• Aucun résultat trouvé

Fluctuations à l'équilibre d'un modèle stochastique non gradient qui conserve l'énergie

N/A
N/A
Protected

Academic year: 2021

Partager "Fluctuations à l'équilibre d'un modèle stochastique non gradient qui conserve l'énergie"

Copied!
71
0
0

Texte intégral

(1)

0.1 Resum´e . . . i

0.2 Introduction . . . iv

1 Equilibrium Fluctuations 1 1.1 Notations and Results . . . 1

1.2 Convergence of the finite-dimensional distributions . . . 6

1.3 Boltzmann-Gibbs Principle . . . 12

1.4 Central Limit Theorem Variances and Diffusion Coefficient . . . 17

1.5 Tightness . . . 36

2 Technical Results 45 2.1 Some Geometrical Considerations . . . 45

2.2 Spectral Gap . . . 47

2.3 The Space Hy . . . 50

(2)
(3)

0.1

Resum´

e

Dans de r´ecents travaux, un mod`ele microscopique pour la conduction de la chaleur dans les solides a ´et´e examin´e (cf [BO], [B],[FNO]). Dans ce mod`ele, les atomes les plus proches voisins int´eragissent comme oscillateurs coupl´es, forc´es par un bruit additif qui ´echange de l’´energie cin´etique entre les plus proches voisins.

Plus pr´ecis´ement, dans le cas de conditions p´eriodiques au bord, les atomes sont marqu´es par x ∈ TN = {0, · · · , N − 1}. L’espace des configurations est d´efinie par

ΩN = (R × R)TN et, pour un ´el´ement typique (p

x, rx)x∈TN ∈ Ω

N, r

x repr´esente la

distance entre les particules x et x + 1, et px la vitesse de la particule x. Le g´en´erateur

formel du syst`eme est d´efini par LN = AN + SN, o`u

AN = X x∈TN {(px+1− px)∂rx+ (rx− rx−1)∂px} , (0.1.1) et SN = 1 2 X x∈TN Xx,x+1[Xx,x+1] , (0.1.2)

avec Xx,y = py∂px− px∂py. Ici AN est l’op´erateur de Liouville d’une chaˆıne d’oscillateurs

harmoniques et SN est l’op´erateur du bruit.

Notons par {(p(t), r(t)), t ≥ 0} le processus de Markov g´en´er´e par N2L

N (le facteur

N2 correspondant `a une acc´el´eration du temps). Notons C(R

+, ΩN) l’espace des

tra-jectoires continues sur l’espace des configurations. Pour un temps T > 0 fix´e et pour une mesure donn´ee µN sur ΩN, la mesure de probabilit´e dans C([0, T ], ΩN), induite par le processus de Markov et l’´etat initial µN, sera not´ee PµN. Comme d’habitude, EµN

d´esigne l’esp´erance par rapport `a PµN.

Dans ce travail, nous nous concentrons sur l’op´erateur de bruit SN qui n’agit que sur

les vitesses, donc nous restreignons l’espace des configurations `a RTN. L’´energie totale

de la configuration (px)x∈TN est d´efinie par

E = 1 2

X

x∈TN

p2x . (0.1.3)

Il est facile de v´erifier que SN(E ) = 0, i.e l’´energie totale est constante en temps.

La dynamique induite par SN se r´ev`ele ˆetre un syst`eme gradient, ce qui signifie que le

courant microscopique instantan´e de l’´energie entre x et x + 1 peut ˆetre exprim´e comme le gradient d’une fonction locale, `a savoir, Wx,x+1 = 12(px+1− px).

(4)

0.1. RESUM ´E ii

Bien que cela simplifie consid´erablement les d´emonstrations, ¸ca n’est pas une condi-tion naturelle. L’´etude de syst`emes non gradient est de grand int´erˆet, car d’un point de vue physique, ils fournissent des mod`eles plus r´ealistes, et du point de vue math´ematique, les syst`emes gradient forment un ensemble de petite dimension (en un certain sens) de l’espace des mod`eles r´eticuleres stochastiques r´eversibles avec lois de conservation locales (voir [W] et r´ef´erences cit´ees).

Jusqu’`a maintenant, la m´ethode la plus g´en´eral pour traiter avec les syst`emes non gradient a ´et´e d´evelopp´ee par S.R.S Varadhan (cf [V]), o`u grosso modo, l’id´ee est de remplacer le courant par un gradient augment´e d’un terme de fluctuation.

Par analogie avec [V], nous introduisons un coefficient dans le g´en´erateur (0.1.2) afin de briser sa structure de gradient,

LN = 1 2 X x∈TN Xx,x+1[a(px, px+1)Xx,x+1] , (0.1.4)

o`u a(x, y) est une fonction diff´erentiable satisfaisant 0 < c ≤ a(x, y) ≤ C < ∞, et avec d´eriv´ees partielles continues et born´ees.

Le comportement collectif du syst`eme peut ˆetre d´ecrit grˆace `a des mesures em-piriques. La mesure empirique de l’´energie associ´ee au processus est d´efinie par

πtN(ω, du) = 1 N X x∈TN p2x(t)δx N(du) , (0.1.5)

o`u δz repr´esente la mesure de Dirac concentr´ee en z.

On d´esigne par YN

t le champ de fluctuations de la mesure empirique de l’´energie, i.e

une forme lin´eaire qui agit sur les fonctions lisses H : T → R de la fa¸con suivante YtN(H) = √1

N X

x∈TN

H(x/N ){p2x(t) − y2} .

Le principal r´esultat de ce travail ´etablit la convergence en loi de YN

t (H) lorsque N tend

vers l’infini.

Afin de ´etudier les fluctuations `a l’´equilibre de syst`emes de particules, Brox et Rost [BR] ont introduit le principe de Boltzmann-Gibbs, et ont prouv´e leur validit´e pour des processus attractives de zero-range. Chang et Yau [CY] ont propos´e une m´ethode alter-native pour prouver le principe de Boltzmann-Gibbs dans le cas des syst`emes gradient, qui `a ´et´e ´etendue par Lu [Lu] et Sellami [Se] pour des syst`emes non gradient.

(5)

En adaptant la m´ethode introduite dans [V], nous identifions le terme de diffusion (Section 1.4), ce qui nous permet de d´eduire le principe de Boltzmann-Gibbs (Sec-tion 1.3). Ceci est le point essentiel pour montrer que les lois finidimensionnelles de YN

t (H) convergent vers les lois finidimensionnelles d’un processus g´en´eralis´e

d’Ornstein-Uhlenbeck (Section 1.2). De plus, en utilisant `a nouveau le principe de Boltzmann-Gibbs nous obtenons aussi la tension du champ de fluctuation de l’´energie dans un certain es-pace de Sobolev (Section 1.5), ce qui avec la convergence des lois finidimensionnelles im-plique la convergence en loi de YN

t (H) vers le processus g´en´eralis´e d’Ornstein-Uhlenbeck

(mentionn´e ci-dessus). L’id´ee d’utiliser le principe de Boltzmann-Gibbs pour d´emontrer la tension, vient de [CLO].

Le deuxi`eme chapitre est consacr´e `a des d´etails plus techniques. Dans la Section 2.3, nous d´efinisons l’espace Hy et en utilisons une caract´erisation pour prouver le th´eor`eme

5. Cette caract´erisation repose essentiellement sur certaines conditions d’int´egrabilit´e d’un syst`eme de Poisson ainsi que sur une estimation optimale du trou spectral pour le g´en´erateur du processus. Ces deux sujets sont respectivement trait´es dans les Section 2.1 et 2.2. Enfin, dans la Section 2.4, nous rappelons sans preuve un r´esultat d’´equivalence des ensembles.

La robustesse de l’approche de Varadhan pour faire face aux syst`emes non gradient a ´et´e largement confirm´e par des mod`eles dans lesquels les quantit´es conserv´ees sont fonctionnelles lin´eaires des coordonn´ees du syst`eme, en d’autres termes, les ensembles invariants sont des hyperplans (cf [KLO], [Q]). D’apr`es (0.1.3), nous pouvons voir que cette propri´et´e n’est pas satisfaite pour le mod`ele qui nous int´eresse. En effet ici les en-sembles invariants sont des hypersurfaces. Ceci introduit des difficult´es suppl´ementaires de nature g´eom´etrique, car la m´ethode de Varadhan implique des manipulations sur des champs de vecteurs sur les hyperplans invariants (ou plus g´en´eralement, sur les hyper-surfaces invariantes). Ces difficult´es apparaissent dans les Sections 1.4, 2.1 et 2.3 , o`u des explications plus d´etaill´ees sont donn´ees.

(6)

0.2. INTRODUCTION iv

0.2

Introduction

A central problem in classical statistical mechanics is to provide a bridge between the macroscopic (thermodynamics) and microscopic (classical mechanics) description of the different phenomena observed in physical systems.

The ideal approach would be to start from a microscopic model of many components interacting with realistic forces and evolving with Newtonian dynamics, and then to de-rive some collective phenomena like the hydrodynamical regime or fluctuation-dissipation relations. For the moment, a rigorous mathematical derivation from deterministic mi-croscopic models seems outside the range of the mathematical techniques.

During the last decades the study of stochastic lattice systems, where particles in-teract randomly, has been largely exploited, basically because such systems may ex-hibit phenomena analogous to that of real physical systems (e.g. hydrodynamical equa-tions, fluctuations-dissipation theorems and metastable states) with precise mathemat-ical properties as counterpart of such phenomena (law of large numbers, central limit theorem and large deviation results).

