• Aucun résultat trouvé

On the coupling property of Lévy processes

N/A
N/A
Protected

Academic year: 2022

Partager "On the coupling property of Lévy processes"

Copied!
13
0
0

Texte intégral

(1)

www.imstat.org/aihp 2011, Vol. 47, No. 4, 1147–1159

DOI:10.1214/10-AIHP400

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

On the coupling property of Lévy processes

René L. Schilling

b

and Jian Wang

a,b

aSchool of Mathematics and Computer Science, Fujian Normal University, 350007, Fuzhou, P. R. China. E-mail:jianwang@fjnu.edu.cn bInstitut für Mathematische Stochastik, TU Dresden, 01062 Dresden, Germany. E-mail:rene.schilling@tu-dresden.de

Received 21 July 2010; revised 25 October 2010; accepted 2 November 2010

Abstract. We give necessary and sufficient conditions guaranteeing that the coupling for Lévy processes (with non-degenerate jump part) is successful. Our method relies on explicit formulae for the transition semigroup of a compound Poisson process and earlier results by Mineka and Lindvall–Rogers on couplings of random walks. In particular, we obtain that a Lévy process admits a successful coupling, if it is a strong Feller process or if the Lévy (jump) measure has an absolutely continuous component.

Résumé. Nous donnons les conditions nécessaires et suffisantes pour le succès du couplage entre des processus de Lévy (avec partie de sauts non-dégénérée). Notre méthode est basée sur les formules explicites pour le semigroupe de transition d’un processus de Poisson composé, et les résultats de Mineka et Lindvall–Rogers sur le couplage d’une marche aléatoire. En particulier, nous montrons qu’un processus de Lévy admet un couplage, s’il est un processus fortement fellerien ou si la mesure de Lévy (mesure de sauts) possède une composante absolument continue.

MSC:60G51; 60G50; 60J25; 60J75

Keywords:Coupling property; Lévy processes; Compound Poisson processes; Random walks; Mineka coupling; Strong Feller property

1. Introduction and main results

The coupling method is a powerful tool in the study of Markov processes and interacting particle systems. There are some comprehensive books on this topic now, see e.g. [2,7,13,15]. Let (Xt)t0 be a Markov process on Rd with transition probability function{Pt(x,·)}t0,x∈Rd. AnR2d-valued process(Xt, Xt)t0is calleda coupling of the Markov process(Xt)t0, if both(Xt)t0and(Xt)t0are Markov processes which have the same transition functions Pt(x,·)but possibly different initial distributions. In this case,(Xt)t0and(Xt)t0are called themarginal processes of the coupling process; the coupling time is defined byT :=inf{t≥0: Xt=Xt}. The coupling(Xt, Xt)t0is said to besuccessfulifT is a.s. finite. A Markov process(Xt)t0admits asuccessful coupling(also: enjoys thecoupling property) if for any two initial distributionsμ1andμ2, there exists a successful coupling with marginal processes possessing the same transition probability functionsPt(x,·)and starting fromμ1andμ2, respectively. It is known, see [7,12], that the coupling property is equivalent to the statement that

tlim→∞μ1Ptμ2PtVar=0 forμ1andμ2∈P Rd

. (1.1)

As usual,μP (A)=

P (x, A)μ(dx)is the left action of the semigroup and · Varstands for the total variation norm.

If a Markov process admits a successful coupling, then it also has the Liouville property, i.e. every bounded harmonic function is constant; in this context a function f is harmonic, if Lf =0 where L is the generator of the Markov process. See [3,4] and references therein for this result and more details on the coupling property.

The aim of this paper is to study the coupling property of Lévy processes by using explicit conditions on Lévy measures. Our work is mainly motivated by the recent paper [14], which contains some interesting results on the

(2)

coupling property of Ornstein–Uhlenbeck processes; the paper [14] uses mainly the conditional Girsanov theorem on Poisson space and assumes that the corresponding Lévy measure has a non-trivial absolutely continuous part. Our technique here is completely different from F.-Y. Wang’s paper [14]. We use an explicit expression of the compound Poisson semigroup and combine this with the Mineka and Lindvall–Rogers couplings for random walks, see [8].

