• Aucun résultat trouvé

On viscosity and fluctuation-dissipation in exclusion processes

N/A
N/A
Protected

Academic year: 2021

Partager "On viscosity and fluctuation-dissipation in exclusion processes"

Copied!
26
0
0

Texte intégral

(1)

EXCLUSION PROCESSES

C. LANDIM, S. OLLA, S. R. S. VARADHAN Dedicated to Gianni Jona Lasinio on his seventieth birthday.

Abstract. We consider the asymmetric simple exclusion process. We review some results in dimension d ≥ 3 concerning the fluctuation-dissipation theorem and we prove regularity of viscosity coefficients.

1. Introduction

In this paper we will review some of the results concerning the fluctuations in equilibrium of the asymmetric simple exclusion process in dimension d ≥ 3. These are examples of nonreversible models of interacting particle systems on Zd with

conservation of particles and a family of ergodic equilibrium distributions indexed by a single parameter, i.e. the density. The model is simple enough to admit a certain amount of explicit calculations. In particular, the equilibrium distributions {να} are Bernoulli product measures with να[η(x) = 1] = α for 0 ≤ α ≤ 1.

The study of fluctuations depends on controlling space-time correlations of the form Eνα   Z T 0 X z∈V τzf (ηt) dt 2 

for large times T and large volumes V . Here f is a suitable local function and τz

is translation by z ∈ Zd. The natural way to control such objects is to invert the

generator of the process L, i.e. to solve the equation Lu =X

z∈V

τzf (1.1)

perhaps with u =P

z∈V τzv for some v. Our goal is to show that we can

approxi-mate, in a proper weak sense, the solution of equation (1.1) by such functions. Our interest is to perform such approximation in a diffusive space-time scaling limit (i.e. T ∼ |V |2/d). We will characterize the functions f such that the corresponding

asymptotic space-time variance is finite and then we will prove that this variance is a smooth function of the density α.

There is a natural orthonormal basis for L2(να) indexed by finite subsets A of Zd

that commutes with τz. Moreover there is a corresponding decomposition of L2(να)

as ⊕n≥0Gn as the direct sum of orthogonal subspaces according to the cardinality

n of A. Since, in this context, one cannot distinguish between f and τzf , A and Date: February 12, 2003.

2000 Mathematics Subject Classification. primary 60K35.

Key words and phrases. Exclusion process, fluctuation-dissipation theorem, regularity of vis-cosity coefficients.

(2)

τzA can be identified. A reasonable cross section, one with a multiplicity of n for

sets of cardinality n can be found for the equivalence classes. Another important point is that that in the symmetric case L leaves each Gn invariant so that the

inversion needs to be done only on each Gn, which is no harder than the analysis of

random walk of n − 1 particles on Zd. If d ≥ 3, the asymmetric case can be treated

as a perturbation of the symmetric one. Although this is a somewhat singular perturbation the transience of the random walk in d ≥ 3 and the fact that the perturbation only causes Gn to be mapped into Gn−1⊕ Gn⊕ Gn+1help the analysis.

The perturbation is controlled with the use of weighted norms with weights growing polynomially in the degree n of Gn.

Duality for the symmetric simple exclusion was first observed by Spitzer and was broadly used in this context (cf. [12]). The use of a dual base in order to study fluctuations in the asymmetric simple exclusion in dimension d ≥ 3 was introduced in [8] (were the fluctuation-dissipation theorem was first proved) and in [13]. The fluctuation dissipation theorem was then applied in order to study the diffusive incompressible limit (cf. [3]), the first order corrections to the hydrodynamic limit (cf. [6]) and the equilibrium fluctuations for the density field [2].

Always with the use of this dual base, in [11] we prove the regularity of the self-diffusion coefficient (as function of the density) for the symmetric simple exclusion (reversible). Using this approach in [1] is proved the regularity of the bulk diffusion coefficient for a non-gradient speed change exclusion that has να as equilibrium

(reversible) measures.

2. Notation and Results

Fix a probability distribution p(·) supported on a finite subset of Zd∗ = Zd\{0}

and denote by L the generator of the simple exclusion process on Zd associated to

p(·). L acts on local functions f on X = {0, 1}Zd as

(Lf )(η) = X

x,y∈Zd

p(y − x)η(x){1 − η(y)}[f (σx,yη) − f (η)] , (2.1) where σx,yη stands for the configuration obtained from η by exchanging the occu-pation variables η(x), η(y):

(σx,yη)(z) =    η(z) if z 6= x, y , η(x) if z = y , η(y) if z = x .

Denote by s(·) and a(·) the symmetric and the anti-symmetric parts of the proba-bility p(·):

s(x) = (1/2)[p(x) + p(−x)] , a(x) = (1/2)[p(x) − p(−x)]

and denote by Ls, La the symmetric part and the anti-symmetric part of the

generator L. Ls, La are obtained by replacing p by s, a in the definition of L. For α in [0, 1], denote by ναthe Bernoulli product measure on X with να[η(x) =

1] = α. Measures in this one-parameter family are stationary and ergodic for the simple exclusion dynamics and in the symmetric case, i.e. p(x) = p(−x), these measures are reversible. Expectation with respect to να is represented by < · >α

and the scalar product in L2

(3)

The main result is Theorem 5.3, known as the fluctuation-dissipation theorem. There are two distinct types of fluctuations. If u is a local function, then fluctuations of the form Z t−2 0 f(ηs)ds where f(η) = X x G(x)(τxLu)(η)

satisfy a central limit theorem and converge to a Brownian Motion when scaled by 1+d2.

On the other hand if v =P

zb(z)[η(z) − η(0)] for some b supported on a finite

subset of Zd ∗, then g = X x G(x)(τxv)(η) = X x,z G(x)b(z)[η(x + z) − η(x)] = X z b(z)X x [G((x − z)) − G(x)]η(x) = −X z X x b(z)z · ∇G(x) η(x) . Consequently fluctuations of the form

Z t−2

0

g(ηs)ds

when normalized again by 1+d

2 correspond to a time average of a density

fluctua-tion.

The fluctuation dissipation theorem asserts that if w ∈ ⊕n≥2Gnis a local function

then there are coefficients b(z, w) such that the fluctuation of w −P

zb(z, w)[η(z) −

η(0)] is well approximated by fluctuations of the first type. The constants b(z, w) are defined for each α. We prove in section 7 that these are C∞functions of α.

This leads immediately to the Ornstein-Uhlenbeck limit for the rescaled density fluctuations in equilibrium. If ξ(t) is a rescaled density fluctuation at time t, then

we get

dξ(t) = A(t)dt + dM(t)

The A decomposes into two terms, one of them being a density fluctuation of the

form Aξ(t) which is linear in the field and the other which is nonlinear

approxi-mating a Brownian Motion. The term dM(t) approaches a Brownian Motion as

well. The covariances can all be worked out resulting in an Ornstein-Uhlenbeck type of relation

dξ(t) = Aξ(t)dt + dβ(t)

In practice A turns out to be second order elliptic partial differential operator and its coefficients are the symmetrized viscosity coefficients that depend smoothly on the underlying density α (cf. [2] for details).

3. Duality

We examine in this section the action of the symmetric part Lsof the generator

(4)

Fix once for all a density α in (0, 1). All expectations in this section are taken with respect to ναand we omit all sub-indices.

3.1. The dual space. For each n ≥ 0, denote by En the subsets of Zd with n

points and let E = ∪n≥0En be the class of finite subsets of Zd. For each A in E , let

ΨAbe the local function

ΨA =

Y

x∈A

η(x) − α pχ(α) ,

where χ(α) = α(1−α). By convention, Ψφ= 1. It is easy to check that {ΨA, A ∈ E }

is an orthonormal basis of L2

α). For each n ≥ 0, denote by Gn the subspace of

L2

α) generated by {ΨA, A ∈ En}, so that L2(να) = ⊕n≥0Gn. Functions in Gnare

said to have degree n.

Consider a local function f . Since {ΨA : A ∈ E } is a basis of L2(να), we may

write f =X n≥0 X A∈En f(A)ΨA.

Note that the coefficients f(A) depend not only on f but also on the density α: f(A) = f(A, α). Since f is a local function, f : E → R is a function of finite support.

For local functions u, v, define the scalar product  · , ·  by  u, v  = X

x∈Zd

{ < τxu, v > − < u >< v > } , (3.1)

where {τx, x ∈ Zd} is the group of translations. That this is in fact an inner product

can be seen by the relation  u, v  = lim V ↑Zd 1 |V | < X x∈V τx(u− < u >), X x∈V τx(v− < v >) > .