The major breakthrough in the study of hydrodynamics for stochastic lattice systems appears with the work of Guo, Papanicolau, and Varadhan [GPV] where, by mean of intensive uses of ideas of large deviations, the authors derive the hydrodynamic behavior of the Ginzburg-Landau model. They consider the discrete torus TN = {0, · · · , N − 1}

and represent by xi the charge at site i. The evolution in time of the vector (x1, · · · , xN)

is given by a diffusion with infinitesimal generator LN =

1 2

X

i∈TN

D2i,i+1− (φ0(xi) − φ0(xi+1))Di,i+1,

where Di,j = ∂xi − ∂xi+1 and φ is a continuously differentiable function such that

R∞

−∞e

−φ(x)dx = 1, R∞

−∞e

λx−φ(x)dx < ∞ for all λ ∈ R and R∞

−∞e

σ|φ0(x)|−φ(x)dx < ∞

for all σ > 0.

The generator LN defines a diffusion process with invariant measure ⊗Ni=1e

−φ(xi)dx

i,

which is not ergodic because the hyperplanes x1 + · · · + xN = N y of average charge y

are invariant sets for all y ∈ R. Nevertheless, the restriction of the diffusion to each of such hyperplanes is nondegenerate and ergodic.

The arguments used in [GPV] provide a robust method to derive the hydrodynamic behavior for a large class of systems. In the beginning it seemed that this method could only be applied to the so-called gradients systems, that is, systems in which the

(7)

instantaneous current can be expressed as the discrete gradient of a local function. While this simplifies the proofs considerably, it is not a natural condition. The study of nongradient systems is of great interest because from a physical point of view they provide more realistic models, and from the mathematical point of view gradient system form only a set of low dimension (in some sense) in the space of stochastic reversible lattice models with local conservation laws (see [W] and references therein).

In a later paper S.R.S. Varadhan [V] managed to apply the method for the following nongradient perturbation of the Ginzburg-Landau model

LN =

1 2

X

i∈TN

{Di,i+1[a(xi, xi+1)Di,i+1] − (φ0(xi) − φ0(xi+1))a(xi, xi+1)Di,i+1},

where a(xi, xi+1) is a function satisfying 0 < c ≤ a(xi, xi+1) ≤ C with bounded

continu-ous first derivatives.

The arguments used in [V] permit to extend the entropy method to reversible non-gradient systems, provided the generator of the system restricted to a cube of size l has a spectral gap that shrinks as l−2.

At this time, the more general method to deal with nongradient systems is the one developed in (cf [V]), where roughly speaking, the idea is to replace the current by a gradient plus a fluctuation term.

On the other hand, in recent works, a microscopic model for heat conduction in solids had been considered (cf [BO], [B],[FNO]). In this model nearest neighbor atoms interact as coupled oscillators forced by an additive noise which exchange kinetic energy between nearest neighbors.

More precisely, in the case of periodic boundary conditions, atoms are labeled by x ∈ TN = {0, · · · , N − 1}. The configuration space is defined by ΩN = (R × R)TN, where

for a typical element (px, rx)x∈TN ∈ Ω

N, r

x represents the distance between particles x

and x + 1, and px the velocity of particle x. The formal generator of the system reads

as LN = AN + SN, where AN = X x∈TN {(px+1− px)∂rx+ (rx− rx−1)∂px} , (0.2.1) and SN = 1 2 X x∈TN Xx,x+1[Xx,x+1] , (0.2.2)

(8)

0.2. INTRODUCTION vi

where Xx,y = py∂px − px∂py. Here AN is the Liouville operator of a chain of interacting

harmonic oscillators and SN is the noise operator.

Denote by {(p(t), r(t)), t ≥ 0} the Markov process generated by N2LN (the factor

N2 corresponds to an acceleration of time). Let C(R

+, ΩN) be the space of continuous

trajectories. Fixed a time T > 0, and for a given measure µN on ΩN, the probability

measure on C([0, T ], ΩN) induced by this Markov process starting in µN will be denoted

by PµN. As usually, expectation with respect to PµN will be denoted by EµN.

In this work we focus on the noise operator SN which acts only on velocities, so we

restrict the configuration space to RTN. The total energy of the configuration (p

x)x∈TN is defined by E = 1 2 X x∈TN p2x . (0.2.3)

It is easy to check that SN(E ) = 0, i.e total energy is constant in time.

As in [GPV], the generator SN defines a diffusion process with invariant measure

⊗N i=1 1 √ 2πe −x2

i/2dxi, which is not ergodic because the hyperspheres p2

1+ · · · + p2N = N y of

average kinetic energy y are invariant sets for all y > 0. Nevertheless, the restriction of the diffusion to each of such hyperspheres is nondegenerate and ergodic.

The dynamics induced by SN turns to be a gradient system, in fact, the microscopic

instantaneous current of energy between x and x + 1 can be expressed as Wx,x+1 = 1

2(px+1 − px).

In analogy to [V] we introduce a coefficient into the generator (0.2.2) to break the gradient structure, namely

LN = 1 2 X x∈TN Xx,x+1[a(px, px+1)Xx,x+1] , (0.2.4)

where a(x, y) is a differentiable function satisfying 0 < c ≤ a(x, y) ≤ C < ∞ with bounded continuous first derivatives.

The collective behavior of the system can be described thanks to empirical measures. The energy empirical measure associated to the process is defined by

πtN(ω, du) = 1 N X x∈TN p2x(t)δx N(du) , (0.2.5)

where δz represents the Dirac measure concentrated on z.

Denote by YN

(9)

acts on smooth functions H : T → R as YtN(H) = √1 N X x∈TN H(x/N ){p2x(t) − y2} .

The principal result of this work states the convergence in distribution of YN

t (H) as

N goes to infinity.

In order to study the equilibrium fluctuations of interacting particle systems, Brox and Rost [BR] introduced the Boltzmann-Gibbs principle and proved their validity for attractive zero range process. Chang and Yau [CY] proposed an alternative method to prove the Boltzmann-Gibbs principle for gradient systems. This approach was extended to nongradient systems by Lu [Lu] and Sellami [Se].

By adapting the method introduced in [V] we identify the correct diffusion term (Sec-tion 1.4), which allows us to derive the Boltzmann-Gibbs principle (Sec(Sec-tion 1.3). This is the key point to show that the energy fluctuation field converges in the sense of fi-nite dimensional distributions to a generalized Ornstein-Uhlenbeck process (Section 1.2). Moreover, using again the Boltzmann-Gibbs principle we also prove tightness for the en-ergy fluctuation field in a specified Sobolev space (Section 1.5), which together with the finite dimensional convergence implies the convergence in distribution to the generalized Ornstein-Uhlenbeck process mentioned above. The idea of using the Boltzmann-Gibbs principle to prove tightness comes from [CLO].

The second chapter is devoted to more technical details. In Section 2.3 the space Hy is introduced an a characterization of this space is used into the proof of Theorem

5. This characterization relies basically on some integrability conditions for a Poisson system and a sharp spectral gap estimate for the generator of the process. This two subjects are treated in Section 2.1 and Section 2.2, respectively. Finally, in Section 2.4, we state without proof an equivalence of ensembles result.

The robustness of the Varadhan’s approach to deal with nongradient systems has been largely corroborated for models in which the conserved quantities are linear func-tionals of the coordinates of the system, in other words, the invariant sets are hyper-planes (cf [KLO], [Q]). After (0.2.3), we can see that this property is not satisfied for the model we are interested in (here the invariant sets are hyperspheres). This fact in-troduces additional difficulties of geometric nature because Varadhan’s method involve manipulations on vector fields on the invariant hyperplanes (or more general, on the invariant hypersurfaces). This difficulties will appear in sections 1.4, 2.1 and 2.3.

(10)
(11)

Equilibrium Fluctuations

1.1

Notations and Results

We will now give a precise description of the model. We consider a system of N particles in one dimension evolving under an interacting random mechanism. It is assumed that the spatial distribution of particles is uniform, so that the state of the system is given by specifying the N velocities.

For a positive integer N , denote by TN the lattice torus of length N : TN =

{0, 1, · · · , N − 1}. The configuration space is denoted by ΩN = RTN and a typical

configuration is denoted by p = (px)x∈TN where px represents the velocity of the particle

x. The velocity configuration p changes with time and, as a function of time undergoes a diffusion in RN.

The diffusion mentioned above have as infinitesimal generator the following operator LN = 1 2 X x∈TN Xx,x+1[a(px, px+1)Xx,x+1], (1.1.1)

where Xx,y = py∂px − px∂py, a(x, y) is a differentiable function satisfying 0 < c ≤

a(x, y) ≤ C < ∞ with bounded continuous first derivatives and all the sums are taken modulo N . Observe that the total energy defined as EN = 12 P

x∈TNp

2

x

satis-fies LN(EN) = 0, i.e total energy is a conserved quantity.

Let us consider for every y > 0 the Gaussian product measure νN

y on ΩN with density

(12)

1.1. NOTATIONS AND RESULTS 2

relative to the Lebesgue measure given by

ΦyN(p) = N Y i=1 e− p2x 2y2 √ 2πy, where p = (p1, p2, . . . , pN). Denote by L2N

y ) the Hilbert space of functions f on ΩN such that νyN(f2) < ∞.

LN is formally symmetric on L2(νyN) . In fact, is easy to see that for smooth functions

f and g in a core of the operator LN, we have for all y > 0

Z RN Xx,z(f )gΦyNdp = − Z RN Xx,z(g)f ΦyNdp, and therefore, Z RN LN(f )gΦyNdp = 1 2 X x∈TN Z RN Xx,x+1[a(px, px+1)Xx,x+1(f )]gΦyNdp = −1 2 X x∈TN Z RN Xx,x+1(f )[a(px, px+1)Xx,x+1(g)]ΦyNdp = Z RN f LN(g)ΦyNdp.

In particular, the diffusion is reversible with respect to all the invariant measures νyN. On the other hand, for every y > 0 the Dirichlet form of the diffusion with respect to νN y is given by DN,y(f ) = h−LN(f ), f iy = 1 2 X x∈TN Z RN a(px, px+1)[Xx,x+1(f )]2ΦyNdp , (1.1.2)

where h·, ·iy stands for the inner product in L2N y ).

