• Aucun résultat trouvé

Boundary driven Kawasaki process with long range interaction: dynamical large deviations and steady states

N/A
N/A
Protected

Academic year: 2021

Partager "Boundary driven Kawasaki process with long range interaction: dynamical large deviations and steady states"

Copied!
37
0
0

Texte intégral

(1)

HAL Id: hal-00634205

https://hal.archives-ouvertes.fr/hal-00634205v2

Preprint submitted on 30 Sep 2012

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.

Boundary driven Kawasaki process with long range interaction: dynamical large deviations and steady states

Mustapha Mourragui, Enza Orlandi

To cite this version:

Mustapha Mourragui, Enza Orlandi. Boundary driven Kawasaki process with long range interaction:

dynamical large deviations and steady states. 2011. �hal-00634205v2�

(2)

RANGE INTERACTION: DYNAMICAL LARGE DEVIATIONS AND STEADY STATES

MUSTAPHA MOURRAGUI AND ENZA ORLANDI

Abstract. A particle system with a single locally-conserved field (density) in a bounded interval with different densities maintained at the two endpoints of the interval is under study here. The particles interact in the bulk through a long range potential parametrized byβ0 and evolve according to an exclu- sion rule. It is shown that the empirical particle density under the diffusive scaling solves a quasi-linear integro-differential evolution equation with Dirich- let boundary conditions. The associated dynamical large deviation principle is proved. Furthermore, forβ small enough, it is also demonstrated that the empirical particle density obeys a law of large numbers with respect to the stationary measures (hydrostatic). The macroscopic particle density solves a non local, stationary, transport equation.

1. Introduction

Over the last few years there have been several papers devoted to understand- ing the macroscopic properties of systems out of equilibrium. Typical examples are systems in contact with two thermostats at different temperature or with two reservoirs at different densities. A mathematical model is provided by interacting particles performing local reversible dynamics (for example reversible hopping dy- namics) in the interior of a domain and some external mechanism of creation and annihilation of particles at the boundary of the domain, modeling the reservoirs.

The full process is non reversible. The stationary non equilibrium states are char- acterized by a flow of mass through the system and long range correlations are present. We refer to [3] and [6] for two reviews on this topic. We study a model with such features but in a regime where phase separation might occur at equilib- rium for the underlying reversible model. To this aim we consider in the interior of the domain a reversible dynamics (Kawasaki dynamics) constructed trough a mean field interaction (Kac potential), see below for more details. This is the first time, to our knowledge, where both, long range dynamics in the bulk and creation and annihilation of particles at the boundary of the domain, are taken into account.

The particle models we consider are dynamic versions of lattice gases with long- range Kac potentials, i.e. the interaction energy between two particles, say one at xand the other aty(xandyare both inZd), is given byJN(x, y) =N−dJ(Nx,Ny)

Date: September 30, 2012.

2000Mathematics Subject Classification. Primary 82C22; Secondary 60F10, 82C35.

Key words and phrases. Kawasaki process, Kac potential, Large deviations, Stationary nonequilibrium states.

M.M. supported by ANR Blanc LHMSHE, ANR-2010-BLANC-0108, Grefi- Mefi 2010, Projet risc.

E.O. supported by MURST/Cofin Prin2011, ROMA TRE University, Rouen University.

(3)

where J is a smooth symmetric probability kernel with compact support and N is a positive integer which is sent to ∞. The equilibrium states for these mod- els have been investigated thoroughly [12], [18] and [22], and have provided great mathematical insight into the static aspects of phase transition phenomena. The dynamical version of these lattice gases in domain with periodic boundary condi- tions (reversible dynamics) has been analyzed in [8], [9], [10] and [1]. This paper starts by studying the dynamics of these systems in a bounded interval with reser- voirs (non-reversible dynamics). We investigate their qualitative behavior in the range of the parameter when at equilibrium phase transition is present. Let us describe informally the dynamics. We consider a one dimensional lattice gas with particle reservoirs at the endpoints. The restriction on the dimension is done only for simplicity. Given an integer N >1 at any given time each site of the discret set {−N, . . . , N} is either occupied by one particle or empty. The interaction en- ergy among particles is given by a modified version of the Kac potential JN and it is tuned by a positive parameterβ. The modification of the Kac potential, see (2.1), takes into account that the particles are confined in a bounded domain. In the bulk each particle jumps at random times on the right or on the left nearest neighborhood, if the chosen site is empty, at a rate which depends on the particle configuration through the Kac potential. Whenβ= 0 we have the simple exclusion process. At the boundary sites,±N, particles are created and removed for the local density to be 0< ρ≤ρ+<1. At rateρ± a particles is created at ±N if the site is empty and at rate 1−ρ± the particle at±N is removed if the site is occupied.

The dynamics described above defines anirreducibleMarkov jump process on a finite state space, i.e. there is a strictly positive probability to go from any state to another. By the general theory on Markov process [17], the invariant measure µstatN is unique and encodes the long time behavior of the system. Let Pβ,N

µstatN be the stationary process. SinceµstatN is invariant, the measurePβ,N

µstatN is invariant with respect to time shifts. The measure Pβ,N

µstatN is invariant under time reversal if and only if the measure µstatN is reversible for the process, i.e if the generator of the process satisfies the detailed balance condition with respect toµstatN . In the case at handµstatN is stationary and reversible only ifβ= 0 andρ+=ρ. In such case the invariant, reversible measure is the Bernoulli product measure with marginals ρ. When β > 0 and /or ρ+ 6= ρ then µstatN is not reversible. In such case the corresponding process is denotednon reversible. We shall consider the latter case.

The lattice space is scaled by N1 and the time by N2 (diffusive limit) while the behavior is studied, as N ↑ ∞, of the empirical density of the particles evolving according to the dynamics described above.