Since  u − τxu , v = 0 for all x in Zd, this scalar product is only semidefinite.

Denote by L2

·,·(να) the Hilbert space generated by the local functions and the

inner product  ·, · . The scalar product of two local functions u, v can be written in terms of the Fourier coefficients of u, v through a simple formula. To this end, fix two local functions u, v and write them in the basis {ΨA, A ∈ E }:

u = X A∈E u(A)ΨA, v = X A∈E v(A)ΨA.

An elementary computation shows that  u, v  = X x∈Zd X n≥1 X A∈En u(A) v(A + x) .

In this formula, B + z is the set {x + z; x ∈ B}. The summation starts from n = 1 due to the centering by the mean in the definition of the inner product  ·, · .

We say that two finite subsets A, B of Zdare equivalent if one is the translation

of the other. This equivalence relation is denoted by ∼ so that A ∼ B if A = B + x for some x in Zd. Let ˜E

n be the quotient of En with respect to this equivalence

relation: ˜En= En/∼, ˜E = E/∼. If, for some n, f : En→ R is a summable function,

X A∈En f(A) = X ˜ A∈ ˜En ˜f( ˜A)

(5)

where, for any equivalence class ˜A and a summable function f : E → R, ˜f( ˜A) = X

z∈Zd

f(A + z) , (3.2)

A being any representative from ˜A. In particular, for two local functions u, v,  u, v  = X x,z∈Zd X n≥1 X ˜ A∈ ˜En u(A + z) v(A + x + z) = X n≥1 X ˜ A∈ ˜En ˜ u( ˜A) ˜v( ˜A) .

We say that a function f : E → R is translation invariant if f(A + x) = f(A) for all sets A in E and all sites x of Zd. Of course, functions ˜f on ˜E are the same as translation invariant functions on E . Fix a subset A of Zdwith n points. There are n sets in the class of equivalence of A that contain the origin. Therefore, summing a translation invariant function f over all equivalence classes ˜A in ˜En is the same as

summing f over all sets B in En which contain the origin and dividing by n:

X ˜ A∈ ˜En f( ˜A) = 1 n X A∈En A30 f(A)

provided f(A) = f(A + x) for all A, x. Let E∗ be the class of all finite subsets of

Zd∗= Zd\{0} and let E∗,n be the class of all subsets of Zd∗ with n points. Then, we

may write  u, v  = X n≥1 1 n X A∈En A30 ˜ u(A) ˜v(A) = X n≥0 1 n + 1 X A∈E∗,n ˜ u(A ∪ {0}) ˜v(A ∪ {0}) .

In conclusion, if for a finitely supported function f : E → R, we define Tf : E∗→ R

by (Tf)(A) = ˜f(A ∪ {0}) = X z∈Zd f([A ∪ {0}] + z) , we have that  u, v  = X n≥0 1 n + 1 X A∈E∗,n Tu(A) Tv(A) . (3.3)

For n ≥ 0, denote by πn the projection that corresponds to E∗,n i.e. (πnf)(A) =

f(A)1{A ∈ E∗,n} and denote by < · , · > the usual scalar product on each set E∗,n:

for f, g : E∗,n→ R,

< f, g > = X

A∈E∗,n

f(A) g(A) .

In view of formula (3.3), it is natural to introduce, for an integer k ≥ −1, the Hilbert spaces L2(E∗, k) generated by finite supported functions f : E∗→ R and the

scalar product  · , · 0,kdefined by

 f, g 0,k=

X

n≥0

(n + 1)k< πnf, πng> .

With this notation, for local functions f , g in ⊕n≥1Gn,

(6)

To summarize some observations on the transformation T, we need some no-tation. For a subset A of Zd∗ and x, y, z in Zd∗, denote by Ax,y the set defined

by Ax,y =    (A\{x}) ∪ {y} if x ∈ A, y 6∈ A, (A\{y}) ∪ {x} if y ∈ A, x 6∈ A, A otherwise ; (3.4) and denote by SzA the set defined by

SzA =



A − z if z 6∈ A, (A − z)0,−z if z ∈ A .

(3.5) Therefore, to obtain SzA from A in the case where z belongs to A, we first translate

A by −z, getting a new set which contains the origin, and we then remove the origin and add site −z.

Remark 3.1.

(a) The restriction of f to E1 is irrelevant for the definition of Tf(A) if A is

nonempty.

(b) Not every function f∗: E∗→ R is the image by T of some function f : E → R

since

(Tf)(A) = (Tf)(SzA) (3.6)

for all z in A.

(c) Let f∗ : E∗ → R be a finitely supported function satisfying (3.6): f∗(A) =

f∗(SzA) for all z in A. Define f : E → R by

f(B) = 

|B|−1f

∗(B \ {0}) if B 3 0 ,

0 otherwise . (3.7)

An elementary computations shows that Tf = f∗. With this choice, which

is natural but not unique, f(0) = f∗(φ).

(d) T maps En into E∗,n−1 lowering the degree of a function by one. Thus the

translations in the inner product  ·, ·  effectively reduce the degree by one while replacing the space Zd by Zd∗.

(e) Formula (3.3) shows also that a local function f is in the kernel of the inner product  ·, ·  if and only if Tf vanishes, i.e., if and only if

X

x∈Zd

f(A + x) = 0

for all finite subsets A such that |A| ≥ 1. Examples of such functions are the constants or the difference of the translations of a local function: τyf − τxf .

To keep notation simple, most of the time, for a function f : E → R, we denote Tfby ¯f. Real functions on E or on E∗ are indistinctively denoted by the symbols f,

g.

3.2. Some Hilbert spaces. We investigate in this subsection the action of the symmetric part of the generator L on the basis {ΨA, A ∈ E∗}. Fix a function

u of degree n ≥ 1 and denote by u its Fourier coefficients. A straightforward computation shows that

Lsu = X

A∈En

(7)

where Lsis the generator of finite symmetric random walks evolving with exclusion

on Zd:

(Lsu)(A) = (1/2)

X

x,y∈Zd

s(y − x)[u(Ax,y) − u(A)] (3.9)

and Ax,y is the set defined by (3.4). Furthermore, an elementary computation,

based on the fact that X

z∈Zd

f([B ∪ {y}] + z) = Tf(SyB)

for all subsets B of Zd

∗, sites y not in B and finitely supported functions f : E → R,

shows that for every set B in E∗

TLsu(B) = LsTu(B) , (3.10) where (Lsu) (B) = (1/2) X x,y∈Zd ∗

s(y − x)[u(Bx,y) − u(B)] +

X

y6∈B

s(y)[u(SyB) − u(B)] .

This computation should be understood as follows. We introduced an equiv-alence relation in E when we decided not to distinguish between a set and its translations. This is the same as assuming that all sets contain the origin. If n particles evolve as exclusion random walks on Zd, one of them fixed to be at the origin, two things may happen. Either one of the particles which is not at the origin jumps or the particle we assumed to be at the origin jumps. In the first case, this is just a jump on Zd∗ and is taken care by the first piece of the generator Ls. In

the second case, however, since we are imposing the origin to be always occupied, we need to translate back the configuration to the origin. This part corresponds to the second piece of the generator Ls.

We are now in a position to define the Hilbert space induced by the local functions C, the symmetric part of the generator L and the scalar product  ·, · . For two local functions u, v, let

 u, v 1=  u, (−Ls)v 

and let H1= H1(C, Ls,  ·, · ) be the Hilbert space generated by local functions

f and the inner product  ·, · 1. By (3.8), (3.3) and (3.10) the previous scalar

product is equal to −X n≥0 1 n + 1 X A∈E∗,n u(A) Lsv(A) = − X n≥0 1 n + 1 X A∈E∗,n u(A) (Lsv)(A) = X n≥0 1 n + 1 < πnu, (−Ls)πnv> because Ls keeps the degree of the functions mapping E∗,n in E∗,n.

This formula leads to the following definitions. For each n ≥ 0, denote by < ·, · >1the scalar product on E∗,n defined by

< f, g >1= < f, (−Ls)g >

and denote by H1(E∗,n) the Hilbert space on E∗,ninduced by the finitely supported

functions and the scalar product < ·, · >1. The associated norm is denoted by

kfk2

(8)

H1(E∗, Ls, k) the Hilbert space induced by the finite supported functions f, g : E∗→

R and scalar product

 f, g 1,k=  f, (−Ls)g 0,k=

X

n≥0

(n + 1)k< πnf, (−Ls)πng> .

The associated norm is denoted by k · k1,k so that kfk21,k= f, f 1,k.