Denote by {p(t), t ≥ 0} the Markov process generated by N2L

N (the factor N2

correspond to an acceleration of time). Let C(R+, ΩN) be the space of continuous

trajectories on the configuration space. Fixed a time T > 0, and for a given measure µN on ΩN, the probability measure on C([0, T ], ΩN) induced by this Markov process starting in µN will be denoted by PµN. As usually, expectation with respect to PµN will

(13)

The diffusion generated by N2LN can also be described by the following system of

stochastic differential equations

dpx(t) =

N2

2 {Xx,x+1[a(px, px+1)]px+1− Xx−1,x[a(px−1, px)]px−1− px[a(px, px+1) + a(px−1, px)]}dt + N [px−1

p

a(px−1, px)dBx−1,x− px+1

p

a(px, px+1)dBx,x+1],

where {Bx,x+1}x∈TN are independent standard Brownian motion.

Then, by Ito’s formula

dp2x(t) = N2[Wx−1,x− Wx,x+1]dt + N [σ(px−1, px)dBx−1,x(s) − σ(px, px+1)dBx,x+1(s)], where, Wx,x+1 = a(px, px+1)(p2x− p 2 x+1) − Xx,x+1[a(px, px+1)]pxpx+1 , (1.1.3) and, σ(px, px+1) = 2pxpx+1 p a(px, px+1) . (1.1.4)

We can think of Wx,x+1as being the instantaneous microscopic current of energy between

x and x + 1.

The collective behavior of the system is described thanks to empirical measures. With this purpose let us introduce the energy empirical measure associated to the process defined by πtN(ω, du) = 1 N X x∈TN p2x(t)δx N(du) ,

where δu represents the Dirac measure concentrated on u.

To investigate equilibrium fluctuations of the empirical measure πN we introduce

the empirical energy fluctuation field YN

t , a linear functional which acts on smooth

functions H : T → R as YtN(H) = √1 N X x∈TN H(x/N ){p2x(t) − y2} .

(14)

1.1. NOTATIONS AND RESULTS 4

and covariances given by

E[Yt(H1)Ys(H2)] = 4y4 p4π(t − s)ˆa(y) Z T du Z R dv ¯H1(u) exp  − (u − v) 2 4(t − s)ˆa(y)  ¯ H2(v) , (1.1.5) for every 0 ≤ s ≤ t. Here ¯H1(u) ( resp ¯H2(u)) is the periodic extension to the real line of

the smooth function H1 (resp H2), and ˆa(y) is the diffusion coefficient determined later

in Section 1.4.

Let us consider for k > 32 the Sobolev space H−k whose definition will be given at

the beginning of Section 1.5. Denote by QN the probability measure on C([0, T ], H−k)

induced by the energy fluctuation field YtN and the Markov process {pN(t), t ≥ 0} defined at the beginning of this section, starting from the equilibrium probability measure νyN. Let Q be the probability measure on the space C([0, T ], H−k) corresponding to the

generalized Ornstein-Uhlenbeck process Yt defined above.

We are now ready to state the main result of this work.

Theorem 1. The sequence of the probability measures {QN}N ≥1 converges weakly to the

probability measure Q .

The proof of Theorem 1 will be divided into two parts. On the one hand, in Section 1.5 we prove tightness of {QN}N ≥1 , where also a complete description of the space H−k

is given. On the other hand, in Section 1.2 we prove the finite-dimensional distribution convergence. These two results together imply the desired result. Let us conclude this section with a brief description of the approach we follow.

Given a smooth function G : T × [0, T ] → R we compute 1 √ N X x∈TN G(x N, t)p 2 x(t) = 1 √ N X x∈TN G(x N, 0)p 2 x(0) + Z t 0 1 √ N X x∈TN ∂sG( x N, 0)p 2 x(s)ds + Z t 0 N32 X x∈TN [G(x + 1 N , s) − G( x N, s)]Wx,x+1(s)ds + Z t 0 √ N X x∈TN [G(x + 1 N , s) − G( x N, s)]σ(px, px+1)dBx,x+1(s) .

(15)

Thus, MNG(t) = YtN(Gt) − YtN(G0) − Z t 0 YtN(∂sGs)ds + Z t 0 √ N X x∈TN ∇NG( x N, s)Wx,x+1(s)ds , (1.1.6)

where the left hand side is the martingale,

MNG(t) = √1 N X x∈TN Z t 0 ∇NG( x N, s)σ(px, px+1)dBx,x+1(s) , whose quadratic variation is given by

hMG Ni(t) = 1 N X x∈TN Z t 0 |∇NG( x N, s)| 2a(p x, px+1)p2xp2x+1ds .

Here ∇N denotes the discrete gradient of a function defined in TN. Recall that if G is

a smooth function defined on T and ∇ is the continuous gradient, then ∇NG( x N) = N [G( x + 1 N ) − G( x N)] = (∇G)( x N) + o(N −1 ). In analogy, ∆N denotes the discrete Laplacian, which satisfies

∆NG( x N) = N 2[G(x + 1 N ) − 2G( x N) + G( x − 1 N )] = (∆G)( x N) + o(N −1 ). where ∆ is the continuous Laplacian.

To close the equation for the martingale MG

N(t) we have to replace the term involving

the microscopic currents in (1.1.6) with a term involving YtN. Roughly speaking, what makes possible this replacement is the fact that non-conserved quantities fluctuates faster than conserved ones. Since the total energy is the unique conserved quantity of the system, it is reasonable that the only surviving part of the fluctuation field represented by the last term in (1.1.6) is its projection over the conservative field YN

t . This is the

content of the Boltzmann-Gibbs Principle (see [BR]).

Observe that the current Wx,x+1 cannot be written as the gradient of a local function,

neither by an exact fluctuation-dissipation equation, i.e as the sum of a gradient and a dissipative term of the form LN(τxh). That is, we are in the so called nongradient case.

(16)

1.2. CONVERGENCE OF THE FINITE-DIMENSIONAL DISTRIBUTIONS 6

In order to perform the replacement mentioned in the previous paragraph, we follow the approach proposed by S.R.S Varadhan in [V], which is briefly described in the following lines.

Denote by C0 the space of cylinder functions with zero mean with respect to all

canonical measures. On C0 and for each y > 0 a semi-inner product  · y is defined.

The Hilbert space Hy induced by C0 and  · y is seen to be generated by two

or-thogonal subspaces, namely, the linear space generated by the function p2

1− p20 and the

subspace L(C0). Here L stands for the natural extension of LN to Z.

Since the current W0,1 belong to C0 we have a decomposition of it in terms of this

two spaces. More precisely, there exists ˆa(y) ∈ R such that for all δ > 0 there exists f ∈ C0 such that

 W0,1+ ˆa(y)[p21− p 2

0] − L(f ) y≤ δ .

The terms corresponding to L(f ) being negligible permit to perform the desired replace-ment.

In fact, in this work we only consider this semi-inner product for the functions in C0

we are interested in (i.e W0,1, [p21− p20], and L(f )).

1.2

Convergence of the finite-dimensional

distribu-tions

To investigate equilibrium fluctuations of the empirical measure πN, we fix once for all

y > 0 and we denote by YN

t the empirical energy fluctuation field, a linear functional

which acts on smooth functions H : T → R as YtN(H) = √1

N X

x∈TN

H(x/N ){p2x(t) − y2} .

We state the main result of the section.

Theorem 2. The finite-dimensional distributions of the fluctuation field YN

t converges,

as N goes to infinity, to the finite-dimensional distributions of the generalized Ornstein-Uhlenbeck process Y defined in (1.1.5).

In this setting convergence of finite-dimensional distributions means that given a positive integer k, for every {t1, · · · , tk} ⊂ [0, T ] and every collection of smooth functions

(17)

{H1, · · · , Hk}, the vector YtN1 (H1), · · · , Y

N

tk(Hk) converges in distribution to the vector

(Yt1(H1), · · · , Ytk(Hk)).

Recall that from Itˆo’s formula we have YtN(H) = Y0N(H) − Z t 0 √ N X x∈TN ∇NH(x/N )Wx,x+1(s)ds − Z t 0 1 √ N X x∈TN ∇NH(x/N )σ(px(s), px+1(s)) dBx,x+1(s) .

Let us focus on the integral in the first line of the above expression. A central idea in the nongradient method proposed by Varadhan in [V], is to consider the current Wx,x+1

as an element of a Hilbert space generated by two orthogonal subspaces, one of which is the space generated by the gradient (in our case p2

x − p2x+1). Since non conserved

quantities fluctuates in a much faster scale than conserved ones, it is expected that just the component corresponding to the projection over the energy fluctuation field survives after averaging over space and time. Recall that total energy is the only conserved quantity of the evolution.

The idea is to use the last observations, together with the fact thatP

x∈TN∆NH(x/N ) =

0, in order to replace the integral term corresponding to the current Wx,x+1 by an

ex-pression involving the empirical energy fluctuation field, namely Rt

0 Y N

s (∆NH)ds.

Firstly, consider the empirical field YN

t acting now on time dependent smooth

func-tions H : T × [0, T ] → R as YtN(H) = √1 N X x∈TN H(x/N, t){p2x(t) − y2} .