A Lévy process(Xt)t0onRdis a stochastic process with stationary and independent increments and càdlàg (right continuous with finite left limits) paths. It is well known thatXt is a (strong) Markov process whose infinitesimal generator is, forfCb(Rd), of the form

Lf (x)=1 2

d i,j=1

qi,j2f (x)

∂xi∂xj + f (x+z)f (x)z· ∇f (x) 1+ |z|2

ν(dz)+b· ∇f (x),

whereQ=(qi,j)di,j=1is a positive semi-definite matrix,b∈Rdis the drift vector andνis the Lévy or jump measure;

the Lévy measureνis aσ-finite measure on(Rd,B(Rd))such that

(1∧ |z|2)ν(dz) <∞. Note that the Lévy triplet (b, Q, ν)characterizes, up to indistinguishability, the process (Xt)t0 uniquely. Our standard reference for Lévy processes is the monograph [11]. We writePt(x, A)=Pt(Ax),A∈B(Rd), for the transition probability ofXt.

Letμandν be two bounded measures on(Rd,B(Rd)). We defineμν:=μν)+, whereν)±is the Jordan–Hahn decomposition of the signed measureμν. In particular,μν=νμand it is easy to see that μν(Rd)=12[μ(Rd)+ν(Rd)μνVar],cf. [2]. We can now state our main result.

Theorem 1.1. Let(Xt)t0be ad-dimensional Lévy process with Lévy triplet(b, Q, ν).For everyε >0,defineνεby

νε(B)= ν(B), ν

Rd

<∞;

ν

zB: |z| ≥ε , ν

Rd

= ∞. (1.2)

Assume that there existε, δ >0such that inf

x∈Rd,|x|≤δ

νεxνε) Rd

>0. (1.3)

Then,there exists a constantC=C(ε, δ, ν) >0such that for allx, y∈Rdandt >0, Pt(x,·)Pt(y,·)

VarC(1+ |xy|)

t ∧2.

In particular,the Lévy processXtadmits a successful coupling.

Remark 1.2. Condition(1.3)guarantees that Pt(x,·)Pt(y,·)

Var=O t1/2

ast→ ∞

holds locally uniformly for allx, y∈Rd.This order of convergence is known to be optimal for compound Poisson processes,see[14],Remark3.1.In[14]it is pointed out that a pure jump Lévy process admits a successful coupling only if the Lévy measure has a non-discrete support,in order to make the process more active.Condition(1.3)is one possibility to guarantees sufficient jump activity;intuitively it will hold if for sufficiently small values ofε, δ >0and allx∈Rdwith|x| ≤δwe havex+supp(νε)∩supp(νε)=∅;heresupp(νε)is the support of the measureνε.

In order to see that(1.3)is sharp,we consider an one-dimensional compound Poisson process with Lévy measure ν supported on Z.Then,for anyδ(0,1)and x∈Rd with|x| ≤δ, νxν)(Z)=0. On the other hand,all functions satisfying f (x+n)=f (x)for x∈Rd and n∈Z are harmonic.By[3,4],this process cannot have the coupling property.

Theorem 3.1 of [14] establishes the coupling property for Lévy processes whose jump measureνhas a non-trivial absolutely continuous part. It seems to us that this condition is not directly comparable with (1.3). In fact, based on the

(3)

Lindvall–Rogers ‘zero–two law’ for random walks [8], Propsotion 1, we give in Section4a necessary and sufficient condition guaranteeing that a Lévy process has the coupling property. In this section we will also find the connection between (1.3) and the existence of a non-trivial absolutely continuous component of the Lévy measure. In particular, we obtain some extensions of Theorem1.1and [14], Theorem 3.1.

Once we know that a Lévy process admits the coupling property, many interesting new questions arise which are, however, beyond the scope of the present paper. For example, it would be interesting to construct explicitly the corresponding successful Markov coupling process and to determine its infinitesimal operator. There are a number of applications of optimal Markov processes and operators; we refer to [2,15] for background material and a more detailed account on diffusions and q-processes. We will discuss those topics for Lévy processes in a forthcoming paper [1].