To prove the hydrodynamic behavior of the system, we follow the entropy method introduced by Guo, Papanicolaou, and Varadhan [11]. It relies on an estimate on the entropy of the state of the process with respect to a reference invariant state.

The main problem in the model considered here is that the stationary state is not explicitly known. We have therefore to consider the entropy of the state of the process with respect to a state which is not invariant, for instance, with respect to a product measure with a slowly varying profile. Since this measure is not invariant, the entropy does not decrease in time and we need to estimate the rate at which it increases. These estimates on the entropy are deduced in Subsection 3.1

(4)

It results that the local empirical particle density, in the diffusive limit, is the solution of a boundary value problem for a quasilinear integro differential para- bolic equation, see (2.9). In addition to this, it is demonstrated that when β is small enough, then the empirical particle density obeys a law of large numbers with respect to the stationary measures (hydrostatic). The result is obtained char- acterizing the support of any limit points ofPβ,N

µstatN . This intermediate result holds for anyβ ≥0. Then it is shown that withβ small enough, the stationary solution of the boundary problem (2.9) is unique, it is a global attractor for the macroscopic evolution with a decay rate uniform with respect to the initial datum. This holds only for β < β0 whereβ0>0 depends on the diameter of the domain and on the chosen interactionJ. Namely the quasilinear non local parabolic equation does not satisfy a comparison principle, which is the main tool used in previous papers, see for example [20] and [7], to show the hydrostatic. For value larger thanβ0 we are not able to show the uniqueness of the stationary solution of the boundary value problem (2.9). We stress that β0 < βc where βc is the value above which phase segregation occurs at equilibrium; with the choice made of the parametersβc =14, see page 1712 of [9]. Further, we prove the dynamical large deviations for the em- pirical particle density. The large deviation functional is not convex as function of the density, it is lower semicontinuous and has compact level sets. Since the large deviation functional is not convex and the underlying dynamics does not satisfy any comparison principle, care needs to be taken to prove the lower bound. The basic strategy to show the lower bound consists in obtaining this bound for smooth paths and then applying a density argument. The argument goes as follows: Given a path ρwith finite rate functional I(ρ) one constructs a sequence of smooth pathsρn so that ρn → ρin a suitable topology and I(ρn)→I(ρ). When the large deviation functional is convex, this argument is easily implemented. In our case, because of the lack of convexity we modify the definition of the rate functional declaring it infinite if a suitable energy estimate does not hold. In this way the modified rate functional when finite provides the necessary compactness to close the argument.

This method has been developed in [21] and we adapted to our model. The mod- ification of the rate functional helps in showing the lower bound but makes more difficult the upper bound. One needs to show that the energy estimate holds with probability super exponentially close to one. A similar strategy was applied in [20]

and [4].

In a recent paper, De Masi et alii, [5], constructed, in the phase transition regime, stationary solutions of a boundary value problem equivalent to (2.9) in which the density ρis replaced by the magnetization m = 2ρ−1. They did not study the derivation of the boundary value problem from the particle system and they did not inquire about the uniqueness of stationary solutions. They investigated the qualitative behavior of constructed stationary solutions of (2.9) as the diameter of the domain goes to infinity. They proved, for the constructed solution, the validity of the Fourier law in the thermodynamic limit showing that, in the phase transition regime, the limit equilibrium profile has a discontinuity (which defines the position of the interface) and satisfies a stationary free boundary Stefan problem.

The paper is organized as follows: The precise feature of the model, notations and results are stated in Section 2, In Section 3, some basic estimates needed along the paper are collected. In Section 4 the hydrodynamic and the hydrostatic limits are shown. Section 5 is split into 5 subsections and deals with dynamical large

(5)

deviations. We prove in Section 6 the weak uniqueness of the solution of the quasi- linear integro-differential equation. Furthermore, for β small enough, it is shown that its stationary solution is unique and it is a global attractor in L2.

2. Notation and Results

Fix an integer N ≥ 1. Call ΛN = {−N, . . . , N} and ΓN = {−N,+N} the boundary points. The sites of ΛN are denoted byx, y and z. The configuration space isSN ≡ {0,1}ΛN which we equip with the product topology. Elements ofSN, called configurations, are denotedη so thatη(x)∈ {0,1}stands for the number of particles at sitexof the configurationη.

We denote Λ = (−1,1) (¯Λ = [−1,1]) the macroscopic open (closed) interval, Γ ={−1,1}its boundary and u∈[−1,1] the macroscopic space coordinate.

2.1. The interaction. To define the interaction between particles we introduce a smooth, symmetric, translational invariant probability kernel of range 1, i.e J(u, v) =J(0, v−u) for allu, v∈R,J(0, u) = 0, for all|u|>1, andR

J(u, v)dv= 1, for all u∈ R, 1. When (u, v)∈ Λ×Λ we define the interaction Jneum(u, v) im- posing a reflection rule: uinteracts withv and with the reflected points ofv where reflections are the ones with respect to the left and right boundary of Λ. For this reason it is referred to as “Neuman” interaction. More precisely, we define for u andv in Λ

Jneum(u, v) :=J(u, v) +J(u,2−v) +J(u,−2−v), (2.1) where 2−vis the image ofvunder reflections on the right boundary{1}and−2−v is the image ofv under reflections on the left boundary{−1}. By the assumption onJ,Jneum(u, v) =Jneum(v, u) andR

Jneum(u, v)dv= 1 for allu∈Λ, see Lemma 3.1. We defined the interaction (2.1) by boundary reflections only for convenience.

It has the advantage to keepJneum a symmetric probability kernel. This choice of the potential has been done already in [5].

The pair interaction betweenxandy in ΛN is given by JN(x, y) =N−1Jneum(x

N, y N).

