• Aucun résultat trouvé

From persistent random walks to the telegraph noise

N/A
N/A
Protected

Academic year: 2021

Partager "From persistent random walks to the telegraph noise"

Copied!
27
0
0

Texte intégral

(1)

HAL Id: hal-00326521

https://hal.archives-ouvertes.fr/hal-00326521

Submitted on 3 Oct 2008

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 abroad, or from public or private research centers.

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 publics ou privés.

From persistent random walks to the telegraph noise

Samuel Herrmann, Pierre Vallois

To cite this version:

Samuel Herrmann, Pierre Vallois. From persistent random walks to the telegraph noise. Stochastics and Dynamics, World Scientific Publishing, 2010, 10 (2), pp.161-196. �10.1142/S0219493710002905�.

�hal-00326521�

(2)

From persistent random walks to the telegraph noise

Samuel Herrmann and Pierre Vallois

Institut de Math´ematiques Elie Cartan - UMR 7502 Nancy-Universit´e, CNRS, INRIA

B.P. 239, 54506 Vandoeuvre-l`es-Nancy Cedex, France {herrmann,vallois}@iecn.u-nancy.fr

October 3, 2008

Abstract

We study a family of memory-based persistent random walks and we prove weak conver- gences after space-time rescaling. The limit processes are not only Brownian motions with drift. We have obtained a continuous but non-Markov process (Zt) which can be easely ex- pressed in terms of a counting process (Nt). In a particular case the counting process is a Poisson process, and (Zt) permits to represent the solution of the telegraph equation. We study in detail the Markov process ((Zt, Nt); t≥0).

1 The setting of persistent random walks.

1) The simplest way to present and define a persistent random walk with value in Z is to introduce the process of its increments (Yt, t∈IN). In the classical symmetric random walk case, this process is just a sequence of independent random variables satisfying IP(Yt= 1) = IP(Yt =−1) = 12 for any t≥0. Here we shall introduce some short range memory in these increments in order to create the persistence phenomenon. Namely (Yt) is a{−1,1}-valued Markov chain: the law of Yt+1given Ft=σ(Y0, Y1, . . . , Yt) depends only on the value ofYt. This dependence is represented by the transition probability π(x, y) = IP(Yt+1=y|Yt =x) with (x, y)∈ {−1,1}2:

π=

1−α α β 1−β

0< α <1, 0< β <1.

The persistent random walk is the corresponding process of partial sums:

Xt=

t

X

i=0

Yi with X0 =Y0= 1 or −1. (1.1)

Let us discuss two particular cases:

• If α+β= 1, then increments are independent and therefore the short range memory disappears. (Xt, t∈IN) is a classical Bernoulli random walk.

• The symmetric caseα=βwas historically suggested by F¨urth [7] and precisely defined by Taylor [14]. Goldstein [8] developed the calculation of the random walk law and clarified the link between this process and the so-called telegraph equation. Some nice presentation of these results can be found in Weiss’ book [17] and [18]. This partic- ular short memory process is often called either persistent or correlated random walk or Kac walks (see, for instance, [5]). An interesting presentation of different limiting distributions for this correlated random walk has been given by Renshaw and Henderson [11].

2) Recently, Vallois and Tapiero [15] studied the influence of the persistence phenomenon on the first and second moments of a counting process whose increments takes their values in

(3)

{0,1} instead of{−1,1}. They obtained some nearly linear behaviour for the expectation.

Using the transformationy→2y−1, it is easy to deduce that, in our setting, we have:

IE−1[Xt] := IE[Xt|X0=Y0=−1] = α−β

1−ρ (t+ 1)− 2α

(1−ρ)2(1−ρt+1). (1.2) IE+1[Xt] := IE[Xt|X0 =Y0= +1] = α−β

1−ρ (t+ 1)− 2β

(1−ρ)2(1−ρt+1). (1.3) An application to insurance has been given in [16].

It is actually possible to determine the moment generating function (see Proposition 6.4 in Section 6).

Φ(λ, t) = IE[λXt], (λ∈R+).

However it seems difficult to invert this transformation; i.e. to give the law ofXt.

3) This leads us to investigate limit distributions. It is well-known that the correctly normal- ized symmetric random walk converges towards the Brownian motion. Let us define the time and space normalizations. Letα0 andβ0 denote two real numbers satisfying:

0≤α0≤1, 0≤β0≤1. (1.4)

Let ∆xbe a positive small parameter so that:

0≤α0+c0x≤1, 0≤β0+c1x≤1, (1.5) wherec0 andc1 belong toR(see in subsection 6.2 the allowed range of parameters).