2. The coupling property of compound Poisson processes

In this section, we consider the coupling property of compound Poisson processes. Let(Lt)t0be a compound Poisson process onRdsuch thatL0=xand with Lévy measureν. Then,Lt can be written as

Lt=x+

Nt

i=1

ξi, t≥0,

whereNt is the Poisson process with rateλ:=ν(Rd)andi)i1is a sequence of i.i.d. random variables onRdwith distributionν(·)/λ; moreover, we assume that theξi’s are independent of Nt. As usual we use the convention that 0

i=1ξi=0.

The transition semigroup for a compound Poisson process is explicitly known. This allows us to reduce the coupling problem for a compound Poisson processes to that of a random walk. LetPtandLbe the semigroup and the generator forLt, respectively. Then, it is well known that for anyfBb(Rd),

Lf (x)= f (x+z)f (x) ν(dz)

=λ f (x+z)f (x) ν0(dz)

and

Pt=et L=

n=0

tnLn n! =eλt

n=0

(t λ)nν0n

n! , t≥0; (2.1)

hereν0nis then-fold convolution ofν0andν00:=δ0.

The following result explains the relationship of transition probabilities of compound Poisson processes and of random walks.

Proposition 2.1. Let Pt(x,·)be the transition probability of the compound Poisson process L=(Lt)t0 and let S=(Sn)n1,Sn=ξ1+· · ·+ξn,be a random walk where(ξi)i1are i.i.d.random variables withξ1ν0:=ν/ν(Rd).

Then,for allx, y∈Rd Pt(x,·)Pt(y,·)

Var≤eλt n=0

(λt )nP(x+Sn∈ ·)−P(y+Sn∈ ·)Var

n!

=eλt

2(1−δx,y)+

n=1

(λt )nP(x+Sn∈ ·)−P(y+Sn∈ ·)Var

n!

,

(4)

whereδx,yis a Kronecker delta function,i.e.δx,y=1ifx=y,and0otherwise.

Proof. LetPt andPnS be the semigroups of the compound Poisson processLand the random walkS, respectively.

Because of (2.1) we find Pt(x,·)Pt(y,·)

Var= sup

f1

Ptf (x)Ptf (y)

=eλt sup

f1

n=0

(λt )nxν0n(f )δyν0n(f )) n!

≤eλt sup

f1

n=0

(λt )n(PnSf (x)PnSf (y)) n!

≤eλt n=0

(λt )nsupf1|PnSf (x)PnSf (y)| n!

=eλt n=0

(λt )nP(x+Sn∈ ·)−P(y+Sn∈ ·)Var

n! ,

which proves the first assertion.

Forn=0 we haveS0=0; thus P(x+S0∈ ·)−P(y+S0∈ ·)

Var= δxδyVar=2(1−δx,y)

and the second assertion follows.

An immediate of Proposition2.1is the following estimate forPt(x,·)Pt(y,·)Varwhich is based on a similar inequality forP(x+Sn∈ ·)−P(y+Sn∈ ·)Var.

Proposition 2.2. Assume that for allx, y∈Rd there is a constantC(x, y) >0such that P(x+Sn∈ ·)−P(y+Sn∈ ·)

VarC(x, y)

n forn≥1. (2.2)

Then,

Pt(x,·)Pt(y,·)

Var≤2eλt(1δx,y)+

√2C(x, y)(1−eλt)

λt .

Proof. A combination of Proposition2.1and (2.2) yields for allx, y∈Rd Pt(x,·)Pt(y,·)

Var≤eλt

2(1−δx,y)+C(x, y) n=1

(λt )n

nn!

.

Jensen’s inequality for the concave functionx1/2gives

n=1

1 n

(λt )n

n! ≤ eλt−1 (eλt−1)1/2

n=1

(λt )n n·n!

