• Aucun résultat trouvé

Equilibrium fluctuations for the weakly asymmetric discrete Atlas model

N/A
N/A
Protected

Academic year: 2021

Partager "Equilibrium fluctuations for the weakly asymmetric discrete Atlas model"

Copied!
20
0
0

Texte intégral

(1)

HAL Id: hal-01591186

https://hal.inria.fr/hal-01591186

Preprint submitted on 21 Sep 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 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.

Equilibrium fluctuations for the weakly asymmetric discrete Atlas model

F Hernández, Marielle Simon

To cite this version:

F Hernández, Marielle Simon. Equilibrium fluctuations for the weakly asymmetric discrete Atlas model. 2017. �hal-01591186�

(2)

Equilibrium fluctuations for the weakly asymmetric discrete Atlas model

F. Hern´andez and M. Simon

AbstractThis contribution aims at presenting and generalizing a recent work of Hern´andez, Jara and Valentim [10]. We consider the weakly asymmetric version of the so-calleddiscrete Atlas model, which has been introduced in [10]. Precisely, we look at some equilibrium fluctuation field of a weakly asymmetric zero-range pro- cess which evolves on a discrete half-line, with a source of particles at the origin. We prove that its macroscopic evolution is governed by a stochastic heat equation with Neumann or Robin boundary conditions, depending on the range of the parameters of the model.

1 Introduction

The discrete analogous of the so-calledAtlas model1which has been recently in- troduced in [10] is defined as a family of semi-infinite one-dimensional interacting particle systems, and more specifically, a family of zero-range processes2 with a source at the origin. Let us first give a formal description of these processes: parti- cles are situated on the semi-infinite latticeN={1,2, ...,}. At each site of the lattice Nthere is a random Poissonian clock of rate 2, which is independent of all the other random clocks attached to the other sites. Each time the clock at sitexNrings, Marielle Simon

Equipe MEPHYSTO, Inria Lille – Nord Europe, France´ andLaboratoire Paul Painlev´e, UMR CNRS 8524, Lille, France.

e-mail: marielle.simon@inria.fr Freddy Hern´andez

Instituto de Matem´atica e Estat´ıstica, Universidade Federal Fluminense, Rua M´ario Santos Braga S/N, Niter´oi, RJ 24020-140, Brazil.

e-mail: freddyhernandez@id.uff.br

1The continuous Atlas model is given by a semi-infinite system of independent Brownian motions onR, see for instance [5, 11], and also [10] for more details.

2We refer to [13] for a review on zero-range processes.

1

(3)

one of the particles at this site moves to one of its two neighboursx1 orx+1, with equal probability. If the particle decides to move to site 0, then the particle leaves the system. In addition, with exponential rateλn>0 a particle is created at site 1.

We consider in this article theweakly asymmetric versionof that discrete Atlas model, whose dynamics is now described as follows: whenever the clock at sitex Nrings and there is a particle at this site, one particle jumps to its right neighbouring sitex+1 with probability (1αn)/2 and to its left neighbouring sitex1 with probability(1+αn)/2, for someαn>0 which will be clarified later on. In the same way, atx=1, if the particle tries to jump left, it disappears; besides, a particle is created at sitex=1 with exponential rateλn(1αn).

In [10] the caseαn=0 (i.e. the symmetric case) andλn=1b/nwithb>0 has been completely investigated. Here we generalize the model by choosingαn=a/nα andλn=1b/nβ for somea,b>0 andα,β >0. For this choice of parameters, the equilibrium state of the system is still given by a product geometric measure on {0,1,2, . . .}N, as in [10]. In this work we show that:

ifα >1 andβ 61, the stationary space-time current fluctuations converge, in a suitable rescaling, to the solution of the stochastic heat equation with Neuman boundary conditions, similarly to [10];

ifα =1 andβ 61, the stationary space-time current fluctuations converge, in a suitable rescaling, to the solution of the stochastic heat equation with Robin boundary conditions.

Very recently, a mathematical breakthrough has been achieved, towards theweak KPZ universality conjecture(see [4, 7, 8, 9] for instance). This conjecture states that the fluctuations of a large class of weakly asymmetric one-dimensional interacting particle systems should converge to the KPZ equation. Theweakly asymmetricAtlas model cannot be directly treated by the approach of [7], because of the presence of a source of particles at one boundary. Therefore, that model needs a special care, which we start here in this work.

