• Aucun résultat trouvé

Propagation of chaos and Poincaré inequalities for a system of particles interacting through their cdf

N/A
N/A
Protected

Academic year: 2021

Partager "Propagation of chaos and Poincaré inequalities for a system of particles interacting through their cdf"

Copied!
33
0
0

Texte intégral

(1)

HAL Id: hal-00127988

https://hal.archives-ouvertes.fr/hal-00127988v2

Submitted on 14 Nov 2008

HAL is a multi-disciplinary open access archive for the deposit and dissemination of sci- entific research documents, whether they are pub- lished or not. The documents may come from teaching and research institutions in France or abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires publics ou privés.

Propagation of chaos and Poincaré inequalities for a system of particles interacting through their cdf

Benjamin Jourdain, Florent Malrieu

To cite this version:

Benjamin Jourdain, Florent Malrieu. Propagation of chaos and Poincaré inequalities for a system of particles interacting through their cdf. Annals of Applied Probability, Institute of Mathematical Statistics (IMS), 2008, 18 (5), pp.1706-1736. �10.1214/07-AAP513�. �hal-00127988v2�

(2)

2008, Vol. 18, No. 5, 1706–1736 DOI:10.1214/07-AAP513

c

Institute of Mathematical Statistics, 2008

PROPAGATION OF CHAOS AND POINCAR´E INEQUALITIES FOR A SYSTEM OF PARTICLES INTERACTING

THROUGH THEIR CDF

By Benjamin Jourdain and Florent Malrieu Ecole des Ponts and Universit´e Rennes 1´

In this paper, in the particular case of a concave flux function, we are interested in the long time behavior of the nonlinear process as- sociated in [Methodol. Comput. Appl. Probab.2(2000) 69–91] to the one-dimensional viscous scalar conservation law. We also consider the particle system obtained by replacing the cumulative distribu- tion function in the drift coefficient of this nonlinear process by the empirical cumulative distribution function. We first obtain a trajec- torial propagation of chaos estimate which strengthens the weak con- vergence result obtained in [8] without any convexity assumption on the flux function. Then Poincar´e inequalities are used to get explicit estimates concerning the long time behavior of both the nonlinear process and the particle system.

Introduction. In this paper, we are interested in the viscous scalar con- servation law withC1 flux function−A

tFt(x) =σ2

2 ∂xxFt(x) +∂x(A(Ft(x)), F0(x) =H∗m(x), (1)

wheremis a probability measure on the real line andH(x) = 1{x0} denotes the Heaviside function. As a consequence,H∗mis the cumulative distribu- tion function of the probability measure m. Since A appears in this equa- tion through its derivative, we suppose without restriction that A(0) = 0.

According to [8], one may associate the following nonlinear process with the conservation law:

Xt=X0+σBtZ t

0 A(H∗Ps(Xs))ds,

∀t≥0,the law of Xt is Pt, (2)

Received January 2007; revised December 2007.

AMS 2000 subject classifications.65C35, 60K35, 60E15, 35K15, 46N30.

Key words and phrases.Viscous scalar conservation law, nonlinear process, particle sys- tem, propagation of chaos, Poincar´e inequality, long time behavior.

This is an electronic reprint of the original article published by the Institute of Mathematical StatisticsinThe Annals of Applied Probability, 2008, Vol. 18, No. 5, 1706–1736. This reprint differs from the original in pagination and typographic detail.

1

(3)

where (Bt)t0is a real Brownian motion independent from the initial random variable X0 with law m and σ a positive constant. The process X is said to be nonlinear in the sense that the drift term of the SDE depends on the entire lawPtofXt. More precisely, according to [8], this nonlinear stochastic differential equation admits a unique weak solution. Moreover, H∗Pt(x) is the unique bounded weak solution of (1). Fort >0, by the Girsanov theorem, Ptadmits a densityptwith respect to the Lebesgue measure on the real line.

We want to address the long time behavior of the nonlinear process solv- ing (2) by studying convergence of the density pt (see [2] and [3] for a simi- lar study in a different setting). Since the cumulative distribution function x→H∗Ps(x) which appears in the drift coefficient is nondecreasing, con- vexity ofA is a natural assumption in order to ensure ergodicity. Then the flux function−A in the conservation law (1) is concave.

In the first section of the paper, after recalling results obtained in [8], we show that trajectorial uniqueness holds for (2) under convexity of A.

Then we introduce a simulable system of nparticles obtained by replacing in the drift coefficient the cumulative distribution function by its empirical version and the derivative A by a suitable finite difference approximation.