1/2

=

eλt−1 λt

1/2

n=1

(λt )n+1 (n+1)!·n+1

n 1/2

(5)

eλt−1 λt

1/2√ 2

eλt−1−λt1/2

√2(eλt−1)

λt .

The required assertion follows form the estimates above.

We will now show thatLhas the coupling property wheneverShas.

Proposition 2.3. Let(Lt)t0be the compound Poisson process with Lévy measureν,and let(Sn)n0,be a random walk whereSn=S0+ξ1+ · · · +ξn,and(ξi)i1are i.i.d.random variables withξ1ν0:=ν/ν(Rd).IfSnadmits a successful coupling,so doesLt.

Proof. Forx, y∈Rd, letLtbe a compound Poisson process starting fromx∈Rd. ThenLt=Nt

i=0ξi, whereξ0=x and wherei)i1are i.i.d. random variables with common distributionν0:=ν/ν(Rd);(Nt)t0is a Poisson process with rateλ:=ν(Rd). Moreover,(Nt)t0andi)i1are independent.

Set S0=x and Sn=n

i=0ξi for n≥1. Since S has the coupling property, there exists another random walk Sn =n

i=0ξi such thatSS0andSS0 have the same law and such that for any starting pointξ0=y ofSthe coupling time

Tx,yS :=inf

k≥1: Sk=Sk

is a.s. finite.

Without loss of generality, we can assume thatSk=Sk forkTx,yS , and thati)i1is independent of(Nt)t0. Define

Lt=

Nt

i=0

ξi, t≥0.

ThenL=(Lt)t0is also a compound Poisson process with Lévy measureν and starting pointL0=y. In order to show thatLhas the coupling property, it is enough to verify that

Tx,yL :=inf

t >0:Lt=Lt

<. (2.3)

We claim that

Tx,yL =Kx,y withKx,y:=inf

t >0:NtTx,yS

. (2.4)

This implies (2.3). By assumption we know that for almost everyω,Tx,yS (ω) <∞. Since the Poisson processNttends to infinity ast→ ∞, there existsτ0(ω) <∞such thatNtTx,yS (ω)for alltτ0(ω). Therefore, (2.4) tells us that Tx,yLτ0<∞.

Let us finally prove (2.4). For this argument we assume thatω is fixed. Let t >0 be such thatNtTx,yS , i.e.

tKx,y. SinceSk=Sk forkTx,yS ,SNt =SN

t and, by construction,Lt=Lt; thus,Tx,yLtand sincetKx,ywas arbitrary, we haveTx,yLKx,y. On the other hand, assume thatKx,y>0. Then, by the very definition ofKx,y, for any ε >0, there existstε>0 such thattε> Kx,yεandNtεTx,yS −1. Hence,SN=SN

, i.e.Ltε =Lt

ε. Therefore, Tx,yLtε> Kx,yε. Lettingε→0, we getTx,yLKx,y and the proof is complete.

Note that the proof of Proposition2.3already gives an estimate for the rate at which coupling occurs: for allt >0 P

Tx,yL > t

=P

Nt< Tx,yS

=eλt

1+ k=1

P

Tx,yS > k(λt )k k!

.

(6)

What remains to be done is to get estimates for the coupling time of the random walkS,P(Tx,yS > k), k≥1. This requires concrete coupling constructions forS; the most interesting random walk couplings rely on a suitable coupling of the steps ξj andξj of S andS, respectively. In the next section we will, therefore, consider the Mineka and Lindvall–Rogers couplings.

We close this section with two comments on Proposition2.3.

Remark 2.4. (1)Theorem4.3below will show,that the converse of Proposition2.3is also true: LetL=(Lt)t0be a compound Poisson process with Lévy measureν, andS=(Sn)n0,Sn=S0+ξ1+ · · · +ξn, be a random walk where i)i1are i.i.d. random variables withξ1ν0:=ν/ν(Rd). IfLhas the coupling property so doesS.