Again from the Itˆo’s formula we have

YtN(Ht) = Y0N(H0) + Z t 0 YsN(∂sHs)ds − Z t 0 √ N X x∈TN ∇NHs(x/N )Wx,x+1(s)ds − Z t 0 1 √ N X x∈TN ∇NHs(x/N )σ(px(s), px+1(s)) dBx,x+1(s) . (1.2.1)

(18)

1.2. CONVERGENCE OF THE FINITE-DIMENSIONAL DISTRIBUTIONS 8

Now we proceed to rewrite the last expression as

YtN(Ht) = Y0N(H0) + Z t 0 YsN(∂sHs+ ˆa(y)∆NHs)ds − IN,F1 (H ·t ) − IN,F2 (H·t) − M1 N,F(H ·t ) − MN,F2 (H·t) , (1.2.2)

where F is a fixed smooth local function and,

IN,F1 (H·t) = Z t 0 √ N X x∈TN ∇NHs(x/N )Wx,x+1− ˆa(y)[p2x+1− p 2 x] − LNτxF ds , IN,F2 (H·t) = Z t 0 √ N X x∈TN ∇NHs(x/N )LNτxF ds , MN,F1 (H·t) = Z t 0 2 √ N X x∈TN ∇NHs(x/N )τx p a(p0, p1) " p0p1+ X0,1( X i∈TN τiF ) # dBx,x+1, MN,F2 (H·t) = Z t 0 2 √ N X x∈TN ∇NHs(x/N ) p a(px, px+1)Xx,x+1( X i∈TN τiF ) dBx,x+1(s) .

Here τxrepresents translation by x, and the notation H·tstress the fact that the functions

depends also on time.

The reason to rewrite expression (1.2.1) in this way, is that a judicious choice of the function H will cancel the integral of the fluctuation field term.

Denote by {St}t≥0the semigroup generated by the Laplacian operator ˆa(y)∆. Given

t > 0 and a smooth function H : T → R, define H(·, s) = St−sH for 0 ≤ s ≤ t. As is

well known, this function satisfies the following properties:

∂sHs + ˆa(y)∆Hs = 0 , (1.2.3) h Hs , ˆa(y)∆Hs i = − 1 2 d dsh Hs, Hsi , (1.2.4) where h·, ·i stands for the usual inner product in L2(T).

Remark 1. The smoothness of H implies the existence of a constant C > 0 such that

|YN s (∆Hs) − YsN(∆NHs)| ≤ C √ N 1 N X x∈TN p2x(s) ! ,

(19)

Therefore, we obtain for all smooth functions H : T → R YtN(H) = Y0N(StH) + O( 1 N)−I 1 N,F(H ·t )−IN,F2 (H·t)−MN,F1 (H·t)−MN,F2 (H·t) , (1.2.5) where O(1

N) denotes a function whose L

2 norm is bounded by C/N for a constant C

depending just on H.

Let {Fk}k≥1 be a sequence of local functions belonging to the Schwartz space, such

that

lim

k→∞ˆa(y, Fk) = y

4ˆa(y) (1.2.6)

where ˆa(y, F ) = Eνy[a(p0, p1)(p0p1+ X0,1( eF ))

2]. The functions eF and ˆa(y) are defined

in (2.3.1) and (2.3.3), respectively.

The proof of Theorem 2 will follow from the following three results.

Lemma 1. For every smooth function H : T → R and local function F in the Schwartz space, lim N →∞Eν N y  sup 0≤t≤T IN,F2 (H·t) + MN,F2 (H·t)2  = 0 . (1.2.7)

Lemma 2. The process MN,F1

k converges in distribution as k increases to infinity after

N , to a Gaussian process characterized by

lim k→∞N →∞lim Eν N y M 1 N,Fk(H ·t 1)MN,F1 k(H ·t 2)  = 4y4 Z t 0 Z T ˆ a(y)H10(x, s)H20(x, s) dxds , (1.2.8) for every smooth functions Hi : T → R for i = 1, 2.

Theorem 3 (Boltzmann-Gibbs Principle). For every smooth function H : T → R, lim k→∞N →∞lim Eν N y (I 1 N,Fk(H ·t )2 = 0 . (1.2.9)

The proofs of Lemma 1 and Lemma 2 are postponed to the end of this section, while the proof of Theorem 3 requires a section on its own.

Now we state some remarks. Firstly, the convergence in distribution of Y0N(H) to a Gaussian random variable with mean zero and variance 2y4hH, Hi as N tends to infinity, follows directly from the Lindeberg-Feller theorem.

Property (1.2.4) together with Lemma 2 imply the convergence in distribution as N increases to infinity after k of the martingale M1

N,Fk(H

·t) to a Gaussian random variable

with mean zero and variance 2y4hH, Hi − 2y4hS

(20)

1.2. CONVERGENCE OF THE FINITE-DIMENSIONAL DISTRIBUTIONS 10

Finally, observe that the martingale MN,F1

k(H

·t) is independent to the initial filtration

F0.

Proof of Theorem 2. For simplicity and concreteness in the exposition we will restrict ourselves to the two-dimensional case (YtN(H1), Y0N(H2)) . Similar arguments may be

given to show the general case.

In order to characterize the limit distribution of (YtN(H1), Y0N(H2)) it is enough to

characterize the limit distribution of all the linear combinations θ1YtN(H1) + θ2Y0N(H2).

From Lemma 1, Theorem 3 and expression 1.2.5 it follows

θ1YtN(H1) + θ2Y0N(H2) = θ1Y0N(StH1) + θ1IN(H1, Fk) − θ1MN,F1 k(H ·t 1) + θ2Y0N(H2) = Y0N(θ1StH1+ θ2H2) + θ1IN(H1, Fk) − θ1MN,F1 k(H ·t 1) , where IN(H, F

k) denotes a function whose L2 norm tends to zero as k increases to

infinity after N .

Thus, the random variable θ1YtN(H1) + θ2Y0N(H2) tends as N goes to infinity to a

Gaussian random variable with mean zero and variance 2y4{θ2

1hH1, H1i+2θ1θ2hStH1, H2i+

θ22hH2, H2i}. This in turn implies

EνN

y [Yt(H1)Y0(H2)] = 2y

4hS

tH1, H2i ,

which coincide with (1.1.5).

Now we proceed to give the proofs of Lemma 1 and Lemma 2. Proof of Lemma 1. Let us define

ζN,F(t) = 1 N3/2 X x∈TN ∇NH(x/N, t)τxF (p(t)) .

From the Itˆo’s formula we obtain ζN,F(t) − ζN,F(0) = 1 2I 2 N,F(H ·t) + Z t 0 1 N3/2 X x∈TN ∂t∇NHs(x/N )τxF (p)ds + Z t 0 1 √ N X x∈TN ∇NHs(x/N ) X z∈TN p a(pz, pz+1)Xz,z+1(τxF ) dBz,z+1(s) .

(21)

Then, (IN,F2 (H·t) + MN,F2 (H·t))2 is bounded above by 6 times the sum of the following three terms (ζN,F(t) − ζN,F(0))2, Z t 0 1 N3/2 X x∈TN ∂t∇NHs(x/N )τxF (p)ds !2 , 1 2M 2 N,F(H ·t ) − Z t 0 1 √ N X x∈TN ∇NHs(x/N ) X z∈TN p a(pz, pz+1)Xz,z+1(τxF ) dBz,z+1(s) !2 .

Since F is bounded and H is smooth, the first two terms are of order N1. Using Doob’s inequality and the fact that F is local, we can show that the expectation of the sup0≤t≤T of the third term is also of order N1.

Proof of Lemma 2. Using basic properties of the stochastic integral and the stationarity of the process, we can see that the expectation appearing in 1.2.8 is equal to

4 Z t 0 1 N X x∈TN ∇NH1,s(x/N )∇NH2,s(x/N )EνN y  τxa(p0, p1) p0p1+ X0,1( X i∈TN τiFk) !2 ds .

Translation invariance of the measure νyN lead us to

4 EνyN  a(p0, p1) p0p1+ X0,1( X i∈TN τiFk) !2  Z t 0 1 N X x∈TN ∇NH1,s(x/N )∇NH2,s(x/N )ds ,

and as N goes to infinity we obtain

4 Eνy  a(p0, p1) p0p1+ X0,1( X i∈Z τiFk) !2  Z t 0 Z T H10(u, s)H20(u, s) du ds .

(22)

1.3. BOLTZMANN-GIBBS PRINCIPLE 12

1.3

Boltzmann-Gibbs Principle

The aim of this section is to provide a proof for Theorem 3. In fact, we will prove a stronger result that will be also useful in the proof of tightness. Namely,

lim k→∞N →∞lim Eν N y   sup 0≤t≤T Z t 0 √ N X x∈TN ∇NHs(x/N )[Vx(p(s)) − LNτxFk(p(s))] ds !2  = 0 , where, Vx(p) = Wx,x+1(p) − ˆa(y)[p2x+1− p 2 x] .

We first localize the problem. Fix an integer M that shall increase to infinity after N , and subdivide the torus TN in non overlapping blocks of size M . Being l and r the

integers satisfying N = lM + r with 0 ≤ r < M , define for j = 1, · · · , l Bj = {(j − 1)M, · · · , jM } ,

B0j = {(j − 1)K, · · · , jM − 1} , Bkj = {(j − 1)K, · · · , jM − sk} ,

where sk is the size of the block supporting Fk. Denote the remaining block by Bl+1 =

{lK + 1, · · · , N }.

With this notation we can write √ N X x∈TN ∇NHs(x/N )[Vx(p(s)) − LNτxFk(p(s))] = V1+ V2+ V3 , with, V1 = √ N X x∈Bl+1 ∇NHs(x/N )Vx − √ N X x∈Bl+1 ∇NHs(x/N )LNτxFk, V2 = √ N l X j=1 ∇NHs(jK/N )VjM − √ N l X j=1 X x∈Bj\Bkj ∇NHs(x/N )LNτxFk , V3 = √ N l X j=1 X x∈B0j ∇NHs(x/N )Vx− √ N l X j=1 X x∈Bk j ∇NHs(x/N )LNτxFk.

(23)

There-fore, in order to prove Theorem 3 it suffices to prove lim k→∞N →∞lim Eν N y " sup 0≤t≤T Z t 0 Vi ds 2# = 0 , (1.3.1)

for each Vi separately.

Before giving the proof, we introduce some notations and state some general results for time continuous Markov processes {Xt}t≥0 with a stationary measure π. Denote by

h , iπ the inner product in L2(π) and by L : D(L) ⊂ L2(π) → L2(π) the infinitesimal

generator of this process.