WhenAis convex, existence and trajectorial uniqueness hold for this system.

Moreover, we prove a trajectorial estimation of propagation of chaos which strengthens the weak convergence result obtained in [8]. Unfortunately, be- cause the empirical cumulative distribution function is a step function and therefore not an increasing one, this estimation is not uniform in time.

The second and main section deals with the long time behavior of both the nonlinear process and the particle system. We address the convergence of the density pt of Xt by first studying the convergence of the associated solution H∗pt of (1) to the solution F with the same expectation of the stationary equation σ22xxF(x) +∂x(A(F(x)) = 0 obtained by removing the time derivative in (1). For this result, no convexity hypothesis is made on A. Instead, one assumesA(u)<0 foru∈(0,1), A(0)<0,A(1) = 0 and A(1)>0. In contrast, to prove exponential convergence of the density of the particle system uniform in the numbern of particles, we suppose that the function A is uniformly convex. This hypothesis ensures the existence of an invariant distribution for the particle system. In [14], a necessary and sufficient condition on the drift sequence is established for existence of the invariant measure and convergence in total variation norm for the law of the particle system at timet to this measure. In the present paper, the key step to derive quantitative convergence to equilibrium consists in obtaining a Poincar´e inequality for the stationary density of the particle system uniform inn. This density has exponential-like tails and therefore does not satisfy a logarithmic Sobolev inequality. So the derivation of the Poincar´e inequality cannot rely on the curvature criterion, used, for instance, in [5,6,12] or [13]

(4)

for the granular media equation. Instead we make a direct estimation of the Poincar´e constant using the specific analytic form of the invariant density.

To our knowledge, our study provides the first example of a particle system, for which a Poincar´e inequality but no logarithmic Sobolev inequality holds uniformly in the numbern of particles.

Assumption. Throughout the paper, we assume thatAis aC1function on [0,1] s.t. A(0) = 0.

1. Propagation of chaos.

1.1. The nonlinear process. Let us first state existence and uniqueness for the nonlinear stochastic differential equation (2).

Theorem 1.1. The nonlinear stochastic differential equation (2) admits a unique weak solution ((Xt, Pt))t0. For t >0, Pt admits a density pt with respect to the Lebesgue measure on R. The function (t, x)7→H∗Pt(x) is the unique bounded weak solution of the viscous scalar conservation law (1).

Moreover,

∀t≥0 Xt−X0 is integrable andE(Xt−X0) =−A(1)t.

(3)

Last, if the function A is convex on [0,1], (2) admits a unique strong solu- tion.

Proof. The first and third statements are consequences of Proposition 1.2 and Theorem 2.1 of [8] [uniqueness follows from uniqueness for (1) and existence is obtained by a propagation of chaos result].

According to the Yamada–Watanabe theorem, to deduce the last state- ment, it is enough to check that whenAis convex, then trajectorial unique- ness holds for the standard stochastic differential equation

dXt=σ dBt−A(H∗Qt(Xt))dt

where (Qt)t0 is the flow of time-marginals of a probability measureQ on C([0,+∞),R). Since for eacht≥0 the functionx7→A(H∗Qt(x)) is nonde- creasing, if (Xt)t0and (Yt)t0 both solve this standard SDE, then|Xt−Yt| is bounded by

|X0−Y0|+ Z t

0 sign(Xs−Ys)(A(H∗Qs(Ys))−A(H∗Qs(Xs)))ds, and then by|X0−Y0|which concludes the proof of trajectorial uniqueness.

Existence of the densitypt for t >0 follows from the boundedness of the drift coefficient and the Girsanov theorem. To prove (3), one first remarks

(5)

that by boundedness of the drift coefficient, for each t≥0, the random variableXt−X0 is integrable and

E(Xt−X0) =− Z t

0

E(A(H∗Ps(Xs)))ds

=− Z t

0

Z

R

A Z x

−∞Ps(dy)

Ps(dx)ds.

Fors >0, since by the Girsanov theorem Ps does not weight points, Z

R

A Z x

−∞Ps(dy)

Ps(dx) = [A(H∗Ps(x))]+−∞=A(1).

Corollary 1.2. Assume thatA is C2 on [0,1]. Then the function H∗ Pt(x) is C1,2 on (0,+∞)×R and solves (1) in the classical sense on this domain.