The range of the parameters for which we obtain the same macroscopic behaviour of [10] may not be sharp. The limit values that we obtain forβandα, come from the limitation of the main tool that we use, namely thefirst order Boltzmann-Gibbs prin- ciple. One could try to improve our results, and treat other ranges ofβ,α, by using other estimating tools. This may lead to new macroscopic limits. For instance, one can think about the recent principle stated by [7], called thesecond order Boltzmann- Gibbs principle, which permits to obtain KPZ-type macroscopic fluctuations, under a stronger asymmetry. We let this prospective research to a future work. Finally, let us mention a related work [6] which studies a similar model but with different scalings.

This article is organized as follows. In Section 2 we define the zero-range process model with a weak asymmetry and a source at the origin, and we introduce the current fluctuation field in the stationary state. In Section 3, we sketch the proof and relate the density field to the current field. We also highlight the main differences with the model of [10], namely the presence of possibly diverging boundary terms.

In Section 4 we finally prove the convergence of the density fluctuation field, and

(4)

we show various estimates related to the variance of some additive functionals of our dynamics. The exposition closely follows [10], therefore we omit some details of the proofs that were already included there.

2 Framework 2.1 Notations

We denoteN0:={0,1,2, ...}andN:={1,2, ...,}. Our system of particles evolves onN, more precisely on each sitexNthere is a numberηt(x)N0of particles which depends on timet>0. The state space of the dynamics is therefore0:=NN

0. For anyx,yNsuch thatx6=yandη(x)>1, we define the configurationηx,yas being obtained fromη0when a particle moves from sitexto sitey, namely:

(ηx,y)(z) =

η(x)1 ifz=x, η(y) +1 ifz=y, η(z) otherwise.

We also defineη0,1 andη1,0as follows. First, η0,1is the configuration obtained fromηwhen a particle is created at site 1, namely:

(η0,1)(z) =

(η(1) +1 ifz=1, η(z) otherwise.

Similarly, ifη(1)>1, the configurationη1,0is obtained fromη when a particle is suppressed at site 1, namely

(η1,0)(z) =

(η(1)1 ifz=1, η(z) otherwise.

We say that a functionϕ:0Rislocalif it depends onη0only through a finite number of coordinates.

2.2 The microscopic dynamics

Let us now rigorously define, as a Markov process{ηtn(x); xN}t>0, the asym- metric dynamics described in the introduction.

Letg:N0Rbe given byg(k) =1 ifk6=0 andg(0) =0, and consider three parametersλ,p,q(0,1)such thatp+q=1. We define an operator acting on local functionsϕ:0Ras

(5)

Lp,q,λϕ(η):=

x=1

g(η(x))

q∇x,x+1ϕ(η) +p∇x,x1ϕ(η)

+λq0,1ϕ(η),

where we denote

x,yϕ(η):=ϕ(ηx,y)ϕ(η).

We let the reader refer to [10] and check that the Markov process associated with the linear operatorLp,q,λis well defined in infinite volume. Letµλdenote the product geometric measure on0given by

µλ(dη) =

xN

(1λ)λη(x)dη(x).

It is not difficult to see that µλ is an invariant measure under the evolution of the Markov process generated byLp,q,λ.

LetnNbe a scaling parameter. Fixa,b,α,βR+withα,β>0, and define λn=1 b

nβ, αn= a

nα, pn=1+αn

2 , qn=1αn

2 . (1)

These parameters are fixed from now on and up to the end. Finally, let us denote Ln:=Lpn,qnn, (2) and let{ηtn;t>0}be theacceleratedprocess in the time scaletnθ, generated by nθLn(for some θ>0), starting from the equilibriumµλn. We denote by Pnthe probability distribution of{ηtn;t>0}and byEnthe corresponding expectation.

2.3 Current fluctuation field

For anyxN, letJtn(x)be the cumulative current of particles betweenxandx+1 up to timet, that is, the signed number of particles which have crossed the bond {x,x+1}up to timet. Analogously, denote byJtn(0)the number of particles created atx=1 minus the number of particles that disappeared atx=1, up to timet.