Define for f ∈ D(L) ⊂ L2(π),

||f ||21 = hf, (−LN)f iπ . (1.3.2)

It is easy to see that || · ||1 is a norm in D(L) that satisfies the parallelogram rule, and

therefore, that can be extended to an inner product in D(L). We denote by H1 the

completion of D(L) under the norm || · ||1, and by h , i1 the induced inner product. In

the same way define

||f ||2

−1 = sup

g∈D(L)

{2 hf, giπ− hg, gi1} , (1.3.3) and denote by H−1 the completion with respect to || · ||−1 of the set of functions in L2(π)

satisfying ||f ||−1 < ∞. Later we state some well known properties of this spaces.

Lemma 3. For f ∈ L2(π) ∩ H1 and g ∈ L2(π) ∩ H−1, we have

i) ||g||−1 = suph∈D(L)\{0} hh,giπ

||h||1 ,

ii) | hf, giπ| ≤ ||f ||1||g||−1 .

Property i) implies that H−1 is the dual space of H1 with respect to L2(π), and

property ii) entails that the inner product h , iπ can be extended to a continuous bilinear form on H−1× H1. The preceding results remain in force when L2(π) is replaced by any

Hilbert space.

The following is a very useful estimate of the time variance in terms of the H−1 norm.

Proposition 1. Given T > 0 and a mean zero function V ∈ L2(π) ∩ H −1, Eπ " sup 0≤t≤T Z t 0 V (ps) ds 2# ≤ 24T ||V ||2 −1 .

(24)

1.3. BOLTZMANN-GIBBS PRINCIPLE 14

See Lemma 2.4 in [KmLO] for a proof .

Remark 2. A slightly modification in the proof given in [KmLO] permit to conclude that, for every smooth function h : [0, T ] → R,

Eπ " sup 0≤t≤T Z t 0 h(s)V (ps) ds 2# ≤ Ch||V ||−1 , where, Ch = 6{4||h||2∞T2+ ||h 0||2 ∞T3} .

Moreover, in our particular case we have

EνN y   sup 0≤t≤T Z t 0 l X j=1 hj(s)VBj(ps) ds !2  ≤ l X j=1 Chj||VBj||−1, for functions {VBj} l

j=1 depending on disjoint blocks.

The proof of (1.3.1) will be divided in three lemmas. Lemma 4. lim k→∞N →∞lim Eν N y " sup 0≤t≤T Z t 0 V1 ds 2# = 0 .

Proof. By Proposition 1, the expectation in the last expression is bounded above by the sum of the following three terms

3ˆa(y)2C H|Bl+1| N X x∈Bl+1 p2 x+1− p 2 x , (−LN)−1p2x+1− p 2 x , 3CH|Bl+1| N X x∈Bl+1 Wx,x+1 , (−LN)−1Wx,x+1 , 3CH|Bl+1| N X x∈Bl+1 h(−LN)τxFk, τxFki .

Here CH represents a constant depending on H and T , that can be multiplied by a

constant from line to line.

Observe that the second term is less than or equal to CH|Bl+1|t N X x∈Bl+1 sup g∈L2N y) {2 hWx,x+1, gi + hg, Lx,x+1gi} ,

(25)

where,

Lx,x+1 =

1

2Xx,x+1[a(px, px+1)Xx,x+1] .

From the definition given in (1.1.3) we have Wx,x+1 = −Xx,x+1[a(px, px+1)pxpx+1].

Per-forming integration by parts in the two inner products, we can write the quantity inside braces as

1

2g∈Lsup2N y )

{4 ha(px, px+1)pxpx+1, Xx,x+1gi − hXx,x+1g, a(px, px+1)Xx,x+1gi} ,

which by the elementary inequality 2ab ≤ A−1a2+ Ab2, is bounded above by 2a(px, px+1)p2xp2x+1

. Then, the second term is bounded above by

CH|Bl+1|2

N .

The same is true for the first term. In fact, observe that the first term is just the second one with a(x, y) ≡ 1.

Since Fk is a local function supported in a box of size sk and νyN is translation

invariant, we have for all x, y ∈ TN

hτxFk, (−LN)τyFki ≤ sk ||X0,1(fFk)||2L2N y) ,

which implies that the third term is bounded by CH|Bl+1|2

N sk||X0,1(fFk)||

2 L2N

y),

ending the proof. Lemma 5. lim k→∞N →∞lim Eν N y " sup 0≤t≤T Z t 0 V2 ds 2# = 0 . Proof. The proof is similar to the preceding one.

Lemma 6. lim k→∞N →∞lim Eν N y " sup 0≤t≤T Z t 0 V3 ds 2# = 0 .

(26)

1.3. BOLTZMANN-GIBBS PRINCIPLE 16

Proof. Recall that the expectation in the last expression is by definition

N EνN y   sup 0≤t≤T   Z t 0 l X j=1    X x∈B0j ∇NHs(x/N )Vx− X x∈Bk j ∇NHs(x/N )LNτxFkds      2 .

The smoothness of the function H allows to replace ∇NHs(x/N ) into each sum in the

last expression by ∇NHs(x∗j/N ), where x ∗

j ∈ Bj (for instance, take x∗j = (j − 1)K + 1 )

, obtaining N EνN y   sup 0≤t≤T   Z t 0 l X j=1 ∇NHs(x∗j/N )    X x∈B0 j Vx− X x∈Bk j LNτxFk ds      2 .

By proposition 1 and Remark 2, the quantity in the preceding line is bounded above by CH N l X j=1 * X x∈B0 j Vx− X x∈Bk j LNτxFk, (−LN)−1 X x∈B0 j Vx− X x∈Bk j LNτxFk + νy ,

Using the variational formula for the H−1 norm given in (1.3.3) and the convexity of the

Dirichlet form, we are able to replace (−LN)−1 by (−LBj0)−1 in the expression above. In

addition, by translation invariance of the measure νyN we can bound this expression by

CH l N * X x∈B10 Vx− X x∈Bk 1 LNτxFk, (−LB01)−1 X x∈B01 Vx− X x∈Bk 1 LNτxFk + νy .

By the strong law of numbers and the fact that l

N ∼

1

M, the lim supN →∞ of the last

expression is bounded above by

CH lim sup M →∞ 1 M * X x∈B0 1 Vx− X x∈Bk 1 LNτxFk, (−LB0 1) −1 X x∈B0 1 Vx− X x∈Bk 1 LNτxFk + νM,√ M y .

The limit as k goes to infinity of the last term is equal to zero, as will be justified in the paragraph following the statement of Theorem 5.

(27)

1.4

Central Limit Theorem Variances and Diffusion

Coefficient

Consider a continuous Markov process {Ys}s≥0, reversible and ergodic with respect to

the invariant measure π. Denote by h , iπ the inner product in L2(π) and let us suppose

that the infinitesimal generator of this process L : D(L) ⊂ L2(π) → L2(π) is a negative

operator.

Let V ∈ L2(π) be a mean zero function on the state space of the process. The central limit theorem proved in [KV] for

Xs=

Z t

0

V (Ys)ds ,

states that if V is in the range of (−L)12, then there exists a square integrable martingale

{Mt}t≥0 with stationary increments such that

lim t→∞ 1 √ t0≤s≤tsup |Xs− Ms| = 0 .

This result in turn implies that the limiting variance σ2(V, π) is given by

σ2(V, π) = lim t→∞ 1 tE[X 2 t] = 2hV , (−L) −1 V iπ . (1.4.1)

Observe that in the particular case when V = L(U ) for some U ∈ D(L) we have

σ2(V, π) = 2hU , (−L)U iπ . (1.4.2)

The right hand side of the above equality correspond to twice the Dirichlet form as-sociated to the generator L and the measure π, evaluated in U . Moreover, in terms of the norms defined in (1.3.2) and (1.3.3), the right hand side of (1.4.1) and (1.4.2) corresponds to

σ2(V, π) = 2||V ||−1 ,

σ2(U, π) = 2||U ||1 .

(28)

1.4. CENTRAL LIMIT THEOREM VARIANCES AND DIFFUSION COEFFICIENT 18 generator defined by LN(f ) = 1 2 X −N ≤x≤x+1≤N Xx,x+1[a(px, px+1)Xx,x+1(f )] ,

and let µN,y be the uniform measure on the sphere

( (p−N, · · · , pN) ∈ R2N +1 : N X i=−N p2i = (2N + 1)y2 ) .

The Dirichlet form DN,y associated to this measure is given by,

DN,y(f ) = 1 2 X −N ≤x≤x+1≤N Z a(px, px+1)[Xx,x+1(f )]2µN,y(dp) .

Is not difficult to see that the measures µN,y are ergodic for the process with generator

LN. We are interested in the asymptotic behavior of the variance

σ2(BN +ba(y)AN − H F N , µN,y) , (1.4.3) where, AN(p−N, · · · , pN) = p2N − p 2 −N , BN(p−N, · · · , pN) = X −N ≤x≤x+1≤N Wx,x+1 , HNF(p−N, · · · , pN) = X −N ≤x−k≤x+k≤N LN(τxF ) ,

with F (x−k, · · · , xk) a smooth function of 2k + 1 variables. Observe that these three

classes of functions are sums of translations of local functions, and have mean zero with respect to every µN,y. Finally, we denote this variance by ∆N,y(BN+ba(y)AN− H

F N) and

the inner product in L2

N,y) by h , iN,y.

Observe that the functions BN and HNF belong to the range of LN, in fact

BN(p−N, · · · , pN) = LN( N X x=−N xp2x) , (1.4.4) HNF(p−N, · · · , pN) = LN(ψNF) , (1.4.5)

(29)

where,

ψNF = X

−N ≤x−k≤x+k≤N

τxF .

This, in particular, implies that the central limit theorem variances and covariances involving BN and HNF exist, moreover, after (1.4.2) they are easily computable.