Let (Yt, t∈IN) be a Markov chain whose transition probabilities are given by the matrix:

π=

1−α0−c0x α0+c0x

β0+c1x 1−β0−c1x

. (1.6)

Let (Xt, t∈IN) be the random walk associated with (Yt) (cf. (1.1)). Define the normalized random walk (Zs, s∈∆tIN) by the relation:

Zs= ∆xXs/∆t, (∆t>0, ∆x>0). (1.7) Set ( ˜Zs, s≥0) the continuous time process obtained by linear interpolation of (Zs).

We introduce two essential parameters:

ρ0= 1−α0−β0 (the asymmetry coefficient), (1.8)

η00−α0. (1.9)

In this paper, we will aim at showing the existence of a normalization (i.e. to express ∆t in terms of ∆x) which depends onα00, so that ( ˜Zs) converges in distribution, as ∆x→0.

Our main results and the organization of the paper will be given in Section 2.

2 The main results

2.1 Case : ρ

0

= 1

Obviously ρ0 = 1 implies thatα00 = 0, and the transition probabilities matrix is given as

π=

1−c0x c0x

c1x 1−c1x

(c0, c1>0).

In order to describe the limiting process, we introduce a sequence of independent identically exponentially distributed random variables (en, n ≥ 1) with IE[en] = 1. We construct the following counting process:

Ntc0,c1=X

k≥1

11e12e2+...+λkek≤t}, (2.1)

(4)

where

λk=

1/c0 ifkis odd

1/c1 otherwise. (2.2)

Finally we define

Ztc0,c1= Z t

0

(−1)Nuc0,c1du. (2.3)

For simplicity of notations, in the symmetric case (i.e. c0 =c1), Ntc0 (resp. Ztc0) will stand for Ntc0,c0 (resp. Ztc0,c0). The process (Ztc0) has been introduced by Stroock (in [13] p. 37).

It is possible to show that if we rescale (Ztc0), this process converges in distribution to the standard Brownian motion. This property has been widely generalized. For instance Bardina and Jolis [1] have given weak approximation of the Brownian sheet from a Poisson process in the plane.

Theorem 2.1. Let ∆x = ∆t andY0 =X0 =−1. Then the interpolated persistent random walk( ˜Zs, s≥0)converges in distribution, as∆x→0, to the process (−Zsc0,c1, s≥0).

In particular ifc0=c1, then(Nuc0) is the Poisson process with parameterc0.

If Y0 = X0 = 1 then the interpolated persistent random walk ( ˜Zs, s ≥ 0) converges in distribution, as∆x→0, to the process (Zsc1,c0, s≥0).

Proof. See Section 4.

Next, in Section 3, we investigate the process (Ztc0,c1, Ntc0,c1; t ≥0). In particular we prove that it is Markov, we determine its semigroup and the law of (Ztc0,c1, Ntc0,c1),tbeing fixed. This permits to prove, when c0 =c1, the well-known relation (cf. [18], [5], [8], [9]) between the solutions of the wave equation and the telegraph equation. For this reason the process (Ztc0,c1) will be called the integrated telegraph noise (ITN for short).

We emphasize that our approach based on stochastic processes gives a better understanding of analytical properties.

We will give in Section 5 below two extensions of Theorem 2.1 to the cases where (Yt) is 1) a Markov chain which takes its values in{y1, . . . , yk},

2) a Markov chain with order 2 and valued in{−1,1}.

2.2 Case : ρ

0

6= 1

In this case, the limit process is Markov. We shall prove two kind of convergence results. The first one corresponds to the law of large numbers and the second one looks like functional central limit theorem.

Recall that ( ˜Zt, t≥0) is the linear interpolation of (Zt) andρ0 (resp.η0) has been defined by (1.8) (resp. (1.9)).

Theorem 2.2. 1) Suppose thatr∆t= ∆xwithr >0. ThenZ˜tconverges to the deterministic limit−1−ρrtη00 when∆x→0.

2) Suppose that r∆t= ∆2x withr >0, then the process(ξt , t≥0)defined by ξt = ˜Zt+ t√

0

(1−ρ0)√

t

converges in distribution to the process (ξ0t, t≥0), as∆x→0, where ξt0= 2r −τ

1−ρ0

+ η0τ (1−ρ0)2

t+

s

r(1 +ρ0) 1−ρ0

1− η02 (1−ρ0)2

Wt, (2.4)

(Wt,t≥0) is a one-dimensional Brownian motion,τ= (c0+c1)/2andτ= (c1−c0)/2.