Three observations are in order. First of all, in the definition of the scalar product  ·, · 1,k, because Lsf(φ) = 0 for all f, it is irrelevant if the summation starts

from n = 0 or from n = 1. On the other hand, it follows from the previous relation that kfk2 1,k = X n≥0 (n + 1)kkπnfk21.

Finally, for every local function u, v,

 u, v 1=  Tu, Tv 1,−1

vanishes for functions u and v of degree 0 or 1. Moreover, for every n ≥ 1 and every finitely supported functions f, g : E∗,n→ R,

< f, g >1 = 1 4 X x,y6=0 s(y − x) X A∈E∗,n

[g(Ax,y) − g(A)] [f(Ax,y) − f(A)]

+ 1 2 X y∈Zd s(y) X A∈E∗,n A63y [g(SyA) − g(A)] [f(SyA) − f(A)] .

To introduce the dual Hilbert spaces of H1, H1, for a local function u, consider

the semi-norm k · k−1 given by

kuk−1 = sup v n 2  u, v  −  v, v 1 o ,

where the supremum is taken over all local functions v. Denote by H−1 = H−1(C,

Ls,  ·, · ) the Hilbert space generated by the local functions and the semi-norm k · k−1.

Recall the definition of the spaces Gn introduced at the beginning of subsection

3.1. Since Lskeeps the degree of a function and since the spaces Gnare orthogonal,

for local functions of fixed degree, we may restrict the supremum to local functions of the same degree so that

kf k2−1 = X

n≥1

kπnf k2−1,

where πnf stands for the projection of f on Gn. Moreover, we will see later in (4.1)

that functions of degree 1 which do not vanish in L2( ·, · ) have infinite H −1

norm.

In the same way, for an integer n ≥ 1 and a finitely supported function u : E∗,n→

R, let kuk2−1 = sup v n 2 < u, v > − < v, v >1 o ,

where the supremum is carried over all finitely supported functions v : E∗,n → R.

Denote by H−1 = H−1(E∗,n) the Hilbert space induced by the finitely supported

(9)

For a integer k ≥ −1, define the H−1,k = H−1(E∗, Ls, k) norm of a finite sup-ported function u : E+→ R by kuk−1,k = sup v n 2  u, v 0,k−  v, (−Ls)v 0,k o ,

where the supremum is carried over all finitely supported functions v : E∗ → R.

Denote by H−1,k = H−1(E∗, Ls, k) the Hilbert space induced by this semi-norm and

the space of finite supported functions. Here again, since Ls does not change the

degrees of a function, for every finitely supported u : E∗→ R,

kuk2 −1,k =

X

n≥1

(n + 1)kkπnuk2−1

and for any local function u,

kuk−1 = kTuk−1,−1.

3.3. The Fourier coefficients of the generator L. We conclude this section deriving explicit expressions for the generator L on the basis {ΨA, A ⊂ Zd}. A

long and elementary computation gives the following dual representation: For every local function u =P A∈Eu(A)ΨA, Lu = X A∈E (Lαu)(A)ΨA, where Lα= Ls+ (1 − 2α)Ld+pχ(α)(L++ L−), (Ldu)(A) = X x∈A,y6∈A

a(y − x){u(Ax,y) − u(A)} ,

(L+u)(A) = 2 X x∈A,y∈A a(y − x) u(A\{y}) , (L−u)(A) = 2 X x6∈A,y6∈A

a(y − x) u(A ∪ {y})

and Lsis defined by (3.9). Furthermore, for any function u : E → R, TLαu= LαTu,

provided

Lα= Ls+ (1 − 2α)Ld+

p

χ(α){L++ L−}

and, for A ∈ E∗, v : E∗→ R a finitely supported function,

(Ldv)(A) =

X

x∈A,y6∈A x,y6=0

a(y − x){v(Ax,y) − v(A)) +

X y6∈A y6=0 a(y){v(SyA) − v(A)} , (L+v)(A) = 2 X x∈A,y∈A a(y − x) v(A\{y}) + 2X x∈A a(x){v(A\{x}) − v(Sx[A\{x}])} , (L−v)(A) = 2 X x6∈A,y6∈A x,y6=0

(10)

4. Approximations in H−1

From this section on, we work in dimension d ≥ 3. All finite constants denoted by C0 are allowed to depend on the transition probability p(·) and only on p(·).

The main goal of this section is to show that finitely supported functions can be approximated in H−1 by finitely supported functions in the image of the operator

Lα. We start proving that all finitely supported functions w : E∗ → R such that

w(φ) = 0 are in H−1.

Several estimates on the operators Ld, L−, L+ are assumed here and proved in

section 7.

Theorem 4.1. A finitely supported function w : E∗ → R belongs to H−1,k(E∗, Ls)

for every k ≥ 1 provided w(φ) = 0.

This result states that any local function w in ⊕n≥2Gn belongs to H−1(C, Ls,

 ·, · ). Notice that we are requiring the degree to be greater or equal to 2. The reason is simple. Any local function f of degree one has H1 norm equal to 0: if

f =P x∈ZdcxΨx,  f, Lf  =  f, Lsf  (4.1) = X x,y∈Zd cxcy X z,w∈Zd s(w) < Ψx+z, Ψy+w− Ψy > = 0 .

There are thus two possibilities. Either P

xcx = 0, in which case f = 0 both in

L2( ·, · ) and in H

1 orPxcx6= 0 in which case f = 0 in H1 but not in L2. In

this later circumstance, f does not belong to H−1. Therefore, a function of degree

one belongs to H−1 only if it vanishes in L2, i.e., only ifPxcx= 0. This explains

the restriction on the degree in the previous theorem.

Theorem 4.1 is proved in the same way as Lemma 2.1 in [13]. One has to show that the Green function associated to the generator Ls restricted to E∗,n, for

some n ≥ 1, is finite. This is done by comparing the Green function with the one associated to independent random walks, which is finite because d ≥ 3.

Recall from Remark 3.1 in the previous section that for a finitely supported function f : E → R, (Tf)(A) = (Tf)(SzA) for all z in A. Thus, for n ≥ 0, denote by

In the closed subspace of L2(E∗,n) of all functions f for which (3.6) holds and let

I = ⊕n≥0In.

We are now in a position to state the main result of this section. We have just seen that all finitely supported functions w : E∗→ R such that w(φ) = 0 belong to

H−1. We now prove that these functions can be approximated in H−1 by finitely

supported functions in the image of the operator Lα.

Theorem 4.2. Fix a finitely supported function w : E∗→ R in I such that w(φ) =

0. For each ε > 0 and k ≥ −1, there exists a finitely supported function f : E∗→ R

such that

kLαf+ wk−1,k ≤ ε .

Moreover, we may take f in I and f(φ) = 0.

The proof of this result requires several estimates on the resolvent equation (4.2). The existence of a solution in L2of the resolvent equation stated in the next lemma requires a proof because Lαis not the generator of a Markov process.

(11)

Lemma 4.3. For each λ > 0, there exists a function uλ in I which solves the

resolvent equation

λuλ− Lαuλ= w . (4.2)

Proof: For n ≥ 1, let Πn be the projection on Snj=1E∗,j: Πn =P1≤j≤nπj and

let Mn = Πn(L++ L−)Πn. We first prove the existence of a solution in L2 of the

truncated resolvent equation

λuλ,n − {Ls+ (1 − 2α)Ld+

p

χ(α)Mn}uλ,n = w . (4.3)

By Lemma 7.1, the operators Ls, Ld and L± are bounded on L2(Snj=1E∗,j) for

each fixed n ≥ 1. There exists, in particular, a solution for λ large enough. Ls is a

symmetric negative semi-definite operator while, by Lemma 7.2, Ld and L++ L−

are anti-symmetric. Since Πn is a projection, for any f in L2(E∗), < f, Mnf>=<

f, Πn(L++ L−)Πnf >= < Πnf, (L+ + L−)Πnf >= 0, so that Mn is also

anti-symmetric. Therefore, taking the scalar product on both sides of the previous identity with respect to uλ,n we obtain by Schwarz inequality that

λkuλ,nk0 ≤ kwk0.

Here k · k0 stand for the L2(E∗) norm. By the proof of Proposition I.2.8 (b) in

[12], there exists a solution of (4.3) for every λ > 0. Moreover, uλ,n belongs to I

because w does.