In the remaining of this section we deal with integration with respect to µN,y. Then,

before entering in our specific context, let us state some classical results referring to integration over spheres. The proofs of these and other results concerning integration over spheres, can be founded in [Ba].

Lemma 7. Given p = (p1, · · · , pN), a = (a1, · · · , aN), a = Pn k=1ak and Sn(r) = {p ∈ Rn: |p| = r} , define E(p, a) = n Y k=1 (x2k)ak, and S N(a, r) = Z Sn(r) E(p, a)dσ then, SN(a, r) = 2Qn k=1Γ(ak+ 1 2) Γ(a + N2) r 2a+N −1.

Corollary 1. There exist a constant C depending on y and the lower bound of a(·, ·) such that, for every u ∈ D(L)

hu, ANiN,y ≤ C(2N ) 1 2D N,y(u) 1 2 .

Proof. Observe that

AN(p−N, · · · , pN) = N −1

X

x=−N

(30)

1.4. CENTRAL LIMIT THEOREM VARIANCES AND DIFFUSION COEFFICIENT 20 then, hu, ANiN,y = N −1 X x=−N Z u(p)Xx,x+1(pxpx+1)µN,y(dp) = N −1 X x=−N Z Xx,x+1(u)pxpx+1µN,y(dp) ≤ Z N −1 X x=−N |pa(px, px+1)Xx,x+1(u)| |pxpx+1| pa(px, px+1) µN,y(dp) ≤ Z N −1 X x=−N a(px, px+1)(Xx,x+1(u))2 !12 N −1 X x=−N |pxpx+1|2 a(px, px+1) !12 µN,y(dp) ≤ Z N −1 X x=−N |pxpx+1|2 a(px, px+1) µN,y(dp) !12 Z N −1 X x=−N

a(px, px+1)(Xx,x+1(u))2µN,y(dp)

!12

≤ C(2N )12DN,y(u) 1 2 .

This, in particular, implies that the central limit theorem variances and covariances involving AN exist. In spite of that, the core of the problem will be to deal with the

variance of AN which is not easily computable.

Corollary 2. Z N −1 X i=−N p2ip2i+1µN,y(dp) = 2N (2N + 1)2 (2N + 3)(2N + 1)y 4 .

Proof. By the rotation invariance of µN,y, the left hand side of (2) is equal to

2N Z

p2ip2i+1µN,y(dp) ,

which in turns, applying Lemma 7 with a = {1, · · · , 1} and r = y√2N + 1, is equal to

2N 2Γ( 1 2) 2N −1Γ(3 2) 2 Γ(2 + 2N +12 ) r 4+2N ! 2Γ( 1 2) 2N +1 Γ(2N +12 )r 2N !−1 .

(31)

Lemma 8. (Divergence Theorem) Let BN(r) = {p ∈ RN : |p| ≤ r} and SN −1(r) = {p ∈ Rn: |p| = r}, then for every continuously differentiable function f : RN → R,

r Z BN(r) ∂f ∂pi (p)dp = Z SN −1(r) f (s1, · · · , sN)sidσN −1,

where dσN −1 denotes surface integration over the sphere.

Corollary 3. Taking r = y√2N + 1 in Lemma 8, we have for −N ≤ i, j ≤ N Z Xi,j(W )pipjµN,y(dp) = r |S2N(r)| Z B2N +1(r)  pi ∂W ∂pi − pj ∂W ∂pj  dp .

Proof. Recalling that Xi,j = pj∂pi − pi∂pj and µN,y is the uniform probability measure

over the sphere of radius y√2N + 1 centered at the origin, we have that the left hand side in the preceding line is equal to

1 |S2N(r)| Z S2N(r) pj  pipj ∂W ∂pi  dσ2N − Z S2N(r) pi  pipj ∂W ∂pj  dσ2N  .

From Lemma 8, the expression above is equal to r |S2N(r)| Z B2N +1(r)  pi ∂W ∂pi − pipj ∂W ∂pi∂pj  dp − Z B2N +1(r)  pj ∂W ∂pj − pipj ∂W ∂pi∂pj  dp  ,

concluding the proof.

Corollary 3 is extremely useful for us, because it provide a way to perform certain telescopic sums over the sphere. In fact, it implies that given −N ≤ i < j ≤ N we have

Z Xi,j(W )pipjµN,y(dp) = j−1 X k=i Z Xk,k+1(W )pkpk+1µN,y(dp) . (1.4.6)

Now we return to the study of AN,BN and HNF. The next proposition entail the

asymptotic behavior of the central limit theorem variances and covariances involving BN or HNF.

(32)

1.4. CENTRAL LIMIT THEOREM VARIANCES AND DIFFUSION COEFFICIENT 22

Theorem 4. The following limits hold locally uniformly in y > 0.

i) lim N →∞ 1 2N∆N,y(BN, BN) = Eνy[a(p0, p1)(2p0p1) 2 ] , ii) lim N →∞ 1 2N∆N,y(H F N, H F N) = Eνy[a(p0, p1)[Xx,x+1( eF )] 2] , iii) lim N →∞ 1 2N∆N,y(AN, BN) = −Eνy[(2p0p1) 2] , iv) lim N →∞ 1 2N∆N,y(BN, H F N) = −Eνy[2p0p1a(p0, p1)Xx,x+1( eF )] , v) lim N →∞ 1 2N∆N,y(AN, H F N) = 0 , v) lim sup N →∞ 1 2N∆N,y(AN, AN) ≤ C , where eF is formally defined by

e F (p) = ∞ X j=−∞ τjF (p) ,

and C is a positive constant depending uniformly on y. Although eF is not really make sense, the gradients Xi,i+1( eF ) are all well defined.

Proof. i ) From (1.4.4) and (1.4.2) we have that

∆N,y(BN, BN) = 2DN,y N X x=−N xp2x ! = 4 Z N −1 X x=−N a(px, px+1)(pxpx+1)2µN,y(dp) .

Since µN,y is rotation invariant we have

1 2N∆N,y(BN, BN) = 4 Z a(p0, p1)p20p 2 1µN,y(dp) ,

(33)

ii ) From (1.4.5) and (1.4.2) we have that ∆N,y(HNF, HNF) = 2DN,y(ψNF) = Z N −1 X x=−N a(px, px+1)[Xx,x+1(ψNF)] 2µ N,y(dp) .

The sum in the last line can be broken into two sums, the first one considering the indexes in {−N + 2k, · · · , N − 2k − 1} and the second one considering the indexes in the complement with respect to {−N, · · · , N − 1}. From the conditions imposed over F , when divided by N , the term corresponding to the second sum tends to zero as N goes to infinity. Then,

lim N →∞ 1 2N∆N,y(H F N, H F N) = lim N →∞ Z 1 2N N −1 X x=−N a(px, px+1)[Xx,x+1(ψFN)] 2 µN,y(dp) = lim N →∞ Z 1 2N N −1 X x=−N τxg(p)µN,y(dp) , where, e F (p) = ∞ X j=−∞ τjF (p) , and g(p) = a(p0, p1)[Xx,x+1( eF )]2.

Once more time, the desired result comes from the rotation invariance of µN,y and

the equivalence of ensembles.

iii ) From (1.4.5), (1.4.2) and the fact that Wx,x+1 = −Xx,x+1[pxpx+1a(px, px+1)], we

have ∆N,y(BN, HNF) = 2 N −1 X x=−N Z Xx,x+1[pxpx+1a(px, px+1)]ψFNµN,y(dp) = −2 N −1 X x=−N Z pxpx+1a(px, px+1)Xx,x+1[ψNF]µN,y(dp) ,

where the last equality is due to an integration by parts. Then, we divide the sum appearing in the last line into two sums in the very same way as was done for ii), obtaining

(34)

1.4. CENTRAL LIMIT THEOREM VARIANCES AND DIFFUSION COEFFICIENT 24 lim N →∞ 1 2N∆N,y(BN, H F N) = −2 Z N −2k−1 X x=−N +2k τxh(p)µN,y(dp) , where h(p) = a(p0, p1)p0p1Xx,x+1( eF ).

Again, the result comes from rotation invariance of µN,y and the equivalence of

ensembles.

iv ) From (1.4.4), (1.4.2) and the fact that AN = XN,−N(pNp−N), we have

∆N,y(AN, BN) = −2 Z XN,−N(pNp−N) N X x=−N xp2x ! µN,y(dp) = 2 Z pNp−NXN,−N[N p2N − N p 2 −N]µN,y(dp) = 8N Z p2Np2−NµN,y(dp) = 8N (2N + 1) 2y4 (2N + 3)(2N + 1) .

The second and fourth equalities come from integration by parts and corollary 2, respec-tively. Thus, we have

lim

N →∞

1

2N∆N,y(AN, BN) = 4y

4.

v ) By the same arguments used in the preceding items, together with the observation after corollary 3, we have

∆N,y(AN, HNF) = −2 Z XN,−N(pNp−N)ψFNµN,y(dp) = 2 Z pNp−NXN,−N(ψFN)µN,y(dp) = 2 Z N −1 X i=−N

pipi+1Xi,i+1(ψFN)µN,y(dp) .

From rotation invariance of µN,y and the equivalence of ensembles we obtain

lim N →∞ 1 2N∆N,y(AN, H F N) = Eνy[p0p1X0,1( eF )] = 0 .

vi ) By duality ∆N,y(AN, AN) = 2c2 where c is the smallest constant such that for

every u ∈ D(L), hu, ANiN,y ≤ cDN,y(u) 1 2 . (1.4.7)

(35)

Corollary 1 says that C(2N )12 is a constant satisfying (1.4.7) for every u ∈ D(L),

there-fore, c is smaller than C(2N )12. Thus

1

2N∆N,y(AN, AN) ≤ 2C

2

Recall that C is a constant depending locally uniformly on y. That conclude the proof of Theorem 4.