Proof. By the Girsanov theorem, fort0>0, the lawPt0 ofXt0 admits a density with respect to the Lebesgue measure onR. Hence (t, x)7→H∗Pt(x) is a continuous function on (0,+∞)×R with values in [0,1]. According to [11], Theorem 8.1, page 495, Remark 8.1, page 495 and Theorem 2.5, page 18, there exists a functionuwith values in [0,1], continuous on [0,+∞)×R and C1,2 on (0,+∞)×Rsuch that

∀x∈R, u(0, x) =H∗Pt0(x),

∀(t, x)∈(0,+∞)×R, ∂tu(t, x) = σ2

2 ∂xxu(t, x) +∂x(A(u(t, x))).

By the uniqueness result for bounded weak solutions of this viscous scalar conservation law recalled in Theorem 1.1, ∀t≥t0, H∗Pt(x) =u(t−t0, x).

The conclusion follows sincet0 is arbitrary.

1.2. Study of the particle system. For n∈N, let (an(i))1in be a se- quence of real numbers. In this section, we are interested in then-dimensional stochastic differential equation

dXti,n=σ dBit−an

n

X

j=1

1{Xtj,nXti,n}

!

dt, X0i,n=X0i, 1≤i≤n, (4)

where (Bi)i1 are independent standard Brownian motions independent from the sequence (X0i)i1 of initial random variables.

In the next section devoted to the approximation of the nonlinear stochas- tic differential equation (2), we will choosean(i) equal to the finite difference approximationn(A(i/n)−A((i−1)/n)) ofA(ni). For this particular choice, the nondecreasing assumption made in the following proposition is implied by convexity ofA.

(6)

Proposition 1.3. Assume that the sequence (an(i))1in is nonde- creasing. Then the stochastic differential equation (4) has a unique strong so- lution. Let(Yt1,n, . . . , Ytn,n)denote another solution starting from(Y01, . . . , Y0n) and driven by the same Brownian motion (B1, . . . , Bn). Then

a.s.,∀t≥0

n

X

i=1

(Xti,n−Yti,n)2

n

X

i=1

(X0i−Y0i)2. (5)

In addition, if the initial conditions (X01, . . . , X0n) and (Y01, . . . , Y0n) are s.t.

a.s., ∀i∈ {1, . . . , n}, X0i< Y0i (resp. X0i ≤Y0i), then

a.s., ∀t≥0,∀i∈ {1, . . . , n} Xti,n< Yti,n (resp. Xti,n≤Yti,n).

(6)

Existence of a weak solution to (4) is a consequence of the Girsanov the- orem. Therefore, according to the Yamada–Watanabe theorem, it is enough to prove (5) which implies trajectorial uniqueness to obtain existence of a unique strong solution. To do so, we will need the following lemma.

Lemma 1.4. Let (a(i))1in and (b(i))1in denote two nondecreasing sequences of real numbers. Then for any permutation τ ∈ Sn,

n

X

i=1

a(i)b(τ(i))≤

n

X

i=1

a(i)b(i).

(7)

Proof. For n= 2, the result is an easy consequence of the inequality (a(2)−a(1))(b(2)−b(1))≥0.

For n >2, we define τ1 asτ if τ(1) = 1 and as τ composed with the trans- position between 1 andτ1(1) otherwise. This way, τ1(1) = 1. In addition, using the result forn= 2, we get Pni=1a(i)b(τ(i))≤Pni=1a(i)b(τ1(i)).

For 2≤j≤n−1, we define inductively τj as τj1 if τj1(j) =j and as τj1composed with the transposition betweenjandτj11(j) otherwise. This way, for 1≤i≤j,τj(i) =i. Again by the result forn= 2, one has

n

X

i=1

a(i)b(τ(i))≤

n

X

i=1

a(i)b(τ1(i))≤

n

X

i=1

a(i)b(τ2(i))≤ · · · ≤

n

X

i=1

a(i)b(τn1(i)).

We conclude by remarking thatτn1 is the identity.

We are now ready to complete the proof of Proposition1.3.

Proof of Proposition1.3. Let (X1,n, . . . , Xn,n) and (Y1,n, . . . , Yn,n) denote two solutions. The difference

n

X

i=1

(Xti,n−Yti,n)2

n

X

i=1

(X0i−Y0i)2

(7)

is equal to 2

Z t

0 n

X

i=1

(Xsi,n−Ysi,n) an

n

X

j=1

1{Ysj,nYsi,n}

!

−an

n

X

j=1

1{Xsj,nXsi,n}

!!

ds.