Our aim is to describe the asymptotic behaviour of the current processes{Jtn(x);t>

0} as n goes to infinity. For that purpose we consider the measure-valued pro- cess {Zn

t ; t >0} defined for any smooth and compactly supported function f C

c ([0,∞),R)by

Zn

t (f):= 1 nγ

xN0

Jtn(x)f nx

, (3)

whereγ>1 is a parameter that will be made precise below, andJtn(x)is the recen- tred current defined as:Jtn(x):=Jtn(x)En[Jtn(x)].

For a reason that will become clear in Section 3, instead ofZn

t we shall actually work with the field{Xn

t ;t>0}defined by

(6)

Xn

t (f):= 1 nγ

xN0

Jtn(x) f xn

+ 1

nγ1

xN

η0n(x)F xn

, (4)

whereηtn(x):=ηtn(x)En[ηtn(x)]andF(u):=Ruf(y)dy.

2.4 Main results

Let ˙Wt be a standard space-time white noise on[0,∞)×[0,∞).

Definition 1.LetA,BR. A measure-valued stochastic process{Xt}t>0is said to be a martingale solution of the stochastic heat equation

tXt+A∇Xt=BYt+

2 ˙Wt, (5)

with boundary condition (BC) atx=0 if for any function f C

c ([0,∞),R)satis- fying the boundary condition (BC) atx=0, the real-valued process

Xt(f) +A Z t

0

Xs(f)dsB Zt

0

Xs(f′′)ds

is a continuous martingale of quadratic variation 2tR0(f(u))2du.

Uniqueness of martingale solutions to the stochastic heat equation with Neumann and Robin boundary condition can be found, for instance, in [3] and [2, Proposition 2.7], respectively.

Theorem 1.Letα>1andβ (0,1].

Assumingθ=2+2βandγ=β+32, the sequence of processes{Xtn;t>0}nN

converges in distribution with respect to the uniform topology to the martingale solution of the stochastic heat equation

tXt=b2

2Xt+

2 ˙Wt, (6)

with Neumann boundary condition f(0) =0.

Theorem 2.Letα=1andβ (0,1].

Assumingθ=2+2βandγ=β+32, the sequence of processes{Xn

t ;t>0}nN

defined in(4)converges in distribution with respect to the uniform topology to the martingale solution of the stochastic heat equation

tXt+ab2Xt=b2

2Xt+

2 ˙Wt, (7)

with Robin boundary condition f(0)2f(0) =0.

Note that [10, Theorem 2.1] can be recovered from Theorem 1 by choosinga=0 (orα=∞) andβ=1.

(7)

3 Sketch of the proof

The proof of Theorem 1 and Theorem 2 follows from standard arguments: after proving tightness of the sequence of processes{Xn

t }n, one has to show that any of its limit points is a martingale solution of (5) with the appropriate initial condition.

For that purpose, we write themartingale decompositionofXn

t (see Section 3.1 below), and investigate each of its terms. The estimate which is going to be crucial is the so-calledfirst order Boltzmann-Gibbs principle, which is proved in Section 4.3 by using very precise bounds for additive functionals of Markov processes. Since the whole proof is close to [10], we do not copy the exposition of all technical tools which were proved there, but we refer to them when needed, and we focus on the estimates which turn out to be different in our case, or even new.

3.1 Martingale decomposition

Using Dynkin’s formula, applied to the Markov process

ηtn,Jtn(x)

;t>0 , we can write, for anyxN

Jtn(x) =Mtn(x) + Z t

0

jnx,x+1(ηsn)ds, (8)

where jx,x+1(η)is the microscopic current between sitesxandx+1 given by jx,x+1(η) =nθh

qng(η(x))png(η(x+1))i

, (9)

and the processes{Mtn(x);t>0}are martingales with quadratic variation given by hM·n(x)it=nθ

Z t 0

qng(ηsn(x)) +png(ηsn(x+1))

ds. (10)

Let us introduce the centred microscopic current

jnx,x+1:=jx,x+1n En[jnx,x+1] =jnx,x+1+nθαnλn, (11) and note that one can easily compute:

Jtn(x) =Jtn(x)En[Jtn(x)] =Jtn(x) +nθαnλnt.