Proof. See Section 6.

Gruber and Schweizer have proved in [10] a weak convergence result for a large class of generalized correlated random walks. However these results and ours can be only compared in the case α00.

(5)

Note that

1− η02

(1−ρ0)2 = 0⇐⇒α0= 0 or β0 = 0.

Suppose for instance that α0= 0. Thenβ0, c0>0 and ξt = ˜Zt+ t√

√ r

t

and ξt0=2rc0

β0

t.

Obviously, the diffusion coefficient of (ξt0) can also cancel whenρ0=−1.

Sinceρ0=−1⇐⇒α00= 1, thenc0, c1<0 and ξt= ˜Zt andξ0t =−rτ t.

This shows that, in the symmetric case (i.e. c0=c1), we have ξ0t = 0. This means that the normalization is not the right one since the limit is null. Changing the rescaling we can obtain a non-trivial limit.

Proposition 2.3. Supposeα00 = 1,c0=c1<0andr∆t= ∆3x withr >0.

The interpolated persistent walk( ˜Zt, t≥0)converges in law, as∆x→0, to(√

−rc0Wt, t≥ 0) where(Wt)is a standard Brownian motion.

Proof. See subsection 6.3

2.3 Organization of the paper

The third section presents few properties of the process (Ztc0,c1, t≥0) which has been defined by (2.3). Theorem 2.1 will be proven in Section 4. Section 5 will be devoted to two extensions of Theorem 2.1. In subsection 6.1 we determine the generating function ofXt(recall thatXt

has been defined by (1.1)). This is the main tool which permits to prove Theorem 2.2 and Proposition 2.3 (see subsections 6.2 and 6.3).

3 Properties of the integrated telegraph noise

The aim of this section is to study the two dimensional process (Ztc0,c1, Ntc0,c1; t≥0) in- troduced in (2.2) and (2.3). In the particular symmetric case c0 =c1, the study is simpler since the process (Ntc0, t≥0) is a Poisson process with ratec0 (IE(Ntc0) =c0t) andN0c0 = 0.

However we shall study the general case.

First, we determine in Proposition 3.1 the conditional density ofZtc0,c1 givenNtc0,c1=n.

As a by product we obtain the distribution ofZtc0,c1 (see Proposition 3.3). Second, we prove in Proposition 3.5 that (Ztc0,c1, Ntc0,c1, t≥0) is Markov and we determine its semi-group. We conclude this section by showing that the solution of the telegraph equation can be expressed in terms of the associated wave equation and (Ztc0,c0)t≥0. For this reason, (Ztc0,c1)t≥0 will be called the integrated telegraph noise (ITN for short). Recall that:

τ=c0+c1

2 , τ =c1−c0

2 . (3.1)

Proposition 3.1. 1)IP(Ntc0,c1 = 0) =e−tc0 and givenNtc0,c1= 0, we have Ztc0,c1=t.

2) The counting process takes even values with probability:

IP(Ntc0,c1= 2k) = (c0c1)kαk(t)

22kk!(k−1)!e−τ t withαk(t) = Z t

−t

(t−z)k−1(t+z)keτ zdz, (3.2) and the conditional distribution ofZct0,c1 is given by

IP(Ztc0,c1∈dz|Ntc0,c1= 2k) = 1

αk(t)(t−z)k−1(t+z)keτ z1[−t,t](z) (k≥1). (3.3) 3) The counting process takes odd values with probability:

IP(Ntc0,c1= 2k+ 1) = ck+10 ck1α˜k(t)

22k+1(k!)2 e−τ t withα˜k(t) = Z t

−t

(t−z)k(t+z)keτ zdz, (3.4)

(6)

and the conditional distribution ofZct0,c1 is given by IP(Zct0,c1∈dz|Ntc0,c1= 2k+ 1) = 1

˜

αk(t)(t−z)k(t+z)keτ z1[−t,t](z) (k≥0). (3.5) Corollary 3.2. In the particular symmetric casec0=c1, the conditional density function of Ztc0 givenNtc0=nis the centered beta density, i.e.

forn= 2k, k∈N: fn(t, z) =χ2k(t+z)k(t−z)k−1

t2k 1[−t,t](z), (3.6) forn= 2k+ 1, k∈N: fn(t, z) =χ2k+1(t+z)k(t−z)k

t2k+1 1[−t,t](z), (3.7) with

χ2k+12k+2= 1

22k+1B(k+ 1, k+ 1) = (2k+ 1)!

22k+1(k!)2 (k≥0), (B is the beta function (first Euler function): B(r, s) = Γ(r)Γ(s)Γ(r+s)).