The total interaction energy among particles is given by the following Hamiltonian HN(η) =−1

2 X

x,y∈ΛN

JN(x, y)η(x)η(y). (2.2) 2.2. The dynamics. We denote by ηx,y the configuration obtained from η by interchanging the values atxandy:

x,y)(z) :=





η(y) if z=x η(x) if z=y η(z) if z6=x, y,

and by σxη the configuration obtained fromη by flipping the occupation number at sitex:

xη)(z) :=

(1−η(x) if z=x η(z) if z6=x.

1One could take the interactionJ so that for allu R,R

J(u, v)dv= awitha >0. The only difference, with the case at hand, is that the the underlying reversible particle model has, at equilibrium, phase transition forββc=4a1, see [22].

(6)

We denote forf :SN →R,x, y∈ΛN andη∈ SN, (∇x,yf)(η) =f(ηx,y)−f(η).

The microscopic dynamics is specified by a continuous time Markov chain on the state spaceSN with infinitesimal generator given by

LN =Lβ,N +L−,N+L+,N , (2.3)

where for functionf :SN →R (Lβ,Nf) (η) = X

x∈ΛN ,y∈ΛN

|x−y|=1

CNβ(x, y;η) [(∇x,yf)(η)] , (2.4)

with rate of exchange occupanciesCNβ given by CNβ(x, y;η) = exp

−β

2[HNx,y)−HN(η)]

, (2.5)

whereHN is the Hamiltonian (2.2);

(L−,Nf)(η) = c η(−N)

f(σ−Nη)−f(η) , (L+,Nf)(η) = c+ η(N)

f(σNη)−f(η) , where for anyρ±∈(0,1), c±:{0,1} →Rare given by

c±(ζ) := ρ±(1−ζ) + (1−ρ±)ζ .

The generatorLβ,N describes the bulk dynamics which preserves the total number of particles, the so called Kawasaki dynamics, whereasL±,N, which is a generator of a birth and death process acting on ΓN, models the particle reservoir at the boundary of ΛN. The rate of the bulk dynamics{CNβ(x, y;η), x∈ΛN, y∈ΛN}, see (2.5), satisfies the detailed balance with respect to the Gibbs measure associated to (2.2) with chemical potentialλ∈R

µβ,λN (η) = 1

ZNβ,λexp{−βHN(η) +λ X

x∈ΛN

η(x)}, η∈ SN , (2.6) whereZNβ,λis the normalization constant. For the bulk rates this means

CNβ(x, y;η) =e−β[HNx,y)−HN(η)]CNβ(y, x;ηx,y).

For the generator it means thatLβ,N is self-adjoint w.r.t. the Gibbs measure (2.6), for anyλ∈R. The corresponding process is denoted reversible.

For a positive functionρ: ΛN →(0,1) we denote byνNρ(·) the Bernoulli measure with marginals

νNρ(·)(η(x) = 1) =ρ(x/N) = 1−νNρ(η(x) = 0), x∈ΛN, η∈ SN. (2.7) The Bernoulli measure νNρ+ρ) is reversible with respect to the boundary gen- eratorL+,N (L−,N).

The Markov process associated to the generatorLN, see (2.3), isirreducibleand we denote byµstatNstatN (β, ρ, ρ+) the unique invariant measure. In the notation we stress only the dependence on the parameters relevant to us. This means that for anyf :SN →R

Z

SLNf(η)dµstatN (η) = 0,

(7)

but the generatorLN is notself-adjoint with respect toµstatN . The corresponding process is called non reversible. The only case where the process is reversible is whenρ+andβ= 0. In such case the product measure associated toρ+

is invariant and the process with generatorLN is also reversible.

We denote byMthe space of positive densities bounded by 1:

M:=

ρ∈L [−1,1], du

: 0≤ρ≤1 ,

wheredustands for the integration with respect to the Lebesgue measure on [−1,1].

We equipMwith the topology induced by the weak convergence of measures and denote byh·,·ithe duality mapping. A sequence{ρn} ⊂ Mconverges toρinMif and only if

n, Gi= Z

Λ

ρn(u)G(u)du→ hρ, Gi

for any continuous functionG: [−1,1]→R. Note thatMis a compact Polish space that we consider endowed with the corresponding Borelσ-algebra. The empirical density of the configurationη∈ SN is defined asπN(η) where the mapπN: SN → Mis given by

πN(η) (u) :=

NX−1 x=−N+1

η(x)1x

N2N1 ,Nx +2N1 (u),

in which1{A}stands for the indicator function of the setA. Let{ηN}be a sequence of configurations withηN ∈ SN. If the sequence{πNN)} ⊂ M converges toρin Mas N → ∞, we say that {ηN} is associated to the macroscopic density profile ρ∈ M.

2.3. Functional Spaces. Fix a positive time T. Let D([0, T],M), respectively D([0, T],SN), be the set of right continuous with left limits trajectoriesπ: [0, T]→ M, resp. (ηt)t∈[0,T] : [0, T] → SN, endowed with the Skorohod topology and equipped with its Borelσ−algebra. TakeµN onSN and denote by (ηt)t∈[0,T] the Markov process with generatorN2LN starting, at timet= 0, byη0distributed ac- cording toµN. Notice that the generator of the process has been speeded up byN2. This corresponds to the diffusive scaling. Denote byPβ,Nµ

N the probability measure on the path spaceD([0, T],SN) corresponding to the Markov process (ηt)t∈[0,T]and byEβ,Nµ

N the expectation with respect toPβ,Nµ

N . WhenµNη for some configura- tionη ∈ SN, we write simply Pβ,Nη =Pβ,Nδ

η and Eβ,Nη =Eβ,Nδ

η . We denote by πN the map from D([0, T],SN) to D([0, T],M) defined byπN·)t = πNt) and by Qβ,NµN = Pβ,Nµ

N ◦(πN)−1 the law of the process πNt)

t≥0. For m ∈L([−1,1]) andu∈Λ we denote

(Jneum⋆ m)(u) = Z

Λ

Jneum(u, v)m(v)dv.