For the sake of clarity, in the following we adopt the conventiong(η(0)) =λn. In particular, expressions (8)-(10) also hold forx=0. Since the currentsJtn(x)have dis- joint jumps, the martingales{Mtn(x);t>0}xN0 are mutually orthogonal. There- fore, with our definition (3), for any f C

c ([0,∞),R), the field Zn

t (f) can be written as

(8)

Ztn(f) =Mtn(f) + Z t

0

1 nγ

xN0

jnx,x+1(ηsn)f xn ds,

whereMtn(f)is a martingale, given by Mtn(f):= 1

nγ

xN0

Mtn(x)f xn ,

and with quadratic variation given by M·n(f)

t=nθ Z t

0

xN0

qng(ηsn(x)) +png(ηsn(x+1)) f2 xn

ds. (12) Explicit computations using (9), (11), and integration by parts, show thatZn

t (f) can be decomposed as follows:

Zn

t (f) = Mtn(f) +Bn

t(f) +Cn

t (f) (13)

where

Bn

t(f):=nθγ1(1+αn) 2

Z t 0

xN

g(ηsn(x))λn

nx1f ds (14) Cn

t (f):=αnnθγ Z t

0

xN

g(ηsn(x))λn f nx

ds, (15)

where we have used the standard notation for the discrete gradient of f: for any xN0,nxf :=n[f x+1n

f nx ].

Thefirst order Boltzmann-Gibbs principlestated ahead shall allow us to replace g(ηsn(x))λnin (14) and (15) above byb2n(ηsn(x)ρn)at a small price (de- pending on the values of parametersθ,γ,α,β), obtaining then

Zn

t (f) =Mtn(f) +b2nθγ1(1+αn) 2

Z t 0

xN

ηsn(x)ρn

nx1f ds

b2αnnθγ Z t

0

xN

ηsn(x)ρn f nx

ds+on(1), (16) whereon(1)denotes a random sequence(εn)which satisfiesEn[εn2]0 asn∞.

From the continuity relationJtn(x1)Jtn(x) =ηtn(x)η0n(x), which is valid for anyxN, we have

xN

ηtn(x)f nx

=

xN

η0n(x)f nx + 1

n

xN

Jtn(x)∇nxf + f 1n

Jtn(0). (17) Therefore, the right hand side of (16) can be rewritten as

(9)

Mtn(f) +b2nθγ1(1+αn)

2 (18)

× Z t

0

xN

η0n(x)∇nx1f + 1 n

xN

Jsn(x)xnf +n0f Jsn(0)

ds

b2αnnθγ Z t

0

xN

η0n(x)f xn + 1

n

xN

Jsn(x)∇nxf + f 1n Jsn(0)

ds+on(1),

where xnf is the discrete Laplacian of f defined asxnf :=n[∇nx+1fnxf] = n2[f x+1n

2f nx

+f xn1 ].

It is exactly this expression that justifies to considerXn

t (see (4) for the defini- tion) instead ofZn

t , as stated in the final of Subsection 2.3. In fact, after replacing in (18)

nxf and xnf by f xn

and f′′ xn

, (19)

respectively (for which the error will be of orderon(1)), we obtain from (16) that the martingaleMtn(f)reads as

Mtn(f) = Xtn(f)X0n(f) +cn Z t

0

Xsn(f)dsdn Z t

0

Xsn(f′′)ds (20)

dnn0fcnf 1n n1γ

Z t 0

Jsn(0)ds+on(1), (21) where

cn:=b2αnnθ1 and dn:=b2

2nθ2(1+αn).

Thanks to this decomposition, we will be able to close the martingale problem (20)–

(21), as explained in the next section.

3.2 Closing the martingale problem

Let us now obtain conditions on the parametersα,β,γ,θthat permit to rewrite (20)- (21) as an approximate closed equation for the fluctuation fieldXn

t . In what follows we denoteηn1 if the real sequence(ηn)converges to a constantc6=0 asn∞.

We say that a functiong(n):NRis of orderO(εn)if there exists a constantC>0 such that, for anyn,|g(n)|6Cεn.

i) Quadratic variation of Mtn(f): Recall (12). In order to makeEn[hM·n(f)it] converge to 2tR0f2(u)duasn∞, we need to impose

θ=2γ1. (22)

ii) Initial fieldXn

0(f):One can easy check the following: from the Central Limit Theorem (recall that the variables{η0(x); xN} are i.i.d. distributed under µλn) and the fact that the variance ofη0(x), namely Var[η0(x)], is of ordern,

(10)

we have that: ifγβ >32 thenXn

0(f)converges, asn goes to infinity, to a Gaussian random variable (possibly degenerate). Moreover if γβ < 32 this term explodes.