Proof of Proposition 3.1. Associated with n≥0 and a bounded continuous function f, we define

n(f) = IEh

f(Ztc0,c1)1{Nc0,c1

t =n}

i. a)Whenn= 0, we obtain

0(f) = IEh

f(Zct0,c1)1{t<λ1e1}

i . Ift < λ1e1, thenZtc0,c1=tand

0(f) =f(t) IP(t < λ1e1) =f(t)e−tc0. b)Whenn≥1, using (2.1) we obtain

n(f) = IEh

f(Ztc0,c1)11e1+...+λnen≤t<λ1e1+...+λn+1en+1}

i . Ifλ1e1+. . .+λnen≤t < λ1e1+. . .+λn+1en+1 then

Zct0,c1 =

Z λ1e1 0

(−1)0du+

Z λ1e12e2 λ1e1

(−1)du+. . .+

Z λ1e1+...+λnen λ1e1+...+λn−1en−1

(−1)n−1du +

Z t

λ1e1+...+λnen

(−1)ndu.

Hence

Ztc0,c11e1−λ2e23e3+. . .+ (−1)n−1λnen+ (−1)n(t−λ1e1−. . .−λnen). (3.8) c)Evaluation of ∆2k(f),k≥1.

We introduce two sequences of random variables associated with (en):

ξek=e2+. . .+e2k, ξko=e1+. . .+e2k−1, (k≥1). (3.9) By (3.8), (2.2) and (3.9) we obtain the simpler expression

2k(f) = IEh

f(t−2ξke/c1)1o

k/c0ek/c1≤t<ξok/c0ke/c1+e2k+1/c0}

i .

Note that from our assumptions, ξekko ande2k+1are independent r.v.’s,ξokandξke are both gamma distributed with parameterk. Consequently:

2k(f) = 1

((k−1)!)2 Z

Dt

exp{−c0(t−y/c0−x/c1)}f(t−2x/c1)e−x−yxk−1yk−1dx dy

= ck0e−tc0 k!(k−1)!

Z tc1 0

f(t−2x/c1)xk−1(t−x/c1)kexpnc0

c1 −1 xo

dx,

(7)

where Dt=R2+∩ {y/c0+x/c1≤t}. Using the change of variablez =t−2x/c1, we obtain x=c1t−z

2 ,t−x/c1=t+z2 and

2k(f) = (c0c1)k 2

e−(c0+c1)t/2 k!(k−1)!

Z t

−t

f(z) t−z 2

k−1t+z 2

k

exp{(c1−c0)z/2}dz. (3.10) Finally (3.10) and (3.1) imply (3.2) and (3.3).

d)Evaluation of ∆2k+1(f) fork≥0. The arguments are similar to those presented in part c).

On the eventξok+1/c0ke/c1≤t < ξok+1/c0ke/c1+e2k+2/c1, we have:Ztc0,c1= 2ξok+1/c0−t;

this implies

2k+1(f) = IEh 1o

k+1/c0ek/c1≤t}exp

−c1(t−ξk+1o /c0−ξek/c1)

f(2ξok+1/c0−t)i . Sinceξk+1o and ξek are independent and gamma distributed with parameterk+ 1 (resp. k), we get

2k+1(f) = ck+10 ck1

2(k!)2e−(c0+c1)t/2 Z t

−t

f(z) t−z 2

kt+z 2

k

expn

(c1−c0)z/2o

dz. (3.11)

This leads directly to (3.4) and (3.5).

Let us recall the definition of the modified Bessel functions:

Iν(ξ) = X

m≥0

(ξ/2)ν+2m m!Γ(ν+m+ 1). Proposition 3.3. The distribution ofZtc0,c1 is given by

IP(Ztc0,c1∈dx) =e−c0tδt(dx) +e−τ tf(t, x)1[−t,t](x), (3.12) where

f(t, x) =1 2

hr

c0c1(t+x) t−x I1

pc0c1(t2−x2) +c0I0

pc0c1(t2−x2)i

eτ x. (3.13) Remark 3.4. Let us focus our attention to the symmetric case c0 =c1. We can introduce some randomization of the initial condition as follows: let ǫ be a {−1,1}-valued random variable, independent from the Poisson process Ntc0, withp:= IP(ǫ= 1) = 1−IP(ǫ=−1). It is easy to deduce from (3.12)that we have

IP(ǫZtc0/t∈dx) =

1(dx) + (1−p)δ−1(dx) +g(t, x)dx

e−c0t, (3.14) with

g(t, x) =c0t 2

n I0

c0tp

1−x2

+1 + (2p−1)x

√1−x2 I1

c0tp

1−x2o

1[−1,1](x) andδ1(dx)(resp. δ−1(dx)) is the Dirac measure at1(resp. −1).