We need some more notations. For integersn andm we denote by Cn,m([0, T]× [−1,1]) the space of functionsG=Gt(u) : [0, T]×[−1,1]→Rwithn derivatives in time and mderivatives in space which are continuous up to the boundary. We denote by C0n,m([0, T]×[−1,1]) the subset of Cn,m([0, T]×[−1,1]) of functions vanishing at the boundary of [−1,1], i.e. Gt(−1) =Gt(1) = 0 fort ∈ [0, T]. We denote by Ccn,m([0, T]×(−1,1)) the subset ofCn,m([0, T]×(−1,1)) of functions with compact support in [0, T]×(−1,1).

(8)

2.4. Results. We denote χ(ρ) = ρ(1−ρ), σ(ρ) = 2χ(ρ) and ∇f, resp. ∆f, the gradient, resp. the laplacian, with respect touof a functionf. ForG∈C01,2([0, T]× [−1,1]),ρ∈D([0, T],M) denote

βG(ρ, ρ0) :=

ρT, GT

− hρ0, G0i − Z T

0

dt

ρt, ∂tGt

− Z T

0

dt ρt,∆Gt

+

Z T 0

dt∇Gt(1)−ρ

Z T 0

dt∇Gt(−1)

−β 2

Z T

0 hσ(ρt),(∇Gt)· ∇(Jneum⋆ ρt)idt.

(2.8)

Denote byA[0,T]⊂D [0, T];M the set A[0,T] =n

ρ∈L2 [0, T], H1(Λ)

: ∀G∈ C01,2([0, T]×Λ), ℓβG(ρ, ρ0) = 0o . Theorem 2.1. For any sequence of initial probability measures(µN)N≥1, the se- quence of probability measures (Qβ,NµN )N≥1 is weakly relatively compact and all its converging subsequences converge to some limitQβ,∗that is concentrated on the ab- solutely continuous measures whose density ρ∈ A[0,T]. Moreover, if for any δ >0 and for any continuous function G: [−1,1]→R

N→0lim µNn 1

N X

x∈ΛN

η(x)G(x/N)− Z

Λ

γ(u)G(u)du≥δo

= 0,

for an initial profile γ : Λ → (0,1), then the sequence of probability measures (Qβ,NµN )N≥1converges to the Dirac measure concentrated on the unique weak solution of the following boundary value problem on(t, u)∈(0, T)×Λ







tρt(u) + β∇ ·n

ρt(u)(1−ρt(u))∇(Jneum⋆ ρt)(u)o

= ∆ρt(u) ρt(∓1) = ρ for 0≤t≤T ,

ρ0(u) =γ(u).

(2.9)

Remark 2.2. By weak solution of the boundary value problem (2.9) we mean ℓβG(ρ, γ) = 0 for G∈C01,2([0, T]×[−1,1]).

Theorem 2.3. There existsβ0 depending onΛandJneum so that, for anyβ < β0, for anyG∈C02([−1,1]), for any δ >0,

N→∞lim µstatN hhπN, Gi − hρ, G¯ i≥δi

= 0,

whereρ¯is the unique weak solution of the following boundary value problem



∆ρ(u)−β∇ ·n

ρ(u)(1−ρ(u))∇(Jneum⋆ ρ)(u)o

= 0, u∈Λ,

ρ(∓1) = ρ . (2.10)

We prove Theorem 2.1 and Theorem 2.3 in Section 4. Recall that the stationary measureµstatN depends onβ.

Next we state the large deviation principle associated to the law of large numbers stated in Theorem 2.1. Let JGβ =JT,G,γβ : D([0, T],M) −→ R be the functional

(9)

given by

JGβ(π) :=ℓβG(π, γ) − 1 2

Z T 0

dt

σ(πt), ∇Gt2

, (2.11)

and ˆITβ(·|γ) :D([0, T],M)→[0,+∞] the functional defined by IˆTβ(π|γ) := sup

G∈C1,20 ([0,T]×[−1,1])

JGβ(π). (2.12)

To define the large deviation rate functional, we introduce the energy functional Q:D([0, T],M)→[0,+∞] given by

Q(π) = sup

G

n Z T

0

dthπt,∇Gti −1 2

Z T 0

dthσ(πt)Gt, Gtio

, (2.13)

where the supremum is carried over allG∈Cc([0, T]×(−1,1)). From the concavity ofσ(·) it follows immediately thatQis convex and therefore lower semicontinuous.

MoreoverQ(π) is finite if and only ifπ∈L2 [0, T];H1(Λ) , and Q(π) = 1

2 Z T

0

dt Z 1

−1

du(∇πt(u))2

σ(πt(u)) · (2.14)

If (2.14) holds, then an integration by parts and Schwarz inequality imply that (2.13) is finite. The converse needs to be proven, for a proof of it we refer to [4], Subsection 4.1. The rate functionalITβ(·|γ) :D([0, T],M)→[0,∞] is given by

ITβ(π|γ) =

(IˆTβ(π|γ) if Q(π)<+∞,

+∞ otherwise . (2.15)

We show in Lemma 5.6 that ITβ(π|γ) = 0 if and only if πt(·) solves the problem (2.9) with initial datumπ0(·) =γ(·).

We have the following dynamical large deviation principle.