(8)

By the Girsanov theorem, for anys >0 the distributions of (Xs1,n, . . . , Xsn,n) and (Ys1,n, . . . , Ysn,n) admit densities w.r.t. the Lebesgue measure onRnand thereforedP⊗ds a.e. the positionsXs1,n, . . . , Xsn,n (resp. Ys1,n, . . . , Ysn,n) are distinct and there is a unique permutation τsX ∈ Sn (resp. τsY ∈ Sn) such thatXsτsX(1),n< XsτsX(2),n<· · ·< XsτsX(n),n (resp. YsτsY(1),n< YsτsY(2),n<· · ·<

YsτsY(n),n). Therefore dP⊗ds a.e.,

n

X

i=1

(Xsi,n−Ysi,n) an

n

X

j=1

1{Ysj,nYsi,n}

!

−an

n

X

j=1

1{Xsj,nXsi,n}

!!

is equal to

n

X

i=1

an(i)((XsτsY(i),n−YsτsY(i),n)−(XsτsX(i),n−YsτsX(i),n)).

The sequence (an(i))1inis nondecreasing. Applying Lemma1.4withb(i) = XsτsX(i),n andτ = (τsX)1◦τsY then withb(i) =YsτsY(i),n andτ= (τsY)1◦τsX, one obtains that the integrand in (8) is nonpositive dP⊗ds a.e. Hence (5) holds.

Let us now suppose that a.s. ∀i∈ {1, . . . , n}, X0i < Y0i and define ν = inf{t >0 :∃i∈ {1, . . . , n}, Xti,n≥Yti,n}with the convention inf∅= +∞. From now on, we restrict ourselves to the event {ν <+∞}. Let i∈ {1, . . . , n} be such that Yνi,n = Xνi,n. There is an increasing sequence (sk)k1 of positive times with limit ν such that ∀k ≥ 1, an(Pnj=11{Xj,n

skXski,n}) <

an(Pnj=11{Yj,n

sk Yski,n}). Since (an(i))1in is nondecreasing, by extracting a subsequence still denoted by (sk)k for simplicity, one deduces the existence of j∈ {1, . . . , n} with j6=i such that ∀k≥1, Xsi,nk < Xsj,nk and Ysj,nk ≤Ysi,nk . Since sk < ν, Xsi,nk < Xsj,nk < Ysj,nk ≤Ysi,nk . By continuity of the paths, one obtainsXνi,n=Xνj,n=Yνj,n=Yνi,n. Now since the probability of the event

∃i1, i2, i3 dist. in {1, . . . , n},∃t >0 X0i1+σBti1=X0i2+σBti2=X0i3+σBit3 is equal to 0, the Girsanov theorem implies that a.s.∀l∈ {1, . . . , n} \ {i, j}, Xνl,n 6=Xνi,n=Xνj,n. In the same way, Yνl,n6=Yνi,n=Yνj,n. By continuity of the paths and definition of ν one deduces that for k large enough, and for everyt∈[sk, ν],

n

X

l=1 l6=i,j

1{Ytl,nYti,n}

n

X

l=1 l6=i,j

1{Xtl,nXti,n};

n

X

l=1 l6=i,j

1{Ytl,nYtj,n}

n

X

l=1 l6=i,j

1{Xtl,nXtj,n}.

(8)

Since a.s.dt a.e.,Yti,n6=Ytj,nand (an(i))1inis nondecreasing, one obtains that a.s.dt a.e. on [sk, ν],

an

n

X

l=1

1{Yl,n t Yti,n}

! +an

n

X

l=1

1{Yl,n t Ytj,n}

!

≤an

n

X

l=1

1{Xtl,nXtj,n}

! +an

n

X

l=1

1{Xtl,nXti,n}

! .

By integration with respect toton [sk, ν], this implies that a.s.Yνi,n−Xνi,n+ Yνj,n−Xνj,n≥Ysi,nk −Xsi,nk +Ysj,nk −Xsj,nk >0. Therefore P(ν <+∞) = 0.

When a.s. for i∈ {1, . . . , n}, X0i ≤Y0i, one obtains that for ε >0 the solution (Yt1,n,ε, . . . , Ytn,n,ε) to (4) starting from (Y01+ε, . . . , Y0n+ε) is such that

a.s., ∀t≥0 ∀i∈ {1, . . . , n} Xti,n< Yti,n,ε. Since by (5), Yti,n,ε≤Yti,n+√

nε, one easily concludes by letting ε→0.