In the particular case p = 1/2, x → g(t, x) is an even function. G.H. Weiss ([18] p.393) proved (3.14)using an analytic method based on Fourier-Laplace transform.

Proof of Proposition 3.3. The proof is a direct consequence of the expression of Proposition 3.1. Indeed, for each bounded continuous functionϕwe denote

∆ = IE[ϕ(Ztc0,c1)] =ϕ(t)e−c0t+X

k≥1

2k(ϕ) +X

k≥0

2k+1(ϕ) =ϕ(t)e−c0t+ ∆e+ ∆o, where ∆n(ϕ) = IE[ϕ(Ztc0,c1)1{Nc0,c1

t =n}]. Using (3.2) and (3.3) we get

e=e−τ t Z t

−t

ϕ(z)Se(z)eτ zdz,

(8)

with

Se(z) = 1 2

X

k≥1

(c0c1)k k!(k−1)!

t−z 2

k−1t+z 2

k

= 1

2

√c0c1

rt+z t−z

X

k≥0

1 k!(k+ 1)!

p

c0c1(t2−z2) 2

2k+1

= 1

2

√c0c1

rt+z t−zI1

pc0c1(t2−z2) . For the odd indexes, by (3.4) and (3.5) we get

o=e−τ t Z t

−t

ϕ(z)So(z)eτ zdz, with

So(z) =1 2

X

k≥0

ck+10 ck1

(k!)2

t2−z2 4

k

= c0

2I0

pc0c1(t2−z2) .

Proposition 3.5. 1)(Ztc0,c1, Ntc0,c1; t≥0)is aR×IN-valued Markov process.

2) Lets≥0andn≥0. Conditionally onZsc0,c1=xandNsc0,c1=n,

(Zt+sc0,c1, Nt+sc0,c1), t≥0 is distributed as





x+Rt

0(−1)Nuc0,c1du, n+Ntc0,c1 , t≥0

when nis even, x−Rt

0(−1)Nuc1,c0du, n+Ntc1,c0 , t≥0

otherwise.

Remark 3.6. Note that Propositions 3.5 and 3.1 permit to determine the semigroup of (Ztc0,c1, Ntc0,c1), t ≥ 0

i.e. IP(Ztc0,c1 ∈ dx, Ntc0,c1 = n|Zsc0,c1 = y, Nsc0,c1 = m) where t > s,n≥mandy∈[−s, s].

Proof of Proposition 3.5. Lett > s≥0. Using (2.3) we get Ztc0,c1 =Zsc0,c1+ (−1)Ncs0,c1

Z t−s 0

(−1)N˜usdu, where ˜Nus=Ns+uc0,c1−Nsc0,c1,u≥0.

Note that ( ˜Nus; u≥0)(d)= (Nuc0,c1; u≥0) ifNsc0,c1 ∈2 IN and ( ˜Nus; u≥0)(d)= (Nuc1,c0; u≥0)

ifNsc0,c1∈2 IN +1. This shows Proposition 3.5.

Next, we determine (in Proposition 3.8 below) the Laplace transform of the r.v. Ztc0,c1. It is possible to use the distribution of Ztc0,c1 (cf Proposition 3.3), but this method has the disadvantage of leading to heavy calculations. We develop here a method which uses the fact that (Zsc0,c1; s≥0) is a stochastic process given by (2.3). The key tool is Lemma 3.7 below.

Roughly speaking Lemma 3.7 gives the generator of the Markov process (Zct0,c1, Ntc0,c1).

Lemma 3.7 is an important ingredient in the proof of Proposition 3.11 besides.

Lemma 3.7. Let F : R×IN → R denote a bounded and continuous function such that z→F(z, n) is of classC1 for alln. Then

d

dtIE[F(Ztc0,c1, Ntc0,c1)] = IEh ∂F

∂z(Ztc0,c1, Ntc0,c1)(−1)Ntc0,c1i + IEh

F(Ztc0,c1, Ntc0,c1+ 1)−F(Ztc0,c1, Ntc0,c1)

×

c11{Nc0,c1

t ∈2 IN +1}+c01{Nc0,c1 t ∈2 IN}

i

. (3.15) Proof. Let us denote by ∆(t) = IE[F(Ztc0,c1, Ntc0,c1)]. In order to compute the t-derivative we shall decompose the increment oft→∆(t) in a sum of two terms:

∆(t+h)−∆(t)

h =Bh+Ch,

(9)

with

Bh= 1 h

nIE[F(Zt+hc0,c1, Nt+hc0,c1)]−IE[F(Ztc0,c1, Nt+hc0,c1)]o , Ch= 1

h

nIE[F(Ztc0,c1, Nt+hc0,c1)]−IE[F(Ztc0,c1, Ntc0,c1)]o .