(2)Proposition2.3can be easily generalized to more general settings,see also[1].More precisely,let(Xt)t0be a Markov process onRdand let(St)t0be a subordinator(i.e.an increasing Lévy process)which is independent of Xt.IfXthas the coupling property and ifSttends to infinity ast→ ∞,then the subordinate processXSt also has the coupling property.

3. The Mineka and Lindvall–Rogers couplings – A review

LetS=(Sn)n1,Sn=ξ1+ · · · +ξn, be a random walk onRd with i.i.d. stepsi)i0such thatξ1ν0. The main result of this section is

Theorem 3.1. Suppose that for someδ >0, η0=η0(δ):= inf

x∈Rd,|x|≤δ

ν0xν0) Rd

>0. (3.1)

Then there exists a constantC:=C(δ, η0) >0such that for allx, y∈Rdandn≥1, P(x+Sn∈ ·)−P(y+Sn∈ ·)

VarC(1+ |xy|)

n . (3.2)

The proof of Theorem3.1is mainly based on Mineka’s coupling [9], see also [7], Chapter II, Section 14, pp. 44–

47, and the coupling argument of the zero–two law for random walks proved in [8], Proposition 1, by Lindvall and Rogers. These papers do not contain an estimate as explicit as (3.2). Therefore we decided to include a detailed proof on our own which again highlights the role of the sufficient condition (3.1).

We begin with an auxiliary result which describes the total variation norm of a signed measure under a non- degenerate linear transformation.

Lemma 3.2. Letμbe a probability measureμonRd.Then we have for allx, y∈Rd δxμδyμVar= δxyμμVar=δ|xy|e1

μRx1y

μRx1y

Var,

where e1=(1,0, . . . ,0)∈Rd and Ra is a non-degenerate rotation such thatRaa= |a|e1.In particular, for any a∈Rd,

δaμμVar= δaμμVar.

Proof. Using the definition of the total variation norm we get δxμδyμVar=2 sup

A∈B(Rd)

δxμ(A)δyμ(A)

=2 sup

A∈B(Rd)

μ(A−x)μ(Ay)

=2 sup

B∈B(Rd)

μ

B(xy)

μ(B)

= δxyμμVar.

(7)

Now leta∈Rdand denote byRathe rotation such thatRaa= |a|e1. Clearly, μRa(A)=μ

Rax∈Rd: xA

=μ

y∈Rd: Ra1(y)A

, A∈B Rd

.

Then,

δaμμVar=2 sup

AB(Rd)

μ(Aa)μ(A)

=2 sup

AB(Rd)

μRa1

Ra(Aa)

μRa1(RaA)

=2 sup

A∈B(Rd)

μ◦Ra1

RaA− |a|e1

μRa1(RaA)

=2 sup

B∈B(Rd)

μ◦Ra1

B− |a|e1

μRa1(B)

|a|e

1

μRa1

μRa1

Var.

Proposition 3.3. Under(3.1),there exists a constantC=C(η0) >0such that sup

|xy|≤δ

P(x+Sn∈ ·)−P(y+Sn∈ ·)

VarC

n. (3.3)

Proof. Step 1.It is easy to see that for anya∈Rd and any probability measureμ, μnRa1=

μRa1n

.

Lemma3.2shows that (3.3) is equivalent to the following estimate sup

|a|≤δ

P

|a|e1+Sa,n∈ ·

−P(Sa,n∈ ·)

VarC

n. (3.4)

Here,Sa,nis a random walk inRdwith i.i.d. stepsξa,1, ξa,2, . . .andξa,1ν0Ra1. On the other hand, Lemma3.2also shows that

ν0Ra1

δ|a|e1

ν0Ra1 Rd

=1−1

2ν0Ra1δ|a|e1

ν0Ra1

Var

=1−1

2ν0aν0)

Var

=ν0aν0) Rd

.