1.3. Trajectorial propagation of chaos. From now on, we set

∀n∈N,∀i∈ {1, . . . , n} an(i) =n

A i

n

−A i−1

n (9)

and assume that the initial positions (X0i)i1of the particles are independent and identically distributed according tom. We prefer to define an(i) with the above finite difference approximation of the choice A(i/n) made in [8]

because the sumPni=1an(i) which plays a role in the long time behavior of the particle system is then simply equal to nA(1). One could also obtain trajectorial propagation of chaos estimates similar to Theorem1.5below for the choice an(i) =A(i/n).

In the present section, we also suppose that A is a convex function on [0,1]. By Theorem 1.1, for each i≥1, the nonlinear stochastic differential equation

Xti=X0i +σBtiZ t

0

A(H∗Ps(Xsi))ds,

∀t≥0,the law ofXti is Pt, (10)

has a unique solution and for allt≥0, the law Ptof Xti does not depend on i. Under a Lipschitz regularity assumption on A, we obtain the following trajectorial propagation of chaos estimation.

Theorem 1.5. If A: [0,1]→Ris convex and A is Lipschitz continuous with constant K, then

∀n≥1,∀1≤i≤n,∀t≥0 E

sup

s[0,t]

(Xsi,n−Xsi)2

≤K2t2 6n .

(9)

Proof. Let us write Pni=1(Xti,n−Xti)2 as 2

Z t

0 n

X

i=1

(Xsi,n−Xsi) an

n

X

j=1

1{XsjXsi}

!

−an

n

X

j=1

1{Xsj,nXsi,n}

!!

ds

+ 2 Z t

0 n

X

i=1

(Xsi,n−Xsi)C(s, Xs1, . . . , Xsn)ds whereC(s, Xs1, . . . , Xsn) is equal to

A(H∗Ps(Xsi))−n A 1 n

n

X

j=1

1{Xj

sXis}

!

−A 1 n

n

X

j=1

1{Xj

sXsi}− 1 n

!!

. Like in the proof of trajectorial uniqueness for (4), because of the convex- ity of A, the first term of the r.h.s. is nonpositive. Moreover, by Lipschitz continuity ofA,

A(H∗Ps(Xsi))−n A 1 n

n

X

j=1

1{XjsXsi}

!

−A 1 n

n

X

j=1

1{XsjXsi}− 1 n

!!!2

= Z 1

0

A(H∗Ps(Xsi))−A 1 n

n

X

j=1

1{Xj

sXis}+θ−1 n

!

!2

≤K2 n2

Z 1 0

X

j6=i

H∗Ps(Xsi)−1

{XsjXsi}

!

+ (H∗Ps(Xsi)−θ)

!2

dθ.

Fors >0, as the variablesXsi are i.i.d. with common law Ps which does not weight points and H∗Ps(Xsi) is uniformly distributed on [0,1],

Z 1 0

E

X

j6=i

(H∗Ps(Xsi)−1{Xj

sXis}) + (H∗Ps(Xsi)−θ)

!2!

=X

j6=i

E((H∗Ps(Xsi)−1

{XsjXsi})2) + Z 1

0

E((H∗Ps(Xsi)−θ)2)dθ

= (n−1)E((H∗Ps(Xsi))(1−H∗Ps(Xsi))) + 1/6

=n/6.

Using the Cauchy–Schwarz inequality, one obtains E sup

s[0,t]

n

X

i=1

(Xsi,n−Xsi)2

!

≤2 Z t

0

v u u tK2

6nE n

X

i=1

(Xsi,n−Xsi)

!2! ds

≤2K

√6 Z t

0

v u u

tE sup

u[0,s]

n

X

i=1

(Xui,n−Xui)2

! ds.

(10)

By comparison with the ordinary differential equationα(t) = 2Kqα(t)6 , one concludes that

∀t≥0 E sup

s[0,t]

n

X

i=1

(Xsi,n−Xsi)2

!

≤K2t2 6 .

Exchangeability of the couples ((Xi,n, Xi))i∈{1,...,n} completes the proof.

Remark 1.6. One could think that assuming thatA is uniformly con- vex:

∃α >0,∀0≤x≤y≤1 A(y)−A(x)≥α(y−x) (11)

would lead to a better estimation. Indeed, then for everyi∈ {1, . . . , n−1}, an(i+ 1)−an(i) =n

Z (i+1)/n i/n

A(x)−A

x−1 n

dx≥α n. But since even in this situation, the nonpositive term

n

X

i=1

(Xsi,n−Xsi) an

n

X

j=1

1{Xj sXsi}

!

−an

n

X

j=1

1{Xj,n s Xsi,n}

!!

vanishes as soon as the order between the coordinates of (Xs1,n, . . . , Xsn,n) is the same as the order between the coordinates of (Xs1, . . . , Xsn), we were not able so far to improve the estimation.