SinceF(·, n) is continuously differentiable with respect to the variablez andt →Ztc0,c1 is differentiable (cf (2.3)), using the change of variable formula we obtain

1 h

nF(Zt+hc0,c1, Nt+hc0,c1)−F(Ztc0,c1, Nt+hc0,c1)o

= 1 h

Z t+h t

∂F

∂z(Zuc0,c1, Nt+hc0,c1)(−1)Nuc0,c1du.

Therefore

h→0limBh= IEh∂F

∂z(Ztc0,c1, Ntc0,c1)(−1)Ntc0,c1i

. (3.16)

In order to study the limit ofCh, we consider two cases: Ntc0,c1 ∈2 IN andNtc0,c1 ∈2 IN +1:

Ch = 1 hIEh

F(Ztc0,c1, Ntc0,c1+ ˜Nhc1,c0)−F(Ztc0,c1, Ntc0,c1) 1{Nc0,c1

t ∈2 IN +1}

i

+ 1

hIEh

F(Ztc0,c1, Ntc0,c1+ ˜Nhc0,c1)−F(Ztc0,c1, Ntc0,c1) 1{Nc0,c1

t ∈2 IN}

i, where ˜Nh=Nt+hc0,c1−Ntc0,c1.

According to Proposition 3.5, conditionally on Ztc0,c1 and Ntc0,c1 ∈ 2 IN (resp. Ntc0,c1 ∈ 2 IN +1), ˜Nh is distributed asNhc0,c1 (resp. Nhc1,c0). Note that Proposition 3.1 implies that IP(Nhc0,c1≥2) =o(h) and

IP(Nhc0,c1= 1) =c0

2

eτ h−e−τ h τ

e−τ h=c0h+o(h).

Consequently

h→0limCh = c1IEh

F(Ztc0,c1, Ntc0,c1+ 1)−F(Ztc0,c1, Ntc0,c1) 1{Nc0,c1

t ∈2 IN +1}

i

+ c0IEh

F(Ztc0,c1, Ntc0,c1+ 1)−F(Ztc0,c1, Ntc0,c1) 1{Nc0,c1

t ∈2 IN}

i

. (3.17) Then, (3.16) and (3.17) clearly imply Lemma 3.7.

Let us introduce the two quantities:

Le(t) = IEh

e−µZtc0,c11{Nc0,c1 t ∈2 IN}

i

andLo(t) = IEh

e−µZtc0,c11{Nc0,c1 t ∈2 IN +1}

i

,(t≥0, µ∈R).

(3.18) Since|Ztc0,c1| ≤t, thenLe(t) andLo(t) are well defined for anyµ∈R. Note thatµ→Le(t) (resp. µ→Lo(t)) is a Laplace transform. We have mentioned thet-dependency only because it will play an important role in our proof of Proposition 3.8 below.

Proposition 3.8. LetLe(t) andLo(t)be defined by(3.18). Then Le(t) = 1

√E

(−µ+τ) sinh(t√ E) +√

Ecosh(t√ E)

e−τ t, (3.19) Lo(t) = c0

√E sinh(t√

E)e−τ t, (3.20)

IE[e−µZc0,c1t ] = 1

√E

h(−µ+τ) sinh(t√ E) +√

Ecosh(t√ E)i

e−τ t, (3.21) whereE =µ2−2τ µ+τ2.

Proof. Applying Lemma 3.7 with the particular functionF(z, n) =e−µz1{n∈2 IN}, we have:

d

dtLe(t) = −µIEh

e−µZct0,c1(−1)Ntc0,c11{Nc0,c1 t ∈2 IN}

i

+ Eh

e−µZtc0,c1 1{Nc0,c1

t ∈2 IN +1}−1{Nc0,c1 t ∈2 IN}

×

c11{Nc0,c1

t ∈2 IN +1}+c01{Nc0,c1 t ∈2 IN}

i

(10)

We deduce

d

dtLe(t) =−(µ+c0)Le(t) +c1Lo(t).