Up to this point we proved the existence of a solution of the resolvent equation (4.3). The previous estimate shows that the sequence {uλ,n, n ≥ 1} is uniformly

bounded in L2 for each λ > 0. Moreover, by Lemma 7.7 and the proof of Lemma

2.5 in [8] or the proof of Theorem 5.1 in [13], since w is finitely supported, for every k ≥ 1,

λX

j≥1

jkkπjuλ,nk0 ≤ C(k, w)

uniformly over n. Let f be a limit point of the sequence {uλ,n, n ≥ 1}. f inherits the

previous bound and belongs therefore to the domain of the operators Ls, Ld, L±.

Furthermore, taking scalar products with finitely supported functions, it is easy to show that any limit point of this sequence is a solution of the resolvent equation (4.2). Finally, f belongs to I because each function uλ,n belongs and I is closed.

This proves the lemma.

By the proof of Lemma 2.5 in [8] or the proof of Theorem 5.1 in [13] we may deduce from Lemma 7.7 the following bound on the solution uλ of the resolvent

equation (4.2).

Theorem 4.4. Let uλ be the solution of the resolvent equation (4.2). For any

positive integer k, there exists a finite constant Ck depending only on k such that

sup

λ≥0

λkuλk20,k+ kuλk21,k

≤ Ckkwk2−1,k . (4.4)

The following estimate on the asymmetric part of the generator that preserves the degree is needed in the proof of Theorem 4.2.

Lemma 4.5. Let uλbe the solution of the resolvent equation (4.2). For any k ≥ 1,

there exists a finite constant Ck, depending only on k, such that

(12)

Proof: Let w1= w +pχ(α)[L++ L−]uλ so that

λuλ− {Ls+ (1 − 2α)Ld} uλ= w1.

By Lemma 7.7 there exists a finite constant C0 such that

kπnw1k2−1 ≤ 2kπnwk2−1 + C0n n+1 X j=n−1 kπjuλk21 (4.5) for all n ≥ 1.

Notice that the operator Ls+(1−2α)Lddoes not change the degree of a function.

We may therefore examine equation (4.6) on each set E∗,n:

λπnuλ− {Ls+ (1 − 2α)Ld} πnuλ= πnw1. (4.6)

Since n is fixed until estimate (4.10), we omit the operator πnin the next formulas.

Following section 6 of [13], we approximate this operator by a convolution opera-tor that can be analyzed through Fourier transforms. Fix n ≥ 1 and let Xn= (Zd)n.

We consider a set A in E∗,n as an equivalent class on n! sets of distinct points of

Zd∗. A function f : E∗,n→ R can be lifted into a symmetric function Bf on Xn, that

vanishes on Xn\ En,∗: Bf(x1, . . . , xn) = 0 if xi= xj for some i 6= j or if xi = 0 for

some 1 ≤ i ≤ n. The operators Ls, Ld can also be extended in a natural way to Xn.

Denote by {ej, 1 ≤ j ≤ n} the canonical basis of Rn, by 1 the vector P1≤j≤nej

and consider on Xn the operators eLs, eLd defined by

(eLsf)(x) = X 1≤j≤n z∈Zd s(z)f(x + zej) − f(x) + X z∈Zd s(z)f(x − z1) − f(x) , (4.7) (eLdf)(x) = X 1≤j≤n z∈Zd a(z)f(x + zej) − f(x) + X z∈Zd a(z)f(x − z1) − f(x) .

In this formula and below, x = (x1, . . . , xn) is an element of Xn, so that each

xj belongs to Zd and x + zej = (x1, . . . , xj−1, xj + z, xj+1, . . . , xn), x + z1 =

(x1+ z, . . . , xn+ z).

Denote by k · kXn,1the H1norm associated to the generator eLs: for each function

f: Xn → R, kfk2Xn,1 = 1 n! X x∈Xn f(x)(−eLs)f(x)

and denote by k · kXn,−1its dual norm defined by

kfk2 Xn,−1 = 1 n! X x∈Xn f(x)(−eLs)−1f(x) .

Lifting the resolvent equation (4.6) to Xn and adding and subtracting eLsBuλ+

(1 − 2α)eLdBuλ, we obtain that

λBuλ−  e Ls+ (1 − 2α)eLd Buλ= w2, (4.8) where w2= Bw1 + BLs− eLsB uλ + (1 − 2α)BLd− eLdB uλ.

(13)

We claim that w2 has finite H−1(Xn) norm. Indeed, for each n ≥ 1, by (7.5) and

Lemma 7.6 below, there exists a finite constant C0 such that

kπnBw1k2Xn,−1 ≤ kπnw1k 2 −1, kBLsπnuλ− eLsBπnuλk2Xn,−1 ≤ C0n 2 nuλk21, kBLdπnuλ− eLdBπnuλk2Xn,−1 ≤ C0n 2 nuλk21 so that kπnw2k2Xn,−1 ≤ 2kπnw1k 2 −1 + C0n2kπnuλk21 (4.9)

for some finite constant C0.

It remains to examine the resolvent equation (4.8) through Fourier analysis. Let Tn,d= [−π, π]nd and denote bybuλ: (T

d)n→ C the Fourier transform of Bu λ: b uλ(k) = X x∈Xn eix·k(Buλ)(x) . In this formula, x · k =P

1≤j≤nxj· kj. It follows from the resolvent equation (4.8)

thatbuλ is the solution of

λbuλ(k) − n b Ls(k) + (1 − 2α)bLd(k) o buλ(k) = wb2(k) , where bLs, bLd are the functions associated to the operators eLs, eLd:

−bLs(k) = 2 X 1≤j≤n z∈Zd s(z){1 − cos(kj· z)} + 2 X z∈Zd s(z){1 − cos( n X j=1 kj· z)} , −bLd(k) = 2i X 1≤j≤n z∈Zd a(z) sin(kj· z) − 2i X z∈Zd a(z) sin( n X j=1 kj· z) .

The H−1(Xn, bLs) norm of a function v : Xn→ R has a simple and explicit expression

in terms of the Fourier transform: kvk2 Xn,−1 = −1 n!(2π)nd Z Tn,d dk |bv(k)|2 1 b Ls(k) ·

Since Buλ is the solution of the resolvent equation (4.8), for every λ > 0,

keLdBuλk2Xn,−1 = −1 n!(2π)nd Z Tn,d b Ld(k) λ − bLs(k) − (1 − 2α)bLd(k) 2 |wb2(k)| 2 b Ls(k) dk .

It follows from the explicit formulas for the functions bLs, bLdand a Taylor expansion

for |k| small that the previous expression is bounded by −C0 n!(2π)nd Z Tn,d |wb2(k)| 2 b Ls(k) dk = C0kw2k2Xn,−1

for some finite constant C0. We have thus proved that

keLdBuλk2Xn,−1 ≤ C0kw2k

2

(14)

We may now conclude the proof of the Lemma 4.5. Fix n ≥ 1. By (7.5), by the estimate presented just before (4.9) and by the inequality just derived, there exists a finite constant C0, which may change from line to line, such that

kπnLduλk2−1 = kLdπnuλk2−1 ≤ C0nkBLdπnuλk2Xn,−1 ≤ C0n n n2kπnuλk21+ keLdBπnuλk2Xn,−1 o ≤ C0 n n3kπnuλk21+ nkπnw2k2Xn,−1 o . In particular, by (4.9) and (4.5), kπnLduλk2−1 ≤ C0 n n3kπnuλk21+ nkπnw1k2Xn,−1 o (4.11) ≤ C0 n n3 n+1 X j=n−1 kπjuλk21+ nkπnwk2Xn,−1 o .

It remains to recall the definition of the norm k·k−1,kand the statement of Theorem

4.4 to conclude the proof.

Notice that the constant Ck, which appears in the statement of Lemma 4.5,

depends on k only because the one which appears in Theorem 4.4 depends on k. We have now all elements to prove Theorem 4.2.

Proof of Theorem 4.2. The proof is done in two steps. We first use the resolvent equation (4.2) to obtain a sequence {vj; j ≥ 1} of functions in L2(E∗, k) such that

lim

j→∞kLαvj+ wk−1,k = 0 .

The sequence vj is obtained through convex combinations (in λ) of the solutions

uλ of the resolvent equation (4.2). Details can be found at the end of the proof of

Lemma 2.1 in [8] or at the beginning of the proof of Lemma 2.8 in [10].