We proceed now to calculate the only missing limit variance (the one corresponding to AN) in an indirect way, as follows.

Using the basic inequality

|∆N,y(AN, BN − HNF)|2 ≤ ∆N,y(AN, AN)∆N,y(BN − HNF, BN − HNF) ,

we obtain, lim inf N →∞ |∆N,y(AN,BN−HNF)|2 (2N )2 ∆N,y(BN−HFN,BN−HNF) 2N ≤ lim inf N →∞ ∆N,y(AN, AN) 2N ,

which in view of Theorem 4 implies (4y4)2 Eνy[a(p0, p1){2p0p1+ X0,1( eF )}2] ≤ lim inf N →∞ ∆N,y(AN, AN) 2N . (1.4.8)

Let us define ˆa(y) by the relation ˆ

a(y) = y−4inf

F a(y, F ) ,

where the infimum is taken over all local smooth functions, and a(y, F ) = Eνy[a(p0, p1)(p0p1+ X0,1( eF ))

2] .

Since the limit appearing in (1.4.8) does not depend of F , we have

lim inf N →∞ ∆N,y(AN, AN) 2N ≥ 4y4 b a(y) . (1.4.9)

Moreover, this limit is locally uniform in y.

(36)

1.4. CENTRAL LIMIT THEOREM VARIANCES AND DIFFUSION COEFFICIENT 26

Theorem 5. The function ba(y) is continuous in y > 0 and

lim N →∞ 1 2N∆N,y(AN, AN) = 4y4 ˆ a(y) . (1.4.10)

Before entering in the proof of Theorem 5, let us show how to conclude from this results the proof of lemma 6 in the preceding section. In fact, for the variance we are interested in we have that

∆N,y(BN +ba(y)AN − H

F

N) = (ba(y))

2

N,y(AN, AN) + ∆N,y(BN, BN) + ∆N,y(HNF, H F N)

−2ba(y)∆N,y(AN, BN) + ∆N,y(AN, HNF) − ∆N,y(BN, HNF)

. Thanks to Theorem 4 and Theorem 5, if we divide by 2N and take the limit as N goes to infinity at both sides of last expression, we can conclude that

lim N →∞ 1 2N∆N,y(BN +ba(y)AN − H F N) = 4a(y, 1/2F ) − 4y 4 b a(y) .

By the definition of ba(y), if we take the infimum over F the left hand side of last expression will turn to be equal to zero.

Proof of Theorem 5. Let us define l(y) = lim sup

N →∞ y0→y

1

2N∆N,y0(AN, AN) . (1.4.11) By definition,ba(y) and l(y) are upper semicontinuous functions.

In order to conclude the proof, it is enough to verify the following inequality

l(y) ≤ 4y

4

ˆ

(37)

In fact, by the upper semicontinuity of ˆa(y) and the lower bound in (1.4.9) we obtain 4y4

ˆ a(y) ≤

4y4

lim supy0→yˆa(y0)

≤ lim sup y0→y 4y04 ˆ a(y0) ≤ lim sup y0→y lim sup N →∞ ∆N,y0(AN, AN) 2N ≤ l(y) ,

from which, we have the equality

l(y) = 4y

4

ˆ

a(y). (1.4.13)

From the definition of l(y), it is clear that lim sup

N →∞

1

2N∆N,y(AN, AN) ≤ l(y) ,

which together with (1.4.9) proves (1.4.10). Moreover, equality (1.4.13) together with the upper semicontinuity of l(y) gives the lower semicontinuity of ba(y) resulting in the continuity ofba(y), ending the proof.

Therefore, it remains to check the validity of inequality (1.4.12). This is equivalent to prove that for every θ ∈ R,

l(y) > θ =⇒ θ ≤ 4y

4

ˆ

a(y) . (1.4.14)

Suppose that l(y) > θ. Then, there exist a sequence yN → y such that

lim N →∞ 1 2N∆N,yN(AN, AN) = A > θ . By i) in Lemma 3 we have {∆N,yN(AN, AN)} 1/2 = sup h∈D(LN)\{0} hh, giN,y N DN,y(h) 1 2 .

(38)

1.4. CENTRAL LIMIT THEOREM VARIANCES AND DIFFUSION COEFFICIENT 28 and hwN, ANiN,yN > √ N θDN,yN(wN) 1 2 .

We can suppose without loss of generality that

hwNiN,yN = 0 . (1.4.15)

Taking βN = 2N1 DN,yN(wN) and vN = β

−1 2

N wN, we obtain a sequence of functions

{vN}N ≥1 such that lim inf N →∞ 1 2NhvN, ANiN,yN ≥ r θ 2 , and, DN,yN(vN) = 2N .

We can renormalize again by taking γN = y12 N

1

2NhvN, ANiN,yN and uN = γ

−1

N vN ,

obtain-ing a new sequence of functions {uN}N ≥1 satisfying

1 2N Z XN,−N(uN)pNp−NµN,yN(dp) = y 2 N , (1.4.16) and lim sup N →∞ 1 2N Z X −N ≤x≤x+1≤N a(px, px+1)[Xx,x+1(uN)]2µN,yN(dp) ≤ 4y4 θ . (1.4.17) The aim of Lemma 9, Lemma 11 and Lemma 13 given below, is to use (1.4.16) and (1.4.17) in order to obtain a sort of weak limit of {uN}N ≥1 which will lead us to a

function ξ satisfying i) Eνy[ξ] = 0 , ii) Eνy[p0p1ξ] = 0 , iii) Eνy[a(p0, p1){p0p1 + ξ} 2 ] ≤ 4y 4 θ , and an additional condition concerning Xi,i+1τjξ − Xj,j+1τiξ.

Condition iii ) obviously implies

θ ≤ 4y

4

Eνy[a(p0, p1){p0p1+ ξ}2]

(39)

Rather less obvious is the fact that i ), ii ) and the extra condition mentioned above implies that ξ belongs to the closure in L2(νy) of the set over which the infimum in the

definition of ba(y) is taken. The proof of this fact is the content Section 2.3.

The fact mentioned above, together with (1.4.18) imply the left hand side of (1.4.14), finishing the proof of the Theorem 5.

Now we state and prove the lemmas concerning the construction of the weak limit ζ endowed with the required properties.

Lemma 9. Given θ > 0, k ∈ N and a convergent sequence of positive real numbers {yN}N ≥1 satisfying (1.4.16) and (1.4.17), there exists a sequence of functions {u

(k) N }N ≥1

depending on the variables p−k, · · · , pk such that

1 2k Z Xk,−k(u (k) N )pkp−kµN,yN(dp) = y 2 N , (1.4.19) and, lim sup N →∞ 1 2k Z X −k≤x≤x+1≤k a(px, px+1)[Xx,x+1(u (k) N )] 2µ N,yN(dp) ≤ 4y4 θ , (1.4.20) where y is the limit of {yN}N ≥1.

Proof. Define for −N ≤ x ≤ x + 1 ≤ N

αNx,x+1 = yN−2EµN,yN[pxpx+1Xx,x+1(uN)] ,

and,

βx,x+1N = y−4N EµN,yN[a(px, px+1)[Xx,x+1(uN)]2] .

It follows from (1.4.16) and (1.4.17) that

αN−N,−N +1+ · · · + αNN −1,N = 2N , and that for every  > 0 there exist N0 such that

β−N,−N +1N + · · · + βN −1,NN ≤ 2N (4 θ + y

−4 N ) ,

(40)

1.4. CENTRAL LIMIT THEOREM VARIANCES AND DIFFUSION COEFFICIENT 30

By using Lemma 10 stated and proved below, we can conclude the existence of a block ΛN of size 2k contained in the discrete torus {−N, · · · , N } such that

γN X x∈ΛN αNx,x+1 !2 ≥ 2k X x∈ΛN βx,x+1N , (1.4.21) with γN = 4θ + yN−4.

Let us now introduce some notation. Denotes by RN be the rotation of axes defined

as

RN : R2N +1 → R2N +1

(p−N, · · · , pN) → (p−N +1, · · · , pN, p−N) .

For an integer i > 0 we denote by Ri

N the composition of RN with itself i times, for

i < 0 the inverse of R−iN by RiN , and R0N for the identity function. As usually, given a function u : R2N +1 → R we define Ri

Nu = u ◦ RiN.

Let us define

wN = RNi uN .

Now we proceed to check that (1.4.19) and (1.4.20) are satisfied by the sequence of functions {uk N}N ≥1 defined as ukN = 2k X x∈ΛN αNx,x+1 !−1 EµN,yN[wN | Λk] ,

where Λk= {p−k, · · · , pk}. Because of the invariance under axes rotation of the measure

µN,yN, together with the relation

Xx,x+1(wN) = R−iXx+i,x+i+1(uN) ,

which follows directly from the definition of wN, we have

EµN,yN[pxpx+1Xx,x+1(uk N)] = 2k X x∈ΛN αNx,x+1 !−1

EµN,yN[px+ipx+i+1Xx+i,x+i+1(uN)] ,

(41)

equal to 2k X x∈ΛN αNx,x+1 !−1 EµN,yN[X x∈ΛN pxpx+1Xx,x+1(uN)] , proving (1.4.19).

Using Jensen’s inequality, and an analogous argument as the one used in the preced-ing lines, we obtain that

EµN,yN[a(p x, px+1)[Xx,x+1(ukN)]2] is bounded above by 4k2 X x∈ΛN αNx,x+1 !−2

EµN,yN[a(px+i, px+i+1)[Xx+i,x+i+1(uN)]2] ,

which implies (1.4.20) after adding over x ∈ Λ , using relation (1.4.21) and taking the lim sup as N goes to infinity.