Theorem 2.4. Fix T > 0 and an initial profile γ in M. Consider a sequence {ηN :N ≥1} of configurations associated to γ. Then, the sequence of probability measures{Qβ,NηN :N≥1}on D([0, T],M)satisfies a large deviation principle with speedN and rate functionITβ(·|γ), defined in (2.15):

Nlim→∞

1

N logQβ,NηN πN ∈ C

≤ −inf

π∈CITβ(π|γ) lim

N→∞

1

N logQβ,NηN πN ∈ O

≥ −inf

π∈OITβ(π|γ),

for any closed setC ⊂D([0, T],M)and open setO ⊂D([0, T],M). The functional ITβ(·|γ)is lower semi-continuos and has compact level sets.

We prove Theorem 2.4 in Section 5.

3. Basic estimate

Next lemma states some properties of the potential Jneum(·,·) easily obtained by its definition.

(10)

Lemma 3.1. The potentialJneum(·,·)is a symmetric probability kernel. Moreover for any regular function G: Λ→R, we have the following:

∇ Z

Λ

Jneum(u, v)G(v)dv ≤

Z

Λ

Jneum(u, v)∇G(v)dv. (3.1) Proof. The symmetry ofJneumfollows immediately by the one ofJ. We have

Jneum(u, v) =J(0, v−u) +J(0,2−(u+v)) +J(0,2 + (u+v)),

which is symmetric in uand v. We now prove that Jneum is probabiltity kernel.

Fixu∈Λ, by a change of variables, Z

Λ

Jneum(u, v)dv =

Z (1−u)∧1 (−1−u)∨(−1)

J(0, v)dv +

Z (3−u)∧1 (1−u)∨(−1)

J(0, v)dv +

Z (−1−u)∧1 (−3−u)∨(−1)

J(0, v)dv . Suppose first thatu∈[0,1], then

Z

Λ

Jneum(u, v)dv= Z 1−u

−1

J(0, v)dv + Z 1

1−u

J(0, v)dv

= Z 1

−1

J(0, v)dv = 1.

and thus,Jneum(u,·) is a probability. The proof foru∈[−1,0] is similar. It remains to prove (3.1):

∇ Z

Λ

Jneum(u, v)G(v)dv

= Z

Λ

uJneum(u, v)G(v)dv

= Z

Λ

h

J(u, v)−J(u,2−v)−J(u,−2−v)i

∇G(v)dv . The result follows from the following inequatity,

J(u, v)−[J(u,2−v) +J(u,−2−v)]

≤Jneum(u, v)

for allu, v∈Λ.

For anyG: Λ→Randx, x+ 1∈ΛN denote by∇NG(Nx) the discrete gradient:

NG(x/N) =N

G((x+ 1)/N)−G(x/N)

. (3.2)

Next, we show that the rate CNβ of Lβ,N is a perturbation of the rate of the symmetric simple exclusion generator.

Lemma 3.2. For any x∈ΛN, with x±1∈ΛN andη∈ SN, CNβ(x, x±1;η) = 1∓β

2 η(x+ 1)−η(x)

N−1N

Jneum

⋆ π(η)

(x/N) +O(N−2). Proof. By definition ofHN, for allx, y∈ΛN andη∈ SN,

x,yHN

(η) = 1

N η(x)−η(y)2

Jneum(x N, y

N)−Jneum(0,0) + η(x)−η(y)1

N X

z∈ΛN

η(z)

Jneum(x N, z

N)−Jneum(y N, z

N) .

This concludes the proof.

(11)

We start recalling the definitions of relative entropy and Dirichlet form, that are the main tools in the [11] approach. Let h: Λ→ (0,1) and νNh(·) be the product Bernoulli measure defined in (2.7). Givenµ, a probability measure onSN, denote byH(µ|νNh(·)) the relative entropy ofµwith respect to νNh(·):

H(µ|νNh(·)) = sup

f

n Z f(η)µ(dη)−log Z

ef(η)νNh(·)(dη)o ,

where the supremum is carried over all bounded functions onSN. SinceνNh(·)gives a positive probability to each configuration,µis absolutely continuous with respect toνNh(·)and we have an explicit formula for the entropy:

H(µ|νh(·)N ) = Z

logn dµ dνNh(·)

o

dµ . (3.3)

Further, since there is at most one particle per site, there exists a constantC, that depends only onh(·), such that

H(µ|νNh(·)) ≤ CN (3.4)

for all probability measuresµonSN (cf. comments following Remark V.5.6 in [15]).

3.1. Dirichlet form estimates. One of the main step for deriving the hydro- dynamic limit and the large deviations, is a super exponential estimate which al- lows the replacement of local functions by functionals of the empirical density.

One needs to estimate expression such as hZ, fiµN in terms of Dirichlet form N2h−LNp

f(η),p

f(η)iµN, where Z is a local function and h·,·iµN represents the inner product with respect to some probability measure µN. In the context of boundary driven process, the fact that the invariant measure is not explicitly known introduces a technical difficulty. We fix as reference measure a product measure νNθ(·), see (2.7), where θ is a smooth function with the only requirement that θ(∓1) = ρ. There is therefore no reasons for N2h−LNp

f(η),p

f(η)iνθ(·) to be positive. Next lemma estimates this quantity.

Define the following functionals fromh∈L2(ν) toR+: D0,N h, ν

=

NX−1 x=−N

Z

h(ηx,x+1)−h(η)2

dν(η),

D+,N h, ν

=1 2

Z

c+ η(N)

h(σN−1η)−h(η)2

dν(η),

D−,N h, ν

=1 2

Z

c η(−N) h(σ−N+1η)−h(η)2

dν(η).