At this point, it remains to show the existence, for each fixed j and ε > 0, of a finitely supported function f : E∗ → R such that kLαf − Lαvjk−1,k ≤ ε. To prove

the existence of such function f, assume that f vanishes onS

j≥nE∗,j and recall the

decomposition of the operator Lα to deduce that

kLαf− Lαvjk−1,k ≤ kL+(f − Πnvj)k−1,k + kL−(f − Πnvj)k−1,k (4.12)

+ kLd(f − Πnvj)k−1,k + kLs(f − Πnvj)k−1,k + kLα(I − Πn)vj)k−1,k,

where Πn =Pj≤nπj and I is the identity. We estimate each term on the right

hand side separately. By Lemma 7.7, the first term on the right hand side is such that kL+(f − Πnvj)k2−1,k = n X `=0 (` + 1)kkL+π`{f − vj}k2−1 ≤ C0 n+1 X `=1 `k+1kπ`f− π`vjk21 ≤ C0 n+1 X `=1 `k+2kπ`f− π`vjk20.

The second term on the right hand side of (4.12) is estimated in the same way. By Lemma 7.4, the third one is such that

kLd(f − Πnvj)k2−1,k = n X `=1 `kkLdπ`(f − vj)k2−1 ≤ C0 n X `=1 `k+1kπ`(f − vj)k20

(15)

for some finite constant C0. The fourth member on the right hand side of (4.12) is

easily estimated by exactly the same arguments and by using again Lemma 7.4. Finally, since vjis a convex combination of the solutions of the resolvent equation

(4.2), by (4.11), kLα(I − Πn)vj)k2−1,k = X `>n `kkLαπ`vjk2−1 ≤ X `≥n `k+3kπ`vjk21 + X `≥n `k+1kπ`wk2−1.

Now, for ε > 0 fixed, since w is finitely supported, by Theorem 4.4 and Theorem 4.1 there exists n0 > 0 large enough for the last quantity to be bounded by ε.

For this fixed n0, find a finitely supported function f :Sn≤n0E∗,n → R for which

all previous expressions are bounded by ε, which is possible because vj belongs to

L2(E∗, k).

It remains to check that we may take f in I with f(φ) = 0. The first property follows from the fact, easy to verify, that the operators Ls, Ld, L+, L− map the

closed subspace I of L2(E∗) into I. In particular, the solutions of the resolvent

equations, as well as their convex combinations, belong to I so that f can be taken in I.

The second requirement follows from the fact that (L∗f)(φ) = 0 for any finitely

supported function f in I, where L∗ stands for any of the four operators Ls, Ld,

L+, L− and from the fact that (L+f)({x}) = 0 for all x in Zd∗. These two properties

show that we may set the value of f at φ to be 0 without changing Lαf.

The special features of the operators L∗ just used are proved in the beginning of

section 7. This concludes the proof.

5. The fluctuation-dissipation theorem

We consider in this section the general asymmetric simple exclusion in dimension d ≥ 3. Here again, all finite constants C0 which appear in the statement of the

theorems may depend only on the transition probability p(·).

We prove in this section a fluctuation–dissipation theorem, Theorem 5.3 below, also called the Boltzmann-Gibbs principle. It allows the replacement of a local function w in ⊕n≥2Gn by the sum of gradients η(x + ej) − η(x) and local functions

in the range of the generator. This result is the main step in the proof of Gaussian fluctuations of the empirical measure around the hydrodynamic limit (cf. [4]).

Denote by Eνα the expectation on the path space D(R+, X ) corresponding to

the Markov process with generator given by (2.1) and starting from the stationary measure να. The first result states that a local function f in H−1(Ls,  ·, · ) has

a finite space-time variance in the diffusive scaling:

Theorem 5.1. Fix a smooth function G : Rd → R with compact support, a local function f of degree n ≥ 2, T > 0 and a vector v in Rd. There exists a finite constant C0 such that

lim sup →0 E να h sup 0≤t≤T  d/2+1 Z t−2 0 X x∈Zd G((x − rv))f (τxηr) dr 2i ≤ C0T kGk2L2k Tf k 2 −1.

(16)

Proof: Set Gr(x) = G((x − rv)) and recall Lemma 4.3 in [2] to obtain that Eνα h sup 0≤t≤T  d/2+1 Z t−2 0 dr X x∈Zd Gr(x)f (τxηr) dr 2i ≤ 14 Z T 0 dr d< X x∈Zd Gr(x)τxf , (−Ls)−1 X x∈Zd Gr(x)τxf > . (5.1)

Fix a local function g of degree n ≥ 2. Let Lsbe the generator defined in (3.9)

and g be the Fourier coefficient of g. Then,

< g, (−Ls)−1g > = < g, (−Ls)−1g> ,

where the last scalar product is on En. Recall from section 4 that Xn= (Zd)n. The

proof of Lemma 7.5 shows that the previous expression is bounded by C0 (n − 1)! < Bg, (− eL s 0) −1Bg> Xn

for some finite constant C0, where eLs0 stands for the generator of n independent

random walks on Zd with transition probability s(·) and Bg for the extension of g to Xn. Denote by Gn(·, ·) the Green function associated to the generator eLs0

restricted to Xn. The previous expression can be written as

C0 (n − 1)! X x,y∈Xn (Bg)(x)Gn(x, y)(Bg)(y) . In particular, for g =P

x∈ZdGr(x)τxf , the right hand side of (5.1) is bounded by

C0 (n − 1)! Z T 0 dr d X z,w∈Zd Gr(z)Gr(w) X x,y∈Xn Bf(x + z1)Gn(x, y)Bf(y + w1) .

A change of variables permits to rewrite the expression inside the integral as d X

z,w∈Zd

Gr(z)Gr(z − w) X

x,y∈Xn

Bf(x)Gn(x, y + w1)Bf(y) .

Since f is a local function, f has finite support. In particular, the sums over x, y are carried over finite sets. On the other hand, replacing Gr(z − w) by Gr(z), we

write the previous expression as d X z∈Zd Gr(z)2 X x,y∈Xn Bf(x) X w∈Zd Gn(x, y + w1)Bf(y) + d X z,w∈Zd Gr(z)nGr(z − w) − Gr(z)o X x,y∈Xn Bf(x)Gn(x, y + w1)Bf(y) .

We claim that the second term is bounded above by C(f, G)1/2for some finite

constant depending on f and G. Indeed, since G has a bounded derivative and Bf a finite support, the second line is less than or equal to

C(k∇Gk∞, f ) 1/2 sup x∈Xf− X w∈Zd |w|1/2G n(0, x + w1) . (5.2)

In this formula, Xf− is the set of all sites which can be written as difference of two points in the support of Bf: Xf− = {y − x, Bf(x)Bf(y) 6= 0}. Since Gn(0, z) is the

(17)

sum over w is thus finite, uniformly over x, as soon as (n − 1)d > 5/2, inequality which is fulfilled because we are assuming d ≥ 3 and n ≥ 2. This proves (5.2).

Collecting all previous estimates we obtain that the expectation which appears in the statement of the lemma is bounded above by

C0 (n − 1)! Z T 0 dr d X z∈Zd Gr(z)2 X x,y∈Xn Bf(x) X w∈Zd Gn(x, y + w1)Bf(y)

plus a remainder which converges to 0 as  ↓ 0. Denote by eGn(·, ·) the Green

function associated to the generator eLs defined in (4.7) and restricted to Xn. An

elementary computation shows that X x,y∈Xn Bf(x) X w∈Zd Gn(x, y + w1)Bf(y) = X x,y∈Xn−1 B¯f(x) eGn−1(x, y)B¯f(y) = (n − 1)! kπn−1B¯fk2−1,Xn−1.

On the right hand side, ¯f stands for Tf. Notice that n is replaced by n − 1 in the right hand side because we are fixing a particle at the origin. The last time integral is thus equal to C0 Z T 0 dr d X z∈Zd Gr(z)2kB¯fk2 Xn,−1 ≤ C0 Z T 0 dr d X z∈Zd Gr(z)2k¯fk2 −1,

where in the last step we used Lemma 7.5. As  ↓ 0 this integral converges to C0T kGk2L2k¯fk2−1 ,

which concludes the proof of the lemma.

Corollary 5.2. Fix a smooth function G : Rd → R with compact support, a local function f in ⊕n≥2Gn, T > 0 and a vector v in Rd. There exists a finite constant

C0 such that lim sup →0 E να h sup 0≤t≤T  d/2+1Z t−2 0 X x∈Zd G((x − rv))f (τxηr) dr 2i ≤ C0T kGk2L2k Tf k2−1,0

Proof: Since the symmetric part of the generator does not change the degree of local functions, for a local function g in ⊕n≥2Gn,

< g, (−Ls)−1g > = X n≥2 < πng, (−Ls)−1πng > = X n≥2 < πng, (−Ls)−1πng>

and we may proceed as in the proof of Theorem 5.1.