Now we state and proof a technical result used in the proof of the preceding lemma. Lemma 10. Let {ai}mi=1 and {bi}mi=1 two sequences of real and positive real numbers,

respectively, satisfying m X i=1 ai = m and m X i=1 bi ≤ mγ , (1.4.22)

for fixed constants m ∈ N, γ > 0 and k << m. Then, there exists a block Λ of size 2k contained in the discrete torus {1, · · · , m} such that

γ X i∈Λ ai !2 ≥ 2kX i∈Λ bi . (1.4.23)

Proof. It is enough to check the case where 2k is a factor of m. In fact, in the opposite case we can consider periodic sequences of size 2km instead of the originals ones {ai}mi=1

and {bi}mi=1.

Therefore we can suppose that m = 2kl for some integer l, and define for i ∈ {1, · · · , l} αi = X x∈Λi ax and βi = X x∈Λi bx ,

(42)

1.4. CENTRAL LIMIT THEOREM VARIANCES AND DIFFUSION COEFFICIENT 32

where Λi = {2k(i − 1), · · · , 2ki}.

We want to conclude that (1.4.23) is valid for at least one of the Λi’s. Let us argue

by contradiction.

Suppose that √2kβi > αiγ

1

2 for every i = 1, · · · , l. Adding over i and using the first

part of hypothesis (1.4.22), we obtain

l X i=1 p 2kβi > mγ 1 2 .

By squaring both sides of the last inequality we have,

l

X

i=1

βi > mγ ,

which is in contradiction with the second part of hypothesis (1.4.22).

Now we proceed to take, for each positive integer k, a weak limit of the sequence {u(k)N }N ≥1 obtained in Lemma 9.

Lemma 11. For each positive integer k there exists a function uek depending on the

variables p−k, · · · , pk such that

1 2kEνy[Xk,−k(uek)pkp−k] = y 2 , (1.4.24) and 1 2kEνy " X −k≤x≤x+1≤k a(px, px+1)[Xx,x+1(euk)] 2 # ≤ 4y 4 θ . (1.4.25)

Proof. Consider the linear functionals ΛNi,i+1 defined for −k ≤ i ≤ i + 1 ≤ k by ΛN i,i+1 : L2(R2k+1; νy) → R w → EµN,yN[X i,i+1(u (k) N )w] .

Let Pk be an enumerable dense set of polynomials in L2(R2k+1; νy). From (1.4.25) and

the Cauchy-Schwartz inequality we obtain the existence of a constant C such that

|ΛNi,i+1(w)| ≤ C Z w2dµN,yN 12 , for every w ∈ Pk.

(43)

By a diagonal argument we can draw a subsequence for which the limits of ΛNi,i+1(w) exist for all w ∈ Pk. Moreover, passing to the limit and extending to the whole space L2(R2k+1; νy), we get linear functionals Λi,i+1 satisfying

|Λi,i+1(w)| ≤ C

Z

w2dνy

12

. (1.4.26)

On the other hand, consider the linear functionals ΛN defined by ΛN : L2(R2k+1; νy) → R

w → EµN,yN[u(k)

N w] .

Because of (1.4.25), (1.4.15) and Poincare’s inequality we have

EµN,yN[(u(k)

N )

2] ≤ 4y4C

θ ,

for a constant C depending only on k. Then, by the very same arguments used above, we get a linear functional Λ satisfying

|Λ(w)| ≤√C Z

w2dνy

12

. (1.4.27)

Finally, it follows from (1.4.26) and (1.4.27) the existence of a function euk satisfying

Λi,i+1(w) = Eνy[Xi,i+1(euk)w] , Λ(w) = Eνy[uekw] , and therefore, satisfying also (1.4.24) and (1.4.25).

Now using the sequence {euk}k∈N we construct a sequence of functions {uk0}k0∈M

indexed on an infinite subset of N, each one depending on the variables p−k0, · · · , pk0.

This sequence will satisfy, besides (1.4.24) and (1.4.25), an additional condition regarding the contribution of the terms near the boundary of {−k0, · · · , k0} to the total Dirichlet form.

(44)

1.4. CENTRAL LIMIT THEOREM VARIANCES AND DIFFUSION COEFFICIENT 34

of N, each one depending on the variables p−k0, · · · , pk0, satisfying (1.4.24), (1.4.25) and

Eνy[(Xx,x+1(uk0))

2] = O(k7/8) ,

for {x, x + 1} ⊂ Ik0∪ Jk0, where Ik0 = [−k0, −k0+ (k0)1/8] and Jk0 = [k0 − (k0)1/8, k0].

Proof. Given k > 0 divide each interval [−k, −k + k1/4] and [k − k1/4, k] into k1/8 blocks of size k1/8, and consider the sequence {euk}k∈N obtained in Lemma 11.

Because of (1.4.25), for every k > 0 there exist k0 ∈ [k − k1/8, k] such that

Eνy   X {x,x+1}∈Ik0∪Jk0 (Xx,x+1(euk)) 2  = O(k 7/8 ) .

Define for each k > 0 the function uk0 = 1

Cy,k,k0Eνy  e uk| Fk 0 −k0 , where Cy,k,k0 = 1 2k0y2    2ky2− Eνy   X {x,x+1}∈Ik0∪Jk0 pxpx+1Xx,x+1(uek)      and Fk0

−k0 denotes the σ-field generated by p−k0, · · · , pk0.

Is easy to see that the sequence {uk0}k0 satisfies the desired conditions.

Finally, we obtain the weak limit used in the proof of Theorem 5. Lemma 13. There exist a function ξ in L2

y) satisfying Eνy[ξ] = 0, (1.4.28) Eνy[p0p1ξ] = 0, (1.4.29) Eνy[a(p0, p1)[p0p1+ ξ] 2] ≤ 4y4 θ , (1.4.30)

and the integrability conditions

Xi,i+1(τjξ) = Xj,j+1(τiξ) if {i, i + 1} ∩ {j, j + 1} = ∅, (1.4.31)

(45)

Proof. For all integer k > 0 define ζk = 1 2k0 k0−1 X i=−k0 Xi,i+1(uk0)(τ−i).

It is clear from the definition of ζk that

Eνy[ζk] = 0, and, p0p1ζk0(ω) = 1 2k0 k0−1 X i=−k0

{pipi+1Xi,i+1(uk0)}(τ−i).

Then, from (1.4.24) it follows

Eνy[p0p1ζk] = y

4

. On the other hand, from (1.4.25) we have

Eνy[a(p0, p1)ζ 2 k] = Eνy  a(p0, p1) 1 2k0 k0−1 X i=−k0 [Xi,i+1(uk0)](T−iω) !2  ≤ Eνy " a(p0, p1) 1 2k0 k0−1 X i=−k0 [Xi,i+1(uk0)]2(T−iω) # = 1 2kEνy "k0−1 X i=−k0

a(pi, pi+1)[Xi,i+1(uk0)]2

#

≤ 4y

4

θ .

Now consider the sequence {ξk}k≥1 defined by

ξk = ζk− p0p1.

Since the preceding sequence is uniformly bounded in L2(νy), there exist a weak limit

function ξ ∈ L2

y). Obviously, the function ξ satisfies (1.4.28),(1.4.28) and (1.4.30).

In addition, an elementary calculation shows that 1.4.31 and 1.4.32 are satisfied by ξk up to an error coming from a finite number of terms near the edge of [−k0, k0]. Then,

(46)

1.5. TIGHTNESS 36

1.5

Tightness

Let us firstly introduce some notation in order to define a space in which fluctuations take place and in which we will be able to prove tightness. Let −∆ be the positive operator, essentially self-adjoint on L2([0, 1]) defined by

Dom(−∆) = C02([0, 1]), −∆ = − d

2

dx2 ,

where C02([0, 1]) denotes the space of twice continuously differentiable functions on (0, 1) which are continuous in [0, 1] and which vanish at the boundary. It is well known that its normalized eigenfunctions are given by en(x) =

2 sin(πnx) with corresponding eigenvalues λn= (πn)2 for every n ∈ N , moreover, {en}n∈N forms an orthonormal basis

of L2([0, 1]).

For any nonnegative integer k denote by Hk the Hilbert space obtained as the

com-pletion of C2

0([0, 1]) endowed with the inner product

hf, gik = hf, (−∆)kgi ,

where h, i stands for the inner product in L2([0, 1]). We have from the spectral theorem

for self-adjoint operators that

Hk= {f ∈ L2([0, 1]) : ∞ X n=1 n2khf, eni < ∞} , (1.5.1) and, hf, g)k = ∞ X n=1 (πn)2khf, enihg, eni . (1.5.2)

This is valid also for negative k. Namely, if we denote the topological dual of Hk by

H−k we have H−k = {f ∈ D0([0, 1]) : ∞ X n=1 n−2kf (en)2 < ∞} . (1.5.3)

The H−k-inner product between the distributions f and g can be written as

hf, gi−k = ∞

X

n=1

Références

Documents relatifs

We extend in this work previous results of regularization for the spatially homogeneous non cutoff 2D radially symmetric Boltzmann equation which held only in the case of

Bound-bound transitions (line transitions) between the lower energy level i and the upper energy level j may occur as radiative excitation, spon- taneous radiative

L’accès aux archives de la revue « Annales de la faculté des sciences de Toulouse » ( http://picard.ups-tlse.fr/~annales/ ) implique l’accord avec les conditions

2014 The Brauer-Manin obstruction for homogeneous spaces with connected or abelian stabilizer, J. für die reine und

Though the corrector problem may not have an exact solution due to non-coercivity of the Hamiltonian, homogenization suffices with an approximate corrector whose existence is

Recall that the projection onto Ker(∆ E ℓ ) is transversely smooth: for the (untwisted) leafwise signature operator; whenever E is a bundle associated to the normal bundle of

Using the weak solution of Inverse mean curvature flow, we prove the sharp Minkowskitype inequality for outward minimizing hypersurfaces in Schwarzschild space... where Ric is the

Changsha, Hunan 410081, People’s Republic of China jiaolongche n@sina.com Hunan Normal University, Department of Mathematics,.. Changsha, Hunan 410081, People’s Republic