3 is a consequence of the next proposition, whose proof follows from Proposition 13.

This proposition applies to the stationary sequence (Xn)n1 defined by (8) and the conditions are expressed in terms of the coefficient δp defined in the second part of Section 2. In what follows, all the proofs will be written with the help of that coefficient, but the arguments are exactly the same for the sequence defined by (1) or (4) and the coefficients ˜δp.

(13)

Proposition 14 Let p > 2. Assume that we can find a non-decreasing sequence of positive integers (mk)k1 such that mk = o(32k/pk1), as k → ∞, in such a way that the following conditions are satisfied:

X

kk0

3k(p2)/2

X

k

δ2(m)m1/2ℓ+1

p

<∞, X

kk0

X

k

δp(m)m1ℓ+11/p

p

<∞, (33) X

k

δ2(m)m1/2ℓ+1=o(3k(2p)/2p/p

logk), (34)

and there existsr∈]p,∞[, such that X

k0

3k(pr)/pm(rk2)/2 <∞, (35)

and X

j1

p(j))p/r

j1/r <∞. (36)

Then,(30)holds. Moreover, ifσ >0, one can redefine(Xn)n1without changing its distribution on a (richer) probability space on which there exist iid random variables (Ni)i1 with common distribution N(0, σ2), such that,

Sn−nE(X1)− Xn i=1

Ni

=o

n1/p

P-a.s. (37)

4.1 Proof of Proposition 14

We consider a process (Xn)n1 satisfying (8), with stationary distribution π. We shall check that the assumptions of Proposition 13 are satisfied.

Set V0 := (ε0, W0). It is not difficult to see thatkE((Xk−E(X1))|V0)k2 ≤δ2(k) ≤δp(k).

Now, sincer≥p≥2, it follows from (36) that X

n1

δ2(n)

√n <∞. (38)

Then, the fact that (30) holds follows from the fact that (see e.g. Lemma 22 of [3]) X

k1

|Cov(X1, Xk+1)| ≪

X

k0

(k+ 1)1/2kE((Xk−E(X1))|V0)k2

2

<∞. (39)

We choose Mk = 3k/p. Since the Xi’s are in Lp, it is easy to see that with this choice of Mk, condition (27) is satisfied (it suffices to write thatE(|gk(X1)|)≤E(|X1|1|X1|>Mk) and to use Fubini’s Theorem).

We shall check the condition (28) with α = p. To do so, we apply the Rosenthal-type inequality given in Proposition 15 of the appendix, to the process ( ˜Xk,ℓ+3k−1−Xk,ℓ+3k−1)1, with the choiceη0= (ε3k−1, . . . , ε1+3k−1mk, W3k−1mk) and for everyℓ≥1,ηℓ+3k−1. We have to bound, forq≥1, the coefficientsδq,k (n) used in Proposition 15. When 1≤n≤mk, we use the boundδq,k(n)≤2kX˜k,3k−1−Xk,3k−1kq ≤2δq(mk). When n > mk we notice that the contribution of ( ˜Xk,ℓ+3k−1)1 toδq,k (n) is null, so thatδq,k(n)≤δq(n).

In particular we infer that, fork≥k0,

(14)

max

13k3k−1|Wfk,ℓ−Wk,ℓ|

p

≪ 3k/2δ2(mk)m1/2k + 3k/pm1k1/pδp(mk) +

3k/2 X

jmk

δ2(j)

√j + 3k/p X

jmk

δp(j) j1/p

≪3k/2X

k

δ2(m)m1/2ℓ+1+ 3k/pX

k

δp(m)m1ℓ+11/p. Hence, (28) holds withα=p, since (33) is satisfied.

We prove now that (29) holds for some r >2. We apply again Proposition 15, but now to the process ( ˜Xk,ℓ+3k−1)13mk and with the choiceη0= (ε3k−1, . . . , ε1+3k−1mk) and for every ℓ≥1,ηℓ+3k−1.

For every q≥1, denote by δq,k(n) the nth coefficientδ associated with the above choice, and notice that δq,k (n) = 0 as soon as n > mk. For every q ≥ 1, denote by δq,k (n) thenth coefficient δ associated with the process (Xk,ℓ+3k−1)1. One can see that for every n ≥ 0, δq,k (n)≤δq,k (n).

For every r ≥ 2, everyk ≥1, with dk the unique integer such that 2dk1 <3mk ≤2dk, Proposition 15 gives

max

13mk

fWk,ℓ

r

≪2dk/2

mk

X

j=0

δ2,k (j)/(j+ 1)1/2+ 2dk/r

mk

X

j=0

δr,k (j)/(j+ 1)1/r. (40) Hence, (29) holds for somer >2, if

X

k0

3k(pr)/pm(rk2)/2<∞ and sup

k0

X

j0

δ2,k (j)/(j+ 1)1/2<∞, (41)

and X

k0

3k(pr)/p X

j0

δr,k (j)/(j+ 1)1/rr

<∞. (42)

The first part of (41) is exactly (35). Moreover since ϕk is 1-Lipschitz, we have δ2,k (n) ≤ δ2(n)≤δp(n). Hence the second part of (41) holds for somer >2 as soon as (36) does.

It remains to prove (42). By H¨older’s inequality, X

j0

δr,k (j) (j+ 1)1/r

X

j0

p(j))p/r (j+ 1)1/r

(r1)/r

X

j0

r,k(j))r (j+ 1)

(j+ 1)(r1)/rp(j))p(r1)/r)

1/r

.

Taking into account (36), we see that (42) holds as soon as X

j0

p(j))p(1r)/r (j+ 1)1/r

X

k0

3k(pr)/pr,k (j))r<∞.

Using the fact thatδr,k(j)≤2kϕk(X0)kr, that for every non negative random variableZ, X

k0

3k(pr)/pE((ϕk(Z))r)≤Cr,pE(Zp), and (36) again, we see that (42) holds.

Références

Documents relatifs

Pro@ Let us suppose that the hypotheses of the lemma are satisfied, so in particular E is an infinite sequence of solutions of (2.1).. Our next goal is to prove the

(éd.), De la mer du Nord à la Méditerranée : Francia Media, une région au cœur de l'Europe (c.. Actes du colloque international (Metz, Luxembourg, Trêves, 8-11

A promising long-term avenue can be the generation of more complex entangled states. Quantum state carving [197, 198] is one of these promising approaches. In this scheme, a

personne s’engage dans une occupation c’est aux conditions suivantes : se sentir efficace dans la réalisation, avoir de l’intérêt, trouver du sens et de la valeur et

Figure 9 shows that the SI images can resolve the γ’ precipitates down to 150 nm, which means all the precipitates of the studied microstructure including the smallest ones

The main purpose of this paper is to investigate the strong approximation of the p-fold integrated empirical process, p being a fixed positive integer. More precisely, we obtain

Even if the residual bath is not strictly Markovian or weakly coupled to the redefined system, one might hope that the ME including the RC provides nevertheless a convenient way

For instance, when deal- ing with his market-clearing model with government debt, Modigliani (1963, 87) states that “an increase in the money supply by increasing P and reducing