Therefore, (3.1) implies that for anya∈Rd

|ainf|≤δ

ν0Ra1

δ|a|e1

ν0Ra1 Rd

= inf

|a|≤δ

ν0aν0) Rd

>0. (3.5)

In order to simplify the notation, we useν:=ν0Ra1andSn:=Sa,n. With this notation (3.4) becomes

|supa|≤δ

P

|a|e1+Sn∈ ·

−P(Sn∈ ·)

VarC

n. (3.6)

(8)

Step 2.Assume that|a| ∈(0, δ]and setν|a|=δ|a|e1νandν−|a|=δ−|a|e1ν. Let(ξ, ξ )∈Rd×Rd be a pair of random variables with the following distribution

P

(ξ, ξ )A×D

=

⎧⎪

⎪⎩

1

2ν−|a|)(A), D=

|a|e1

;

1

2ν|a|)(A), D=

−|a|e1

; ν12ν−|a|+νν|a|)

(A), D= {0};

whereA∈B(Rd)andD∈ {{−|a|e1},{0},{|a|e1}}. We see from (3.5) that P

ξ= |a|e1

=1 2

ν−|a|e1ν) Rd

≥1 2 inf

|a|≤δν0aν0) Rd

.

By Lemma3.2, P

ξ= −|a|e1

=1 2

ν|a|e1ν) Rd

=1 2

ν−|a|e1ν) Rd

=P

ξ= |a|e1

.

It is clear that the distribution ofξ isν. Letξ=ξ+ξ. We claim that the distribution ofξis alsoν. Indeed, for anyA∈B(Rd),

P ξA

=P

ξ− |a|e1A, ξ= −|a|e1 +P

ξ+ |a|e1A, ξ= |a|e1

+PA, ξ=0)

=δ−|a|e1ν|a|)(A)

2 +δ|a|e1ν−|a|)(A)

2 +

ννν−|a|+νν|a| 2

(A)

=μ(A),

where we have used thatδ−|a|e1ν|a|)=νν−|a| andδ|a|e1ν−|a|)=νν|a|. Now we construct the coupling(Sn, Sn)ofSnwith the i.i.d. pairsi, ξi), i≥1, where1, ξ1)(ξ, ξ). Sinceξiξi=ξ, we know that ξiξiis, for alli≥1, symmetrically distributed, takes only the values−|a|e1, 0 and|a|e1. Because of (3.5), we have

P (a):=P

ξi1ξi1=0

=

ν−1

2ν−|a|+νν|a|) Rd

= 1−νν−|a| Rd

≤ 1− inf

|a|≤δν0aν0) Rd

=:γ (δ) <1,

whereξi =i1, ξi2, . . . , ξid) andξi =i1, ξi2, . . . , ξid). Set Sjn =n

i=1ξij andSnj=n

i=1ξij. We observe that (S1S1)is a random walk, whose step sizes are−|a|, 0 and|a|with probability12(1P (a)),P (a)and12(1P (a)), respectively. SinceSnj=Snjfor 2≤jd, we get

P

|a|e1+Sn∈ ·

−P(Sn∈ ·)

Var≤2P TS> n

, (3.7)

where

TS=inf

k≥1:Sk1=Sk1+ |a| .

(9)

Step 3.We will now estimate P(TS> n). Let V1, V2, . . .be i.i.d. symmetric random variables, whose common distribution is given by

P(Vi=x)=

⎧⎪

⎪⎩

1 2

1−P (a)

, x= −|a|;

1 2

1−P (a)

, x= |a|;

P (a), x=0.

DefineZn=n

i=1Vi. We have seen in Step 2 thatTS=inf{n≥1: Zn= |a|}. Then, by the reflection principle, P

TS> n

=P

maxknZk<|a|

≤2P

0≤Zn≤ |a| .

SinceZis the sum of i.i.d. random variables with mean 0 and varianceσ2= |a|2(1P (a)), we can use the central limit theorem to deduce, for sufficiently large values ofn,

P

TS> n

=2P

0≤ Zn

|a|√

1−P (a)

n≤ 1

√1−P (a)n

≤2P

0≤ Zn

|a|√

1−P (a)

n≤ 1

√1−γ (δ)n

C

√2π 1/

1γ (δ) n 0

ex2/2dx

C1,γ (δ)

n .