Corollary 1.7. Under the hypotheses of Theorem 1.5, let m˜ be a probability measure on R such that ∀x∈R, H∗m(x)˜ ≤H∗m(x). If for some random variableU1 uniform on[0,1]independent from(Bi)i1, X01= inf{x:H∗m(x)≥U1} and (Yt1)t0 denotes the solution of the nonlinear stochastic differential equation

Yt1=Y01+σBt1Z t

0

A(H∗P˜s(Ys1))ds,

∀t≥0,the law ofYt1 is P˜t, (12)

withY01= inf{x:H∗m(x)˜ ≥U1}, then

P(∀t≥0, Xt1≤Yt1) = 1.

Moreover ∀t≥0, ∀x∈R, H∗P˜t(x)≤H ∗Pt(x). Last, the function t7→

E|Yt1−Xt1| is constant.

Remark 1.8. At least when m and ˜m do not weight points, one has a.s. A(H∗P0(X01)) =A(H∗P˜0(Y01)) since H∗m(X01) =H∗m(Y˜ 01) =U1. Therefore a.s. d(Y1 −X1)0 = 0 and one may wonder whether a.s. Yt1

(11)

Xt1 does not depend on t. If this property holds, necessarily, a.s. dt a.e.

A(H∗Pt(Xt1)) =A(H∗P˜t(Yt1)). If A is increasing, a.s. for all t >0, H∗ pt(Xt1) =H∗p˜t(Yt1) with pt and ˜pt denoting the respective densities of Pt and ˜Pt. IfA isC2, the Brownian contribution ind(H∗pt(Xt1)−H∗p˜t(Yt1)) given by Itˆo’s formula vanishes, that is, pt(Xt1) = ˜pt(Yt1) and ∀u∈]0,1[, pt((H∗pt)1(u)) = ˜pt((H∗p˜t)1(u)) or equivalently ((H∗pt)1)(u) = ((H∗

˜

pt)1)(u). HenceYt1=Xt1+cfor a deterministic constantcwhich does not depend ont according to (3). Letting t→0, one obtains Y01=X01+c. This necessary condition turns out to be sufficient as (Xt1+c)t0 obviously solves the nonlinear stochastic differential equation (2) starting fromX01+c.

Proof of Corollary1.7. For (Ui)i2 a sequence of independent uni- form random variables independent from (U1,(Bi)i1), we set

∀i≥2 X0i= inf{x:H∗m(x)≥Ui} and Y0i= inf{x:H∗m(x)˜ ≥Ui}. SinceH∗m˜ ≤H∗m, a.s.∀i≥1,Y0i≥X0i. From Proposition1.3, one deduces that the solutions (Xt1,n, . . . , Xtn,n) and (Yt1,n, . . . , Ytn,n) to (4) respectively starting from (X01, . . . , X0n) and (Y01, . . . , Y0n) are such that

a.s., ∀n≥1,∀i∈ {1, . . . , n},∀t≥0 Yti,n≥Xti,n.

Since, by Theorem1.5, for fixedt≥0, one may extract from (Xt1,n, Yt1,n)n1 a subsequence almost surely converging to (Xt1, Yt1), one easily deduces that P(∀t≥0, Xt1≤Yt1) = 1. Hence

∀t≥0,∀x∈R H∗P˜t(x) =P(Yt1≤x)≤P(Xt1≤x) =H∗Pt(x).

Since|Yt1−Xt1| − |Y01−X01|=Yt1−Y01−(Xt1−X01), (3) ensures thatE|Yt1− Xt1| ∈[0,+∞] does not depend ont.

2. Long time behavior. In this section we are interested in the long time behavior of both the nonlinear process and the particle system. According to (3) and the equality Pni=1an(i) =nA(1) which follows from (9), we have to suppose A(1) = 0 in order to obtain convergence of the densities as t tends to infinity. We address the convergence of the densityptofXtby first studying the convergence of the associated cumulative distribution function Ft under the following hypothesis denoted by (H) in the sequel:

A(0) =A(1) = 0, A(0)<0, (H)

A(1)>0 and ∀u∈(0,1) A(u)<0.

These assumptions determine the spatial behavior at infinity of the drift coefficient in (2).