(3.5)

Lemma 3.3. Let θ : Λ → (0,1) be a smooth function such that θ(∓1) = ρ. There exists a positive constant C0 ≡ C0(k∇θk) so that for any a > 0 and for f ∈L2 νNθ(·)

, Z

SN

f(η)Lβ,Nf(η)dνNθ(·)(η)≤ − 1−a

D0,N f, νNθ(·) +C0

a N−1kfk2L2Nθ(·)), (3.6) Z

SN

f(η)L±,Nf(η)dνNθ(·)(η) =−D±,N f, νNθ(·)

. (3.7)

(12)

Proof. The proof of (3.7) follows from the reversibility of the Bernoulli measure νNθ(·)with respect to L±,N. Next, we show (3.6). By Lemma 3.2,

Z

SN

f(η)Lβ,Nf(η)dνNθ(·)(η)≤

NX−1 x=−N

Z h ∇x,x+1f (η)i

f(η)dνNθ(·)(η)

+A1

N

NX−1 x=−N

Z ∇x,x+1f

(η)f(η)dνNθ(·)(η)

(3.8)

for some positive constantA1depending only onβ andJ. We write the first term of the right hand side of (3.8) as

NX−1 x=−N

Z h ∇x,x+1f (η)i2

Nθ(·)(η)

+

N−1X

x=−N

Z

RN(x, x+ 1;θ, η)h

x,x+1f (η)i

f(ηx,x+1)dνθ(·)N (η),

(3.9)

where

RN(x, x+ 1;θ, η) =

1−e−N−1Nλ(θ(x/N)) x,x+1η(x)

, (3.10)

λis the chemical potential defined by

λ(r) = log [r/(1−r)] (3.11)

and∇N stands for the discrete derivative defined in (3.2). By the inequality for allA, B∈R and a >0, AB≤ a

2A2+ 1

2aB2 (3.12) and Taylor expansion, the formula (3.9) is bounded by

− 1−a 2

N−1X

x=−N

Z h ∇x,x+1f (η)i2

Nθ(·)(η) +A2

a N−1kfk2L2Nθ(·)) (3.13) for alla >0. HereA2 is a positive constant.

The second term on the right hand side of (3.8) is handled in the identical way.

It is bounded by a 2

NX−1 x=−N

Z h ∇x,x+1f (η)i2

Nθ(·)(η) +A3

a N−1kfk2L2θ(·)N ). (3.14) The lemma follows from (3.8),(3.9), (3.13) and (3.14).

Denote forh∈L2(ν) Dβ,N h, ν

=

NX−1 x=−N

Z

CNβ(x, x+ 1;η) h(ηx,x+1)−h(η)2

dν(η).

Lemma 3.4. There exists a positive constant C1 = C1(β, J) such that, for any measureν and forh∈L2(ν),

1−C1

N

D0,N h, ν

≤ Dβ,N h, ν

≤ 1 + C1

N

D0,N h, ν .

Proof. The proof is elementary sinceCNβ(x, x+ 1, η)−1is uniformly bounded in

N,xandη.

(13)

Lemma 3.5. Letρ, ρ0: Λ→(0,1)be two smooth functions. There exists a positive constant C0 ≡C0(k∇ρ0k,k∇ρk) such that for any probability measure µN on SN,

D0,Ns dµN

Nρ(·), νNρ(·)

≤ 2D0,Ns dµN

Nρ0(·), νNρ0(·)

+ C0N−1. (3.15)

Proof. Denote byf(η) = N

ρ(·)N (η) andh(η) = N

Nρ0 (·)(η). Sincef(η) =h(η)

ρ0 (·) N (η) Nρ(·)(η)

we obtain for−N ≤x≤N−1 the following Z

SN

h∇x,x+1p f(η)i2

Nρ(·)(η)

= Z

SN

h√

h(ηx,x+1)R2(x, x+ 1;η) +∇x,x+1√ h(η)i2

Nρ0(·)(η)

≤2 Z

SN

h∇x,x+1√ h(η)i2

Nρ0(·)(η) + 2

Z

SN

h(ηx,x+1)

RN(x, x+ 1;ρ, η)2

Nρ0(·)(η), where

R2(x, x+ 1;η) = exp

(1/2)N−1N[λ(ρ(x/N))−λ(ρ0(x/N))]∇x,x+1η(x) −1 and λ is the chemical potential defined by (3.11). We conclude the proof using

Taylor expansion and integration by parts.

3.2. Superexponential estimates. For a positive integerℓandx∈ΛN denote Λ(x) = ΛN,ℓ(x) ={y∈ΛN : |y−x| ≤ℓ}.

Whenx= 0, we shall denote Λ(0) simply by Λ, that is, for all 1≤ℓ≤N, Λ≡ΛN,ℓ(0) ={−ℓ,· · ·, ℓ}.

Denote the empirical mean density on the box Λ(x) byη(x):

η(x) = 1

(x)| X

y∈Λ(x)

η(y). (3.16)

For a cylinder function Ψ, that is a function on {0,1}Z depending onη(x), x∈Z, only trough finitely manyx, denote byΨ(ρ) the expectation of Ψ with respect toe νρ, the Bernoulli product measure with density ρ:

Ψ(ρ) =e Eνρ[Ψ]. (3.17)

Further, denote forG∈ C([0, T]×[−1,1]) and ε >0 VN,εG,Ψ(s, η) = 1

N X

x∈ΛN

Gs(x/N)h

τxΨ(η)−Ψ(ηe [εN](x))i

, (3.18)

where the sum is carried over allxsuch that the support ofτxΨ belongs to ΛN and [·] denotes the lower integer part.

(14)

Proposition 3.6. Let {µN :N ≥1} be a sequence of probability measures onSN. For everyδ >0,

ε→0lim lim

N→∞

1

N logPβ,Nµ

N

h Z T

0

VN,εG,Ψ(s, ηs)ds> δi

= −∞.