iii) Boltzmann-Gibbs principle: To apply the Boltzmann-Gibbs principle, that is used twice forBn

t(f)andCn

t (f)(see Section 4.3), we will need two conditions:

2(θγ1)<θ13β

2 and 2(θγα)<θ13β 2 . iv) Real values cn anddn :First, note that the conditionscn1 ordn1 are

equivalent to:

cn1 θ2β =1+α

dn1 θ2β =2. (23)

In addition, sincenαndn=1+α2ncn, we have:

ifα(0,1),

cn1dn0 and dn1cn;

ifα=1,

cn1dn1 ;

ifα(1,∞),

cn1dn and dn1cn0.

This means that we can already exclude two cases: first,α(0,1)anddn1;

and second,α(1,∞)andcn1. Furthermore, we can exclude altogether the caseα (0,1). In fact, in view of (22) and (23), ifα(0,1)andcn1 then γβ=1+α2, and hence the initial field would explode, as noticed in ii) above.

v) Border term:In Section 4.4, we will prove the following estimate:

Enh

Jtn(0)2i

6Ctnθ1, (24)

for some positive constantC>0. Then, observe that (24) combined with (22) implies thatEn

n1γJtn(0)2

is bounded.

Furthermore, one can easily see that dnn0fcnf 1n

is of order

(11)

nθ1α if

α(0,1)andf(0)6=0 or

1<α<2 and f(0) =0 , nθ2 if

α(0,1) and f(0) =0 or

α=1 and (f(0)2a f(0))6=0 or

α>1 and f(0)6=0 , nθ3 if

α=1 and (f(0)2a f(0)) =0 or

α>2 and f(0) =0 .

(25) To sum up, if we want simultaneously:

i) the quadratic variation onMtn(f)to be of orderO(1)and not vanish asn∞;

ii) the initial fieldXn

0 not explode;

iii) the Boltzmann-Gibbs principle to be valid;

iv) cnordnto be of orderO(1)and neithercnnordngoes to infinity asngoes to infinity;

we have to impose:

θ=2+2β, γ=β+32, β 0,43

. (26)

If in addition we want

(v) the border term (21) to go to zero asngoes to infinity, in view of (25) and (26) we have to impose:

α=1 and f(0)2a f(0) =0

or

α>1 andf(0) =0

. (27)

Remark 1.Note that Theorems 1 and 2 are stated with a stronger condition onβ, namely β 61 instead of β < 43. This assumption only comes out when proving tightness of the sequenceXn

t , as explained in the next section.

From now on, the parameters of our model satisfy (26) and (27). With these choices, and recalling (20)–(21), we have shown that any possible limit point ofXn

t satisfies Definition 1, with the suitable boundary conditions. The only missing argument to conclude the proof of Theorems 1 and 2 is tightness of the processes, which is given in the next section. We will also expose all technical proofs that are still missing, namely the Boltzmann-Gibbs principle, and the proof of (24).

4 Proof details and technical estimates

In this section we prove the results which will turn the sketch of the proof of The- orems 1 and 2, exhibited in Section 3, into a rigorous proof. The arguments closely follow the ones in [10], therefore the exposition will be inspired by [10, Sections 3-4], and we will explain the differences on the go.

Références

Documents relatifs

With the specific aim of comparing various models on the same loading history, we also propose in this paper a method to calculate a complete set of parameters from two creep

of standard linear primitives (half-spae, point, m-at and m-simplex) and. how they an

5.2 A version of the Harthong-Reeb line based on Ω-numbers An instantiation of the minimal axiom system for non-standard arith- metic All properties of our minimal axiom system

The initial condition is the following: at time zero, each site in x ∈ Z bears a random number of particles whose distribution is Poisson with parameter ρ &gt; 0, the.. Key words

In the initial step, all lists are sorted (using quicksort, average O(n log n)) and one axis is used to create initial interactions: the range between lower and upper bound for

We have chosen to translate the Frenkel-Kontorova model into the framework of discrete Lagrangian Aubry-Mather theory mainly to have a simple notion of minimizing holonomic

Our local transition function[8], is able to represent phenomena on a wide time scale such as : a dip of the lower subducted plate (80 km / 1 million years), a landward migration of

It is the purpose of this manuscript to formulate a numerical model that is capable of describing dynamic shear band propagation in a porous medium, with a solid skeleton and