To prove exponential convergence of the density of the particle system uniform in the numbern of particles, we make the stronger assumption of

(12)

uniform convexity onA. The key step in the proof is to obtain a Poincar´e in- equality uniform inn for the stationary density of the particle system. This density has exponential-like tails and therefore does not satisfy a logarith- mic Sobolev inequality. So the derivation of the Poincar´e inequality cannot rely on the curvature criterion, used, for instance, by Malrieu [12,13] when dealing with the granular media equation. Instead, we take advantage of the following nice feature: up to reordering of the coordinates, the stationary density is the density of the image by a linear transformation of a vector of independent exponential variables. And it turns out that the control of the constant in the n-dimensional Poincar´e inequality relies on the Hardy inequality stated in Lemma 2.18 which is a one-dimensional Poincar´e-like inequality. To our knowledge, our study provides the first example of a par- ticle system, for which a Poincar´e inequality but no logarithmic Sobolev inequality holds uniformly in the numbern of particles.

2.1. The nonlinear process. In this section, we are first going to obtain necessary and sufficient conditions on the function A ensuring existence for the stationary Fokker–Planck equation obtained by removing the time- derivative in the nonlinear Fokker–Planck equation

tpt2

2 ∂xxpt+∂x(A(H∗pt)pt) (13)

satisfied by the density of the solution of (2). Under a slightly stronger condition, the solutions satisfy a Poincar´e inequality.

Lemma 2.1. A necessary and sufficient condition for the existence of a probability measure µ solving the stationary Fokker–Planck equation

σ2

2 ∂xxµ+∂x(A(H∗µ(x))µ) = 0

in the distribution sense is A(1) = 0 and A(u)<0 for all u∈(0,1). Under that condition, all the solutions are the translations of a probability measure with a C1 density f which satisfies

∀x∈R f(x) =− 2

σ2A(H∗f(x)) and (14)

f(x) =− 2

σ2A(H∗f(x))f(x).

If A(0)<0 and A(1)>0, then

f(x)∼

−2A(0) σ2

Z x

−∞

f(y)dy, when x→ −∞, 2A(1)

σ2

Z + x

f(y)dy, when x→+∞,

(13)

(15)

Z x

0

dy f(y) ∼

−σ2

2A(0)f(x), when x→ −∞, σ2

2A(1)f(x), when x→+∞,

and all the solutions satisfy a Poincar´e inequality and have a finite expecta- tion. Last, if the function A is C2 on [0,1], thenf is C2 and satisfies

f′′(x) =− 2

σ2A′′(H∗f(x))f2(x) +f2(x) f(x) . (16)

Proof. Let µ be a probability measure on R solving the stationary Fokker–Planck equation. The equality σ22xxµ=−∂x(A(H∗µ(x))µ) ensures that µ does not weight points. Hence the stationary equation is equivalent to ∂xx(σ22µ+A(H∗µ(x))) = 0. One deduces that µpossesses a C1 density f such that

∀x∈R f(x) =− 2

σ2A(H∗f(x)) +αx+β, (17)

for some constants α and β. Since A(0) = 0, letting x→ −∞ then x→ +∞ in the last equality, one obtains α=β=A(1) = 0. For u∈(0,1), since u=H∗f(x) for some x∈R and H∗f is not constant and equal to u, the Cauchy–Lipschitz theorem and (17) imply that A(u)6= 0. Since f is nonnegative, A(u)<0. Hence A(1) = 0 and A(u)<0 for all u∈(0,1) is a necessary condition.

Under that condition, a probability measureµsolves the stationary Fokker–

Planck equation if and only if its cumulative distribution functionH∗µ(x) is aC2 solution to the differential equation

ϕ(x) =− 2

σ2A(ϕ(x)), x∈R. (18)

By the Cauchy–Lipschitz theorem, for each v∈[0,1] this equation admits a unique solution ϕv defined onR with values in [0,1] such thatϕv(0) =v.

Moreover, asA(0) =A(1) = 0, ϕ0≡0 andϕ1≡1 and

∀v∈(0,1),∀x∈R 0< ϕv(x)<1.

(19)

For v∈(0,1), since ϕv is nondecreasing andϕv(x) =v−σ22R0xA(ϕv(y))dy, necessarily limy+ϕv(y) = 1. In the same way, limy→−∞ϕv(y) = 0 and ϕv is an increasing function fromRto (0,1) with inverse denoted by ϕv1. The uniqueness result for (18) implies that ∀v∈(0,1),∀x∈R, ϕv(x) =ϕ1/2(x+ ϕ1/21(v)). Therefore the solutions to the stationary Fokker–Planck equation

(14)

are the probability measures obtained by spatial translation of the proba- bility measure with densityf(x) =ϕ1/2(x) which satisfies (14) according to (18).