Proof. Fixc > 0 that will decreases to 0 afterε and a smooth function ρc: Λ→ (0,1) which is constant in Λ(1−c) = [−1 +c,1−c] and equal toρ±at the boundary, i.e ρc(±1) = ρ±. The constant can be arbitrarily chosen and we denote it γ0. Divide ΛN in two subsets, Λ[(1−2c)N] and ΛN[(1−2c)N] and split the sum overx in the definition ofVN,εG,Ψinto the sum over these two sets. Since

sup

η,ε,N,x∈ΛN

n

Gs(x/N)h

τxΨ(η)−Ψ(ηe [εN](x))i o

<∞, we have that

1 N

X

x∈ΛN[(1−2c)N]

Z T 0

dsGs(x/N)h

τxΨ(η)−Ψ(ηe [εN](x))i≤cT K0 (3.19) for some positive constant K0 which depends on G. By Chebyshev exponential inequality, for alla >0,

1

N logPβ,N

µN

h Z T

0

VN,εG,Ψ(s, ηs)ds> δi

≤ −a δ−T K0c + 1

N logPβ,Nµ

N

hexp aN Z T

0

Vc,G,Ψ

N,ε (s, ηs)ds

!i , (3.20) where

Vc,G,Ψ

N,ε (s, η) = 1 N

X

x∈Λ[(1−2c)N]

Gs(x/N)h

τxΨ(η)−Ψ(ηe εN(x))i

. (3.21) It is immediate to see that the Radon-Nikodym derivative

dPβ,Nµ

N

dPβ,N

νNρc(·)

t)t∈[0,T]

= dµN

Nρc(·) ≤eN K1(c)

for some positiveK1(c) that depends onc. The right hand side of (3.20) is bounded by

−a δ−T K0c

+ K1(c) + 1

N logPβ,N

νNρc(·)

h

exp aN Z T

0

Vc,G,ΨN,ε (s, ηs)ds

!i . (3.22) Sincee|x|≤ex+e−xand

limN−1log{aN +bN} ≤max{limN−1logaN , limN−1logbN}, (3.23) we may remove the absolute value in the third term of (3.22), provided our estimates remain in force if we replaceGby−G. Denote by

(LN)s=1

2(LN +LN)

(15)

whereLN is the adjoint ofLN inL2Nρc(·)). By the Feynman-Kac formula, 1

N logPβ,N

νρc(·)N

hexp aN Z T

0

Vc,G,Ψ

N,ε (s, ηs)ds

!i

≤ 1 N

Z T 0

λN,ε(Gs)ds , (3.24) whereλN,ε(Gs) is the largest eigenvalue of{N2(LN)s+N aVc,G,Ψ

N,ε (Gs,·)}. By the variational formula for the largest eigenvalue, for eachs∈[0, T],

1

N,ε(Gs) = sup

f

n Z aVc,G,Ψ

N,ε Gs, η

f(η)νNρc(·)(dη) + NhLNp f ,p

fiνρcN(·)o . (3.25) In this formula the supremum is carried over all densitiesf with respect toνρNc(·). By Lemma 3.3, since (3.7) gives a negative contribution, it is enough, to get the result, to choosec such thatc < T Kδ0 and to show that, there existsM > 0 that depends only onGandc, such that, for alla >0

ε→0lim lim

N→∞sup

f

n Z

aVc,G,Ψ

N,ε Gs, η

f(η)νNρc(·)(dη) − ND0,N(p

f , νNρc(·))o

≤ M . We then let a ↑ ∞. Notice that for N large enough the function Vc,G,Ψ

N,ε Gs, η depends on the configuration η only through the variables {η(x), x ∈ Λ(1−c)N}. Since ρc is equal to γ0, in Λc, we replace νNρc(·) in the previous formula by νNγ0 with respect to which the operatorL0,N is reversible. ThereforeD0,N(·, νNγ0) is the Dirichlet form associated to the generatorL0,N. Since the Dirichlet form is convex, it remains to show that

ε→0lim lim

N→∞sup

f

n Z aVc,G,Ψ

N,ε Gs, η

f(η)νNγ0(dη) − ND0,N(p

f , νNγ0(·))o

= 0, for anya >0. This follows from the usual one block and two blocks estimates (cf.

Chap 5 of [15]).

Forx=±N, a configurationηand ℓ≥1, let WN±,ℓ(η) =η(±N)−ρ±

, (3.26)

where, see (3.16),η(N) =ℓ+11

η(N−ℓ)+· · ·+η(N) andη(−N) =ℓ+11

η(−N+ ℓ) +· · ·+η(−N) .

Proposition 3.7. Fix a sequence {µN : N ≥ 1} of probability measures on SN. For everyδ >0,

ℓ→∞lim lim

N→∞

1

N logPβ,Nµ

N

h Z T

0

WN±,ℓs)ds > δi

= −∞.

Proof. Consider first the limit with the term WN+,ℓ. Fix a smooth function γ : Λ→(0,1) such that γ(−1) =ρ, andγ(u) =ρ+ for u∈[0,1]. Since the Radon- Nikodym derivative N

Nγ(·) is bounded by exp(N K1) for some positive constantK1, it is enough to show that

ℓ→∞lim lim

N→∞

1

N logPβ,N

νNγ(·)

h Z T

0

WN+,ℓs)ds > δi

= −∞.

(16)

We follow the same steps as in Proposition 3.6. Applying Chebyshev exponential inequality and Feynman-Kac formula, the expression in the last limit is bounded for alla >0 by

−aδ + T

N eλN,ε(a), (3.27)

where for alla >0,N1N,ε(a) is the largest eigenvalue of theνNγ(·)-reversible operator f →N(LN)s(f) +a

WN+,ℓ f.

Here (LN)sis the symmetric part of the operatorLN inL2Nγ(·)). By the variational formula for the largest eigenvalue, we have

1

NeλN,ε(a) = sup

f

n Z

aWN+,ℓ(η)f(η)νNγ(·)(dη) + N <LNp

f ,p f >νγ(·)

N

o .