Similarly

d

dtLo(t) = −µIEh

e−µZtc0,c1(−1)Ntc0,c11{Nc0,c1

t ∈2 IN +1}

i

+ Eh

e−µZtc0,c1 1{Nc0,c1

t ∈2 IN}−1{Nc0,c1 t ∈2 IN +1}

×

c11{Nc0,c1

t ∈2 IN +1}+c01{Nc0,c1 t ∈2 IN}

i . We get therefore

d

dtLo(t) = (µ−c1)Lo(t) +c0Le(t).

To sum up

d dt

Le(t) Lo(t)

=

−µ−c0 c1

c0 µ−c1

Le(t) Lo(t)

. We deduce the expressions ofLe(t) andLo(t):

Le(t) =a+eλ+t+aeλt Lo(t) =b+eλ+t+beλt, (3.22) whereλ±=−τ±p

µ2−2τ µ+τ2=−τ±√ E.

The constantsa±andb±are evaluated with the initial conditions:

Le(0) = IP(N0c0,c1∈2 IN) = 1, Lo(0) = IP(N0c0,c1∈2 IN +1) = 0, dLe

dt (0) =−(µ+c0)Le(0) +c1Lo(0) =−µ−c0, dLo

dt (0) = (µ−c1)Lo(0) +c0Le(0) =c0. We obtain

a+= 1 2√

E(−µ+τ+√

E) and a= 1 2√

E(µ−τ+√

E), (3.23)

b+= c0

2√

E and b=− c0

2√

E (3.24)

Using (3.22), (3.23) and (3.24), Proposition 3.8 follows.

It is easy to deduce two direct consequences of Proposition 3.8. First, taking µ= 0 we obtain IP(Ntc0,c1 ∈2 IN) and IP(Ntc0,c1∈2 IN +1). Second, taking the expectation in (2.3) we get the mean ofZtc0,c1.

Corollary 3.9. We have:

IP(Ntc0,c1∈2 IN) = 1 τ h

τsinh(τ t) +τcosh(τ t)i e−τ t, IP(Ntc0,c1∈2 IN +1) =c0

τ sinh(τ t)e−τ t, and

IE[Ztc0,c1] = τ τt+ c0

2(1−e−2τ t).

Remark 3.10. The Laplace transform with respect to the time variable can also be explicitly computed. We define F(µ, s) =R

0 e−stIE[e−µZtc0,c1]dt. Integrating (3.21)with respect todt we get

F(µ, s) = 1 2√ E

E+ (−µ+τ) 1 s−√

E+τ + 1 2√ E

E −(−µ+τ) 1 s+√

E+τ

= (√

E −µ+τ)(s+√

E+τ) + (√

E+µ−τ)(s−√ E+τ) 2√

E((s+τ)2− E)

= 2s√

E+ 4τ√

E −2µ√ E 2√

E((s+τ)2− E) = s+ 2τ−µ (s+τ)2− E

(11)

In the symmetric case,E equalsµ2+c20, then

F(µ, s) = s+ 2c0−µ

s2+ 2sc0−µ2. (3.25)

Let(Zt)be the symmetrization of (Ztc0)which is defined by an initial randomization:

Zt=ǫZtc0, t≥0, whereǫis independent ofZtc0 andIP(ǫ=±1) = 1/2.

Relation (3.25) implies Z

0

e−stIE[e−µZt]dt= s+ 2c0

s2+ 2sc0−µ2. This identity has been obtained by Weiss in [18].

Let us now present a link between the ITN process and the telegraph equation in the particular symmetric case c0 = c1 = c > 0. Recall that (Ntc) is a Poisson process with parameterc.

Letf :R→Rbe a function of classC2 whose first and second derivatives are bounded. We define

u(x, t) = 1 2 n

f(x+at) +f(x−at)o

, x∈R, t≥0.

Then (cf [5])uis the unique solution of the wave equation





2u

∂t2 =a22u

∂x2, u(x,0) =f(x), ∂u

∂t(x,0) = 0.

Proposition 3.11. The function w(x, t) = IEh

u x,

Z t 0

(−1)Nscdsi

, (x∈R, t≥0) is the solution of the telegraph equation (TE)





2w

∂t2 + 2c∂w

∂t =a22w

∂x2, w(x,0) =f(x), ∂w

∂t(x,0) = 0.

This result can be proved using asymptotic analysis applied to difference equation asso- ciated with the persistent random walk [8] or using Fourier transforms [18]. Here we shall present a new proof.