Fix a local function w in ⊕n≥2Gn and denote by w its Fourier coefficients and

by w the finitely supported function Tw : E∗ → R. For each λ > 0, let uλ be the

solution of the resolvent equation

λuλ− Lαuλ= w .

We proved in section 4 the existence of the solution uλ of the resolvent equation

(18)

show the existence of a subsequence λk for which the sequence uλk(α, z) converges,

as k ↑ ∞, uniformly in α in [0, 1], to some limit, denoted by Dz(α):

Dz(α) = lim

k→∞uλk(α, {z}) (5.3)

for each z in Zd ∗.

Theorem 5.3. Fix a local function w in ⊕n≥2Gn and a smooth function G : Rd→

R with compact support. There exists a sequence of local functions umsuch that

lim sup m→∞ lim sup →0 E να h sup 0≤t≤T  d/2+1 Z t−2 0 X x∈Zd G(x)τxWm(ηs) ds 2i = 0 , where Wm(η) = Wα,m(η) = w − Lum + p χ(α)X z∈Zd a(z)Dz(α){Ψz− Ψ0} .

We will refer to Theorem 5.3 as the fluctuation-dissipation theorem. It will be the basic ingredient to study the equilibrium fluctuations for the density field. Proof: Recall that ¯w = Tw. Since w belongs to ⊕n≥2Gn, by Theorem 4.2 with

k = 0, there exists a sequence of finitely supported functions vm : E∗ → R such

that vm(φ) = 0, vmbelongs to I and

lim

m→∞k ¯w− Lαvmk−1 = 0 .

Since vm satisfies (3.6), in view of (3.7), there exists a finitely supported function

um : E → R such that Tum = vm. Moreover, um(A) 6= 0 only if A contains the

origin and um({0}) = 0 because vm(φ) = 0. Let um be the local function defined

by um=PA∈Eum(A)ΨAand let ˆWmbe the local function in ⊕n≥2Gn defined by

ˆ Wm = w − {Lum− π1Lum} . By Corollary 5.2, lim sup →0 E να h sup 0≤t≤T  d/2+1 Z t−2 0 X x∈Zd G(x)τxWˆm(ηs) ds 2i ≤ C0T kGk2L2kT ˆWmk2−1 (5.4)

for some finite constant C0. An elementary computation gives that

π1(Lu) = −

X

x,y∈Zd

s(y − x){u(y) − u(x)}{Ψy− Ψx}

− 1 − 2α 2

X

x,y∈Zd

a(y − x){u(y) + u(x)}{Ψy− Ψx}

+ pχ(α) X

x,y∈Zd

a(y − x)u(x, y){Ψy− Ψx} .

Since um({x}) = 0 for every x in Zd, um(A) 6= 0 only if A contains the origin,

um(A) = |A|−1vm(A \ {0}), we have that

π1(Lu) =

p

χ(α) X

x∈Zd

(19)

which vanishes in L2(χ,  ·, · ) so that Tπ1(Lu) = 0. In particular, T ˆWm =

¯

w− Lαvmand the right hand side of (5.4) is bounded above by

C0T kGk2L2k ¯w− Lαvmk2−1.

Since this expression vanishes as m ↑ ∞, all we need to prove is that lim sup m→∞ lim sup →0 E να h sup 0≤t≤T  d/2+1 Z t−2 0 X x∈Zd G(x)τxW˜m(ηs) ds 2i = 0 , where ˜ Wm = π1Lum − p χ(α)X z∈Zd a(z)Dz(α){Ψz− Ψ0} .

By the explicit formula for π1Lumobtained above,

˜ Wm = p χ(α) X x∈Zd a(x){vm(x) − Dx(α)}{Ψx− Ψ0} .

Since ναis a stationary state, by Schwarz inequality, the previous expectation is

bounded above by d−2T2Eνα h X x∈Zd G(x)τxW˜m(η) 2i . A change of variables gives that this expression is equal to

d−2χ(α)T2 X

x∈Zd

 X

y∈Zd

a(y){G([x − y]) − G(x)}{vm(y) − Dy(α)}

2 .

A Taylor expansion shows that this expression converges, as  ↓ 0, to χ(α)T2 Z Rd da {(∇G)(a) · Rm}2, where Rm= (R1m, . . . , Rmd), Rjm= P

y∈Zda(y)yj{vm(y)−Dy(α)}. By the definition

of Dy(α), Rmj → 0 as m ↑ ∞ because vm is constructed as convex combinations

of uλk, the solution of the resolvent equation (4.2). The integral therefore tends to

0.

6. Regularity of Viscosity coefficients

Fix a local function w in ⊕n≥2Gn and denote by w its Fourier coefficients and

by w the finitely supported function Tw : E∗ → R. For each λ > 0, let uλ be the

solution of the resolvent equation

λuλ− Lαuλ= w .

We proved in section 4 the existence of the solution uλof the resolvent equation. We

prove in this section the existence of a subsequence λk for which uλk(·, z) converges

uniformly in [0, 1], as well as all its derivatives, to a smooth function Dz(·).

The proof is very close to the one presented in ([11]) for the self diffusion in the symmetric case, so we only show here the main points.

We want to show that there exists a subsequence λk ↓ 0, such that, for each z

(20)

To prove the existence of such subsequence it is enough to show that uλ(α, {z}) are

smooth function of α for each λ > 0 and each z, and sup

λ>0

sup

0≤α≤1

|u(j)λ (α, {z})| < ∞ where u(j)λ (α, {z}) stands for the jth derivative of u

λ.

By Lemma 3.1 of [13], we have the bound |u(j)λ (α, {z})| ≤ C0ku

(j) λ (α, ·)k1

for some finite constant C0 depending only on p(·). With the iterated use of

Theo-rem 4.4, we will prove that, for any k and j, sup

λ>0

sup

0≤α≤1

ku(j)λ (α, ·)k1,k< ∞ .

Since the coefficients of Lαare not smooth at the boundary of [0, 1], we

reparame-terize by α = sin2t, t ∈ [0, 2π], as done in [11]. We obtain

L(t) = Ls+ (cos2t − sin2t)Ld+ (sin t cos t) {L++ L−}

and then we consider the equation

λvλ(t) − L(t)vλ(t) = w

Since w does not depends on α, we have uλ(α(t)) = vλ(t). So if we prove that

vλ(j)(t) are uniformly bounded in the k · k1,k norm, we obtain the boundedness in

the same norm for uλ(α) for α in the interior of [0, 1]. The extra argument to

extend this smoothness up to the boundary is identical to the one used in [11] (see the end of the proof of Theorem 5.1 in there).

Differentiating formally L(t) in t, we obtain

L0(t) = −4(sin t cos t)Ld+ (cos2t − sin2t) {L++ L−} .

By Lemma 5.2 of [11], vλ(t) is differentiable in t, and its derivative v0λ(t) satisfies

λv0λ(t) − L(t)v0λ(t) = L0(t)vλ(t) . (6.1)

As a consequence of Lemma 7.7, Lemma 4.5 and the explicit form of L(t), there exists a constant Ck depending only on k

kL0(t)vλ(t)k−1,k ≤ Ckkwk−1,k+3.

Then if we apply Theorem 4.4 to equation (6.1), we obtain the bound kv0λ(t)k−1,k

uniform in t and λ. The argument can be iterated exactly as done in [11], obtaining similar bounds for all the derivatives vλ(j)(t).

7. Estimates on the operators Ld, L+, L−

We prove in this section some elementary identities or estimates involving the operators Ld, L+, L−. Recall that all functions f : E∗→ R which come from a local

function f through the transformation T of the Fourier coefficients of f are such that f(SzA) = f(A) for all z in A. Recall also the definition of the spaces In given

just before the statement of Theorem 4.2.

A simple computation shows that the space I is left invariant by the operators Ls, Ld, L+, L−: For every n ≥ 1 and every f : In→ R,

(21)

This claim can be proved in two different ways. Either by a direct computation or by reconstructing a local function f from f. More precisely, to prove that Lsf∗

belongs to Inif f∗belongs to In, let f be given by (3.7) so that Tf = f∗. By (3.10),

TLsf= Lsf∗, which proves that Lsf∗ belongs to I.

We turn now to an elementary identity to illustrate the fact that the space I enjoys some special properties. For every f : E∗,1→ R,

(L−f)(φ) = −2

X

x6=0

a(x)f({x}) .