In the first inequality above we have used the fact that 1−P (a)≥1−γ (δ) >0.Therefore we find from (3.7) for all largen≥1

P

|a|e1+Sn∈ ·

−P(Sn∈ ·)

Var≤ 2C2,γ (δ)

n .

Since the right-hand side is bounded by 2, this estimate actually holds for alln≥1.

We can now use Lemma3.2to get P

±|a|e1+Sn∈ ·

−P(Sn∈ ·)

Var≤2C2,γ (δ)

n

which immediately yields (3.6), since|a| ≤δwas arbitrary.

We close this section with the proofs of Theorems3.1and1.1.

Proof of Theorem3.1. For anyx, y∈Rd, setk= [|xδy|] +1. Pickx0, x1, . . . , xk∈Rdsuch thatx0=x,xk=yand

|xixi1| ≤δfor 1≤ik. By Proposition3.3, P(x+Sn∈ ·)−P(y+Sn∈ ·)

Vark i=1

P(xi+Sn∈ ·)−P(xi1+Sn∈ ·)

Var

C(δ, η0)(1+ |xy|)

n ,

which is what we claimed.

(10)

Proof of Theorem1.1. Step 1.Assume first that the Lévy triplet is of the form(0,0, ν)and that the Lévy measure satisfiesλ=ν(Rd) <∞. This means thatXt is compound Poisson process. We use the notations from Section2. For allx∈Rd,

νxν) Rd

=λ

ν0xν0) Rd

>0.

Therefore, we can apply Proposition2.2and Theorem3.1to get Theorem1.1in this case.

Step 2.If(Xt)t0is a general Lévy process with Lévy triplet(b, Q, ν), we splitXt into two independent parts Xt=Xt+Xt,

whereXt is the compound Poisson process with Lévy triplet(0,0, νε)νεis defined in (1.2) – andXt is the Lévy process with triplet(b, Q, ννε). Denote byPt, Pt,Pt(x,dy), Pt(x,dy)the transition semigroups and transition functions of the processesXt andXt, respectively. Then,Pt=PtPt. Observe thatPt is a contraction semigroup, i.e.Ptf≤1 wheneverf≤1. Therefore,

Pt(x,·)Pt(y,·)

Var= sup

f1

Ptf (x)Ptf (y)

= sup

f1

PtPtf (x)PtPtf (y)

≤ sup

h1

Pth(x)Pth(y)

=Pt(x,·)Pt(y,·)

Var,

which reduces the general case to the compound Poisson setting considered in the first part.

4. Extensions: The Lindvall–Rogers ‘zero–two’ law

Motivated by Lindvall–Rogers’s zero–two law for random walks [8], Propsotion 1, we present a necessary and suf- ficient condition for the coupling property of a Lévy process. We will add a few simple sufficient criteria in terms of the Lévy measure which are easy to verify. Throughout this section we assume that(Xt)t0 is ad-dimensional Lévy process with Lévy measureν≡0; as usual,X0=0. ByPt(x,·)andPt we denote the transition probability and transition semigroup, respectively.

Theorem 4.1 (Criterion for successful couplings). The following statements are equivalent:

(1) The Lévy process(Xt)t0has the coupling property.

(2) There existst0>0 such that for any tt0, the transition probability Pt(x,·)has (with respect to Lebesgue measure)an absolutely continuous component.

In either case,for everyx, y∈Rd,there exists a constantC(x, y) >0such that Pt(x,·)Pt(y,·)

VarC(x, y)

t , t >0. (4.1)

If the Lévy process has the strong Feller property, i.e. the corresponding semigroup mapsBb(Rd)intoCb(Rd), Theorem4.1becomes particularly simple.

Corollary 4.2. Suppose that there exists somet0>0such that the semigroupPt0mapsBb(Rd)intoCb(Rd).Then,the Lévy processXt has the coupling property.In particular,every Lévy process which enjoys the strong Feller property has the coupling property.

(11)

Proof. By assumptionPt0 is a convolution operator which mapsBb(Rd)intoCb(Rd). Due to a result by Hawkes, cf.