Proof of Proposition 3.11. Applying twice Lemma 3.7 to (z, n) → u(x, z) and (z, n) →

∂u

∂t(x, z)(−1)n we obtain:

∂w

∂t(x, t) = IEh∂u

∂t

x, Z t

0

(−1)Nscds (−1)Ntci

. and

2w

∂t2 (x, t) = IEh∂2u

∂t2 x,

Z t 0

(−1)Nscdsi

−2cIEh∂u

∂t x,

Z t 0

(−1)Nscds

(−1)Ncti . Sinceusolves the wave equation we have

2w

∂t2 (x, t) =a22w

∂x2(x, t)−2c∂w

∂t(x, t).

The functionwis actually the solution of the telegraph equation. It is easy to prove thatw

satisfies the boundary conditions.

Let us note that Proposition 3.11 can be extended to the asymmetric case c0 6=c1. In this general case the telegraph equation is replaced by a linear system of partial differential equations.

(12)

Remark 3.12. 1) In [5], [9], an extension of Proposition 3.11 has been proved. Let Abe the generator of a strongly continuous group of bounded linear operators on a Banach space. Ifw is the unique solution of this abstract ”wave equation”:

2w

∂t2 =A2w; w(·,0) =f, ∂w

∂t(·,0) =Ag (f, g∈ D(A)) thenu(x, t) = IEh

w x,Rt

0(−1)Nscdsi

solves the abstract ”telegraph equation”:

2u

∂t2 =A2u−2c∂u

∂t, u(·,0) =f, ∂u

∂t(·,0) =Ag.

2) In the same vein as [9], Enriquez [6] has introduced processes with jumps to represent solutions of some linear differential equations and biharmonic equations in the presence of a potential term. Moreover useful references are given in [6].

3) It is easy to deduce from Lemma 3.7 that the functions we(x, t) = IEh

u x,

Z t 0

(−1)Nsc0,c1ds 1{Nc0,c1

t ∈2 IN}

i, (x∈R, t≥0)

wo(x, t) = IEh u

x, Z t

0

(−1)Nsc0,c1ds 1{Nc0,c1

t ∈2 IN +1}

i, (x∈R, t≥0) are solutions of the general telegraph system (TS)













2we

∂t2 = (c0c1−c20)we+ (c0c1−c21)wo−2c0

∂we

∂t +a22we

∂x2 ,

2wo

∂t2 = (c0c1−c20)we+ (c0c1−c21)wo−2c1

∂wo

∂t +a22wo

∂x2 , we(x,0) =f(x), wo(x,0) = 0 ∂we

∂t (x,0) =−c0f(x) ∂wo

∂t (x,0) =c0f(x).

4 Convergence of the persistent walk to the ITN

Suppose ρ0 = 1. The aim of this section is to prove the convergence of the interpolated persistent random walk towards the generalized integrated telegraph noise (ITN) i.e. Theorem 2.1. Let us start with preliminary results.

First, let us recall that (Xn, n∈ IN) is the persistent random walk starting in 0 defined by the increments process (Yn, n∈IN) (see Section 1) with transition probabilities

π=

1−c0t c0t

c1t 1−c1t

. Let (Tk; k≥1) be the sign changes sequence of times :

T1= inf{t≥1 :Yt6=Y0}

Tk+1= inf{t > Tk :Yt6=YTk}; k≥1.

(4.1) We putT0= 0 and

Ak=Tk−Tk−1 k≥1 (4.2)

LetNt be the number of times over [0, t] so that the sign of (Yn) changes:

Nt=X

j≥1

1{Tj≤t} (4.3)

The definition ofNt implies that:

Nt=k⇐⇒Tk≤t < Tk+1

Références

Documents relatifs

Consider the first passage Green function of walks which always remain within a given triangle at level.. Level 1 corresponds to the smallest

4 Transient λ-biased random walk on a Galton-Watson tree 111 4.1 The dimension of the harmonic measure... 4.2 Comparison of flow rules on a

The problem of how many trajectories of a random walker in a potential are needed to reconstruct the values of this potential is studied.. We show that this problem can be solved

When the increments (X n ) n∈ N are defined as a one-order Markov chain, a short memory in the dynamics of the stochastic paths is introduced: the process is called in the

We observe that these stochastic processes can be seen as Lévy walks for which the persistence times depend on some internal Markov chain: they admit Markov random walk skeletons..

In that case, the recurrent and the transient property are related to the behaviour of some embedded random walk with an undefined drift so that the asymptotic behaviour depends

In this section, we detail the functional convergence (3.1) when α P p0 , 1q. It turns out that the limit process is no longer a stable Lévy process nor even a Markov process.

We consider in [10] more general random walks W with transitions the weights of a finite- dimensional irreducible g-module V where g is a Lie algebra. The law of W is constructed