In this formula the supremum is carried over all densitiesf with respect toνNγ(·). By Lemma 3.3, we just need to show that, there exists M >0, such that, for all a >0

ℓ→∞lim lim

N→∞sup

f

n Z

aWN+,ℓ(η)f(η)νNγ(·)(dη)

− ND0,N(p

f , νNγ(·))−ND+,N(p

f , νNγ(·))o

≤ M . Recall that the profile γ is constant equal to ρ+ on [0,1]. SinceWN+,ℓ(η) depends only on coordinates in a box Λ(N), we replaceνNγ(·)in the previous formula byνρN+. On the other hand,νNρ+ is reversible forL0,N+L+,N and thereforeD0,N(·, νNρ+) + D+,N(·, νNρ+) is the Dirichlet form associated to the generator L0,N +L+,N. Since the Dirichlet form is convex, it remains to show that

ℓ→∞lim lim

N→∞sup

f

n Z aWN+,ℓ(η)f(η)νNρ+(dη)

− ND0,N(p

f , νNρ+)−ND+,N(p

f , νNρ+)o

= 0.

for any a >0. This follows from the law of large numbers by applying the same device used in the proof of the one block and two blocks estimates, (cf. Chap 5 of

[15], and Lemma 3.12, Lemma 3.13 in [20]).

3.3. Energy estimate. We prove in this subsection an energy estimate which is one of the main ingredient in the proof of large deviations and hydrodynamic limit.

It allows to prove Lemma 4.3 and to exclude paths with infinite energy in the large deviation regime. Forδ >0,G∈Cc([0, T]×Λ) define

δG(π) = Z T

0

dthπt,∇Gti −δ Z T

0

dthσ(πt)Gt, Gti, (3.28) Q˜δ(π) = sup

G∈Cc([0,T]×Λ)

nQ˜δG(π)o

. (3.29)

Notice that

δ(π) = 1 2δQ(π), whereQ(·) is defined in (2.13).

(17)

For a functionminM, letmε: Λ→R+ be given by mε(u) = 1

2ε Z

[u−ε,u+ε]∩Λ

m(v)dv .

Whenu∈[−1 +ε,1−ε],mε(u) = (m∗ιε)(u), whereιεis the approximation of the identity defined by

ιε(u) = 1

11{[−ε, ε]}(u).

Lemma 3.8. There exists a positive constantC1depending only on ρ± so that for any given δ0 >0, for any δ,0≤δ≤δ0, for any sequence{ηN ∈ SN :N ≥1}and for anyG∈ Cc([0, T]×Λ), we have

ε→0lim lim

N→∞

1

N logPβ,N

ηN

h

exp δ NQ˜δG0N ∗ιε)i

≤ C1(T+ 1).

Proof. Assume without loss of generality thatεis small enough so that the support of G(·,·) is contained in [0, T]×[−1 +ε,1−ε]. Let θ : Λ → (0,1) be a smooth function such that θ(∓1) = ρ. Since νNθ(·)N) ≥ exp{−C1N} for some finite constant C1 depending only on θ, it is enough to prove the lemma with Pβ,N

νθ(·)N in place ofPβ,N

ηN .

Set Ψ1(η) = [η(1)−η(0)]2 and note thatΨe1(a) =Eνa1] =σ(a) = 2a(1−a), where νa is the Bernoulli measure with parametera ∈ [0,1]. Denote BG,ΨN,ε,δ10 the set of trajectories (ηt)t∈[0,T] so that

BN,ε,δG,Ψ10 = n

η·∈D([0, T],SN) :

Z T 0

VN,εG21(t, ηt)dt ≤ 1

δ20 o

,

where VN,εG21 is defined in (3.18). By (3.23) and Proposition 3.6, it is enough to show

ε→0lim lim

N→∞

1

N logPβ,N

νNθ(·)

h

e δ NQ˜δG0N∗ιε)

11{BG,ΨN,ε,δ10}i

≤ C1(T+ 1). Recalling the definition ˜QδG, see (3.28), we have

Z T 0

dthπNt ∗ιε,∇Gi= Z T

0

dt

NX−1 x=−N+1

t(x)−ηt(x+ 1)}Gt(x/N) +OG(ε).

Further on the setBN,ε,δG,Ψ10 δ0

Z T 0

dthσ(πtN ∗ιε), G2ti ≥δ0

Z T 0

dt 1 N

N−1X

x=−N+1

G2t(x/N)τxΨ1t)

−δ0OG2(N, ε)− 1 δ0

,

where OG(ε) is absolutely bounded by a constant which vanishes as ε ↓ 0 and OG2(N, ε) is is absolutely bounded by a constant which vanishes asN ↑ ∞. There- fore to conclude the proof it is enough to show that

Nlim→∞

1

N logPβ,N

νNθ(·)

hexp N

Z T 0

dt VGδ(t, ηt)i

≤ C1T (3.30)

Références

Documents relatifs

In this section, we consider the perturbation of the original process (2.2), and we prove some results needed either to caracterize the behavior of the empirical current and

The rate functional associated with the large deviation principle depends on the statistical properties of the external random field, it is lower semicontinuous with compact

This paper is aimed to derive large deviations for statistics of Jacobi process already conjectured by M.. Zani in

Probability Laws Related To The Jacobi Theta and Riemann Zeta Functions, and Brownian Excursions.. Integrals and

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des

Key-words: Kinetic equations, Hamilton-Jacobi equations, large deviations, perturbed test func- tion method, Piecewise Deterministic Markov Process..

76–79 An intriguing prop- erty of those systems is the apparent paradox between the unique microscopic fixed point of the dynamics and the potentially numer- ous attractors that

This paper is the second of a series of three in which we address the compu- tation of the large deviation Hamiltonian H, and of the large deviation parameter ε, for the three