In particular, (L−f)(φ) = 0 for all f in I1because in this space f({x}) = f({−x}) and

a(·) is anti-symmetric. In contrast, (L+g)({x}) = 0 for all functions g : E∗,0→ R

so that, for all f in I1 and all g : E∗,0→ R,

L−f = 0 , L+g = 0 . (7.2)

We turn now to the proof of some estimates involving the operators Ls, Ld, L−

and L+. The first lemma states that the operators Ls, Ld, L−and L+ are bounded

in L2(E

∗,n) for each fixed n ≥ 1.

Lemma 7.1. There exists a finite constant C0 such that

kAfk2

0 ≤ C0n2kfk20

for each f in L2(E

∗,n), were the operator A stands for Ls, Ld, L+ or L−.

Proof: We only prove the estimate concerning L−, the other ones being elementary.

Fix a function f : E∗,n→ R in L2and keep in mind that L−fmaps E∗,n−1 in R. By

the explicit expression for L− and a change of variables,

kL−fk20 ≤ 4 X A∈E∗,n−1 n X x,y6∈A x,y6=0

a(y − x)f(A ∪ {y})o

2

.

SinceP

x∈Zda(x) = 0, the previous expression is less than or equal to

8 X A∈E∗,n−1 n X y6∈A y6=0 a(y)f(A ∪ {y})o 2 + 8 X A∈E∗,n−1 n X y6∈A, y6=0 x∈A

a(y − x)f(A ∪ {y})o

2

.

By Schwarz inequality, since a(·) is absolutely bounded andP

x∈Zd|a(x)| is finite,

this expression is less than or equal to C0n X A∈E∗,n−1 X y6∈A y6=0 f(A ∪ {y})2

for some finite constant C0. To sum over A in E∗,n−1 and over y 6= 0, y 6∈ A is

the same as to sum over B in E∗,n with a multiplicity factor n because all sets are

counted n times. The previous expression is thus equal to C0n2

X

A∈E∗,n

f(A)2, which concludes the proof of the lemma.

It follows from the next statement that the operators Ld and L++ L− are

anti-symmetric on I ∩ L2 0,−1(E∗).

(22)

Lemma 7.2. For every n ≥ 1 and every finitely supported functions u, v : E∗,n→ R

< Ldu, v > = − < u, Ldv> .

For every finitely supported functions f, g in In−1, In respectively,

1

n + 1< L+f, g > = − 1

n < f, L−g> . Proof: The first identity is elementary and relies on the fact thatP

x,y∈Aa(y −x) =

0. Note, however, that both pieces of the operator are needed.

The proof of the second statement is more demanding. Fix finitely supported functions f, g in In−1, In, respectively. By the explicit form of L+,

< g, L+f> = 2

X

A∈E∗,n

X

x,y∈A

a(y − x) g(A)f(A \ {y})

+ 2 X

A∈E∗,n

X

x∈A

a(x) g(A)nf(A \ {x}) − f(Sx[A \ {x}])

o .

Since Sx[A \ {x}] = SxA \ {−x} and since g(SxA) = g(A) for x in A because g

belongs to In, a change of variables B = SxA, x0 = −x in the second piece of the

second term permits to rewrite the second term on the right hand side as 4X x6=0 a(x) X A∈E∗,n A3x g(A)f(A \ {x})

because a(−x) = −a(x). We claim that X x6=0 a(x) X A∈E∗,n A3x g(A)f(A \ {x}) = 1 n − 1 X x,y6=0 a(y − x) X A∈E∗,n A3x,y g(A)f(A \ {y}) . (7.3)

We conclude the proof of the lemma assuming (7.3), whose proof is presented at the end. It follows from identity (7.3) and the previous expression for < g, L+f>

that < g, L+f> = 2  1 + 1 n  X A∈E∗,n X x,y∈A

a(y − x) g(A)f(A \ {y})

+ 21 + 1 n  X A∈E∗,n X x∈A a(x) g(A)f(A \ {x}) . The first term of the right hand side, which can be written as

21 + 1 n  X y6=0 X A∈E∗,n A3y g(A)f(A \ {y})X x∈A a(y − x) , is equal to − 21 + 1 n  X y6=0 a(y) X A∈E∗,n A3y g(A)f(A \ {y}) − 21 + 1 n  X x,y6=0 a(y − x) X A∈E∗,n A3y,A6∈x g(A)f(A \ {y})

(23)

becauseP

x∈Aa(y − x) = −a(y) −

P

x6=0,x6∈Aa(y − x). The first term of this formula

cancels with the second one in the last expression for < g, L+f>. Therefore,

< g, L+f> = − 2  1 + 1 n  X x,y6=0 a(y − x) X A∈E∗,n A3y,A6∈x g(A)f(A \ {y}) .

To conclude the proof of the lemma, it remains to change variables B = A \ {y} and to recall the definition of the operator L−.

We turn now to the proof of Claim (7.3). Since for y in A, g(A) = g(SyA) and

since |A| = n, the left hand side of (7.3) is equal to 1 n X x6=0 a(x) X A∈E∗,n A3x X y∈A∪{0} y6=x g(SyA)f(A \ {x}) = 1 n X x,y6=0 y6=x a(x) X A∈E∗,n A3x,y g(SyA)f(A \ {x}) + 1 n X x6=0 a(x) X A∈E∗,n A3x g(A)f(A \ {x}) .

Notice that the second term on the right hand side is precisely the original one. Consider the first term. Perform a change of variables B = SyA, rewrite (S−yA) \

{x} as S−y(A \ {x − y}) and recall that f(S−y(A \ {x − y}) = f(A \ {x − y}) if −y

belongs to A because f is in In−1, to rewrite this expression as

1 n X x,y6=0 y6=x a(x) X A∈E∗,n A3x−y,−y g(A)f(A \ {x − y}) .

A change of variables x0= x − y, y0= −y, shows that this expression is equal to 1 n X x,y6=0 a(x − y) X A∈E∗,n A3x,y g(A)f(A \ {x}) .

To prove (7.3), it remains to recollect all previous identities.

It follows from this result that the operator L+ + L− is anti-symmetric in

L20,−1(E∗) ∩ I:

Corollary 7.3. The operator L++ L− is anti-symmetric in L20,−1(E∗) ∩ I:

 f, (L++ L−)g 0,−1= −  (L++ L−)f, g 0,−1

for all finitely supported functions f, g in I. The same statement remains in force if L++ L−is replaced by Πn(L++ L−)Πn for every n ≥ 1, where Πn=P1≤j≤nπj.

The proof of Corollary 7.3 is elementary and left to the reader. One needs only to recall identities (7.2). The next result states that Ldis a bounded operator from

L2(E

∗,n) to H−1(E∗,n).

Lemma 7.4. There exists a finite constant C0, independent of n, such that for

each n ≥ 1 and for any finitely supported functions f, g : E∗,n→ R,

< Ldf, g > ≤ C0 √ n kgk1kfk0. In particular, kLdfk−1 ≤ C0 √ n kfk0.

(24)

The very same estimates remain in force if Ld is replaced by Ls.

The proof of Lemma 7.4 is elementary and left to the reader. Next estimate is Lemma 2.3 of [8]. Recall from section 4 the definition of the operator B.

Lemma 7.5. There exists a finite constant C0such that for any function f : En,∗→

R in H1,

kfk2

1≤ kBfk2Xn,1≤ C0nkfk

2 1.

Proof: The first inequality is elementary and follows from the explicit formulas for the respective H1 norms. To prove the second inequality, for x = (x1, . . . , xn) ∈

En,∗, let W1(x) = n X j=1 s(xj) , W2(x) = n X i,j=1 s(xi− xj) .

We also denote these quantities by W1(A) and W2(A), for A = {x1, . . . , xn}. A

simple computation shows that there exists a finite constant C0 such that

BLsf(x) − eLsBf(x) ≤ C0{W1(x) + W2(x)} |Bf(x)| (7.4) for every x in En,∗ and f : En,∗→ R.

We are now in a position to prove the second bound. By definition, kBfk2 Xn,1 = − 1 n! X x∈Xn (Bf)(x)(eLsBf)(x) .

Since Bf vanishes outside En,∗, we may restrict the sum to En,∗. Now, adding and

subtracting (BLsf)(x) in this expression and recalling (7.4), we obtain that

kBfk2 Xn,1 ≤ kfk 2 1 + C0 n! X x∈En,∗ {W1(x) + W2(x)}{Bf(x)}2 = kfk2 1 + C0 X A∈En,∗ {W1(A) + W2(A)}f(A)2.

By Theorem 4.7 in [8] or Lemma 3.7 in [13] the second term of the previous formula is bounded by C0nkfk21, which concludes the proof of the lemma.