[5] or [6], Lemma 4.8.20,Pt0 and allPt withtt0are of the formPt(x)=ptf (x), wherept(x)is the transition density of the process. Therefore condition (2) of Theorem4.1is satisfied.

Now, we turn to the proof of Theorem4.1.

Proof of Theorem4.1. As mentioned in Section1, the coupling property is equivalent to (1.1). Observe that μ1Ptμ2PtVarμ1PtPtVar+ μ2PtPtVar

δxPtδ0PtVarμ1(dx)+

δxPtδ0PtVarμ2(dx).

Thus, if

tlim→∞δxPtδ0PtVar=0 forx∈Rd, (4.2)

then we can use the dominated convergence theorem to see that (1.1) holds. Therefore, the assertions (1.1) and (4.2) are equivalent.

SinceδxPtδ0PtVaris decreasing int, we see that (4.2) is also equivalent to

nlim→∞δxPnδ0PnVar=0 forx∈Rd. (4.3)

Letξ1, ξ2, . . .be i.i.d. random variables onRdwithξ1μ1:=P(X1∈ ·)and setSn=n

i=1ξiforn≥1 andS0=0.

Since the increments of a Lévy process are independent and stationary, (4.3) is the same as

nlim→∞P(x+Sn∈ ·)−P(Sn∈ ·)

Var=0 forx∈Rd. (4.4)

According to [8], Remark (ii), pp. 124–125, or [10], Chapter 3, Section 3, Theorem 3.9, (4.4) holds if, and only if, μ1is spread out, i.e. for somem≥1,μ1m=Pm(0,·):=P(Xm∈ ·)has an absolutely continuous component. Since the semigroup of a Lévy process is a convolution semigroup, it is easy to see that for every tm, the transition probabilityPt(x,·)has an absolutely continuous part. Combining all the assertions above, we have proved that the statements (1) and (2) are equivalent.

Moreover, the arguments used in the proof of Proposition3.3together with [8], Proposition 1, show that P(x+Sn∈ ·)−P(Sn∈ ·)

Var=O n1/2

asn→ ∞

whenever the random walkSnhas the coupling property. Therefore, (4.1) follows from the arguments used in the first part of the proof, in particular sincetδxPtδ0PtVaris decreasing and

P(x+Sn∈ ·)−P(Sn∈ ·)

Var= δxPnδ0PnVar.

Let us finally derive some sufficient conditions in terms of the Lévy measure, which extend Theorem1.1and [14], Theorem 3.1.

Theorem 4.3 (Sufficient criteria for successful couplings). Let(Xt)t0be a Lévy process onRdwith Lévy measure ν≡0and defineε >0as in(1.2),i.e.forB∈B(Rd)

νε(B)= ν(B), ν

Rd

<∞;

ν

zB: |z| ≥ε , ν

Rd

= ∞.

If there exists someε >0such that one of the following conditions is satisfied (1) For somel≥1,νεlhas an absolutely continuous component.

Références

Documents relatifs

Lévy process, supremum process, reflected process, Itô measure, entrance law, stable process, subordinate Brownian motion, integral

The L 2 -regularity property is established here by using the asymptotic behavior of the density of the process solution of (2.1) in small time as well as its derivative with respect

Ces processus, not´ es SP pour Switching Processes, sont construits ` a partir d’un processus dit intrins` eque qui remplace le flot d´ eterministe de la construction des PDMP..

Keywords: Bartlett’s formula, continuous time moving average process, estimation of the Hurst index, fractional L´evy process, L´evy process, limit theorem, sample

We study the smooth-fit property of the American put price with finite maturity in an exponential L´evy model when the underlying stock pays dividends at a continuous rate.. As in

The L 2 - regularity property is established here by using the asymptotic behavior of the density of the process solution of (2.1) in small time as well as its derivative with

This work characterizes the law of the pair U t formed by the random variable X t and the running supremum X t ∗ of X at time t, with a valued-measure partial dierential equation

Together with known results from the fluctuation theory of spectrally negative L´ evy processes and control theory we apply these results in Section 5 to investigate the optimality