It follows from this result that 1 C0n kfk2 −1 ≤ kBfk2Xn,−1 ≤ kfk 2 −1. (7.5)

Lemma 7.6. There exists a finite constant C0 such that

kBLsf− eLsBfk2Xn,−1 ≤ C0n

2kfk2

1, kBLdf− eLdBfk2Xn,−1≤ C0n

2kfk2 1

for all n ≥ 1 and all functions f : En,∗→ R.

Proof: We prove the first estimate and leave to the reader the details of the second. Fix n ≥ 1 and a function h : Xn → R. We need to estimate the scalar product

1 n! X x∈Xn h(x)nBLsf(x) − eLsBf(x) o (7.6) in terms of the H1(Xn) norm of h and the H1(En,∗) norm of f. There are two

(25)

In the first case, by (7.4), the expression inside braces in previous formula is absolutely bounded by C0W∗(x)|Bf(x)| for some finite constant C0, where W∗(x) =

W1(x) + W2(x). Therefore, the corresponding piece in the previous formula is

bounded above by 1 n! X x∈En,∗ W∗(x)|h(x)||Bf(x)| ≤ 1 2` 1 n! X x∈En,∗ W∗(x)h(x)2 + ` 2 X A∈En,∗ W∗(A)f(A)2 for every ` > 0.

If x does not belong to En,∗, the corresponding piece of the scalar product writes

−2 n! X x6∈Xn z∈Zd 1≤j≤n s(z)h(x)Bf(x + zej) − 2 n! X x6∈Xn z∈Zd s(z)h(x)Bf(x + z1)

because in this case BLsf(x) = 0. We estimate the first term and claim that the

second can be handled in the same way. Since Bf vanishes outside En,∗, it is implicit

in the previous formula that the first sum is restricted to all x such that x + zej

belongs to En,∗. Since x + zej ∈ En,∗ and x 6∈ En,∗, either xj = 0 or xj = xk for

some k. In particular, since 2ab ≤ `a2+ `−1b2 for every ` > 0, a change of variables gives that the first term of the previous formula is bounded above by

1 n!` X x∈Xn h(x)2Wf∗(x) + ` n! X x∈En,∗ Bf(x)2W∗(x) , where fW∗(x) =P1≤j≤n1{xj = 0} + P

j6=k1{xj= xk}. We may of course replace

the sum over Xn by a sum over En,∗ in the second term, loosing the factor n!.

Adding together all previous estimates, we obtain that the scalar product (7.6) is bounded above by C0 n!` X x∈Xn h(x)2{fW∗(x) + W∗(x)} + C0` X A∈En,∗ f(A)2W∗(A) .

The same proof of Theorem 4.7 in [8] or Lemma 3.7 in [13] shows that in dimension d ≥ 3, the first term is bounded above by C1nA−1khk2Xn,1for some finite constant

C1, while the second, by the quoted results, is less than or equal to C1nAkfk21. To

conclude the proof of the lemma, it remains to minimize over A and to recall the variational formula for the H−1 norm.

We conclude this section with a central estimate involving the operators L+, L−.

This result is Theorem 4.4 in [8]. The reader can find in Lemma 4.1 of [13] a clearer proof.

Lemma 7.7. There exists a finite constant C0 depending only on the transition

probability p such that

n X A∈E∗,n+1 (L+f)(A)g(A) o2 ≤ C0n kfk21kgk21, n X A∈E∗,n f(A)(L−g)(A) o2 ≤ C0n kfk21kgk 2 1

(26)

for all n ≥ 1 and all finite supported functions f : E∗,n → R, g : E∗,n+1 → R. In

particular,

kL+fk2−1 ≤ C0n kfk21, kL−gk2−1 ≤ C0n kgk21.

References

[1] Bernardin C.: Regularity of the diffusion coefficient for lattice gas reversible under Bernoulli measures. Stochastic Process. Appl. 101 , no. 1, 43–68, (2002).

[2] C. C. Chang, C. Landim, S. Olla: Equilibrium fluctuations of asymmetric exclusion processes in dimension d ≥ 3. Probability Theory and Related Field, 119, 381-409, (2001).

[3] Esposito, R., Marra, R., Yau, H.T.: Diffusive limit of asymmetric simple exclusion. Rev. Math. Phys. 6, 1233-1267 (1994).

[4] Kipnis, C., Landim, C.: Scaling Limit of Interacting Particle Systems, Springer Verlag, Berlin (1999).

[5] Kipnis C., Varadhan S. R. S.; Central limit theorem for additive functionals of reversible Markov processes and applications to simple exclusion. Commun. Math. Phys. 106 1–19 (1986).

[6] C. Landim, S. Olla, H. T. Yau; First order correction for the hydrodynamic limit of asym-metric simple exclusion processes in dimension d ≥ 3. Communications on Pure and Applied Mathematics 50, 149–203, (1997).

[7] C. Landim, S. Olla, H. T. Yau; Some properties of the diffusion coefficient for asymmetric simple exclusion processes. The Annals of Probability 24, 1779–1807, (1996).

[8] Landim C., Yau H. T.; Fluctuation–dissipation equation of asymmetric simple exclusion pro-cesses, Probab. Th. Rel. Fields 108, 321–356, (1997)

[9] Landim C., Olla S., Varadhan S. R. S.; Finite-dimensional approximation of the self-diffusion coefficient for the exclusion process. Ann.Prob. 30, No. 2, 1-26 (2002).

[10] Landim C., Olla S., Varadhan S. R. S.; Asymptotic behavior of a tagged particle in simple exclusion processes. Bol. Soc. Bras. Mat. 31, 241–275, (2000).

[11] Landim C., Olla S., Varadhan S. R. S.; Symmetric simple exclusion process: regularity of the self-diffusion coefficient. Commun. Math. Phys. 224, 302-321 (2001).

[12] Liggett T. M.; Interacting Particle Systems, Springer Verlag, New York, 1985.

[13] Sethuraman S., Varadhan S. R. S., Yau H. T.; Diffusive limit of a tagged particle in asym-metric exclusion process, Comm.Pure Appl.Math. 53, 972-1006, (2000).

[14] Spohn H.: Large Scale Dynamics of Interacting Particles, Springer 1991.

[15] Varadhan S.R.S.; Non-linear diffusion limit for a system with nearest neighbor interactions II. In Asymptotic Problems in Probability Theory: Stochastic Models and diffusion on Fractals, Elworthy K.D. and Ikeda N. editors. Pitman Research Notes in Mathematics vol. 283, 75–128, John Wiley & Sons, New York (1994).

[16] Varadhan S.R.S.; Self diffusion of a tagged particle in equilibrium for asymmetric mean zero random walks with simple exclusion, Ann. Inst. H. Poincar´e (Probabilit´es), 31, 273–285 (1995). IMPA, Estrada Dona Castorina 110, CEP 22460 Rio de Janeiro, Brasil and CNRS UMR 6085, Universit´e de Rouen, 76128 Mont Saint Aignan, France.

e-mail: landim@impa.br

Ceremade, UMR CNRS 7534, Universit´e de Paris IX - Dauphine, Place du Mar´echal De Lattre De Tassigny 75775 Paris Cedex 16 - France.

e-mail: Stefano.Olla@ceremade.dauphine.fr

Courant Institute of Mathematical Sciences, New York University, 251 Mercer street, New York, NY, 10012, U.S.A.

Références

Documents relatifs

L’action de cet extrait est plus rapide que l’extrait aqueux de la petite centaurée qui a mis plus 2 heures pour réduire les glycémies des rats hyperglycémiques. - Dans une

We show that the fluctuations of density are Gaussian and that, as for the symmetric case, the direct calculation of the two point function by fluc- tuating hydrodynamics agrees

Stationary nonequilibrium states, hydrodynamic fluctuations, bound- ary driven systems, exclusion process.. This research has been partially supported by the agreement France-Br´

The theorem is subsequently used to identify the statistical properties of the climatic variables themselves, and thus to characterize climatic variability from

The constant tensor E ij is the average strain applied to the unit cell and v is a periodic fluctuation, meaning that it takes the same value at homologous points ( x − , x + ) of

Thereby, a consistent picture of the out-of-equilibrium state of colloidal gels appears, with no violation of the FDT for a fast-relaxing observable, like velocity, and an

By using the ACFD formula we gain a deeper understanding of the approximations made in the interatomic pairwise approaches, providing a powerful formalism for further development

Moving to examples of specific Markov processes, in Proposition 5.1 we use the simple sufficient condition of Proposition 2.9 for the existence of a response function, to show