Let us now suppose thatA(0)<0 and A(1)>0. Whenx→+∞, f(x) =− 2

σ2A

1− Z +

x f(y)dy

∼2A(1) σ2

Z +

x f(y)dy.

By (14), ff(x)(x) = (logf(x))=−σ22A1/2(x)) converges to −2Aσ2(1) as x→ +∞. This implies that log(fx(x)) converges to−2Aσ2(1) and thatxf(x)1{x0} is integrable. Moreover, since R0+f(y)dy = +∞, R0x f(y)dy2Aσ2(1)R0xff2(y)(y)dy∼

σ2

2A(1)f(x), asx→+∞. In the same way, one obtains the equivalents given in (15) whenx→ −∞and checks the integrability of the functionxf(x)1{x0}. From (15), one has

x→−∞lim Z x

−∞

f(y)dy Z 0

x

dy

f(y)= σ4 4(A(0))2 and

xlim+

Z +

x

f(y)dy Z x

0

dy

f(y) = σ4 4(A(1))2.

By Theorem 6.2.2, page 99 of [1], one concludes that the measure with densityf satisfies a Poincar´e inequality.

By (14), the function f is C2 as soon as the function A is C2 on [0,1].

Moreover,f′′(x) =−σ22A′′(H∗f(x))f2(x)−σ22A(H∗f(x))f(x) which com- bined with (14) implies (16).

Remark 2.2. WhenAis aC1convex function on [0,1] such thatA(0) = A(1) = 0 andA(u)<0 for some u∈(0,1), then the necessary and sufficient condition in Lemma2.1is obviously satisfied. Since (14) implies

(logf(x))′′=

f(x) f(x)

=

−2/σ2A(H∗f(x))f(x) f(x)

=− 2

σ2A′′(H∗f(x))f(x)≤0,

the probability measures solving the stationary Fokker–Planck equation admit log-concave densities with respect to the Lebesgue measure. Log- concavity is a property stronger than the existence of a Poincar´e inequality (see [7]).

Example 2.3. Using (18) and (19), the following two choices forAlead to exact computations and different tails for the stationary densities:

(15)

• ifA(x) =12x(x−1), one gets log(1ϕ1/2ϕ (x)

1/2(x)) =x/σ2, that is, ϕ1/2(x) = ex/σ2

1 +ex/σ2 and ϕ1/2(x) = 1

2cosh2(x/2σ2);

• ifA(x) =x3−x=x(x−1)(x+ 1), ϕ√1/2(x) = 1

p1 +e4x/σ2 and ϕ

1/2(x) = 2e4x/σ2 σ2(1 +e4x/σ2)3/2. When A(1) = 0 and A(u)<0 for all u∈(0,1), a natural question is how to link the translation parameter of the candidate long time limit of the marginal Pt solving the stationary Fokker–Planck equation to the initial marginalm. WhenRR|x|m(dx)<+∞, by (3), for allt≥0,E(Xt1) =E(X01).

Therefore the translation parameter is chosen in order to ensure that the invariant measure has the same mean as the initial measure m.

Let us denote by pt the density of Pt and byFt=H∗Pt its cumulative distribution function.

Theorem 2.4. Let A be C2 on [0,1] satisfying (H). Assume that m admits a density p0 such that RR|x|p0(x)dx <+∞ and RR

(p0(x)p(x))2 p(x) dx is small enough where p denotes the stationary distribution with same expectation asp0. Last, we suppose thatAandp0 are such thatpis a smooth solution of (13). Then RR

(pt(x)p(x))2

p(x) dx converges to 0 exponentially fast as t→+∞.

By a smooth solution of (13), we mean thatppossesses enough regularity and integrability so that the formal computations made in the proof below are justified.

Example 2.5. WhenA(x) =12(x2−x), one easily checks that the func- tionφ(t, x) =−Ft(x+2t) solves Burgers’ equation

tφ=σ2

2 ∂xxφ−1

2∂xφ2, φ(0, x) =−F0(x).

By the Cole–Hopf transformation, ψ(t, x) = exp(−σ12R−∞x φ(t, y)dy) solves the heat equation

tψ=σ2

2 ∂xxψ, ψ(0, x) = exp 1

σ2 Z x

−∞F0(y)dy

. Since Ft(x) =σ2xψψ(t, x−2t), one deduces that

Ft(x) = R

Re(xt/2y)2/2σ2tF0(y)ψ(0, y)dy/(σ√ 2πt) R

Re(xt/2y)2/2σ2tψ(0, y)dy/(σ√

2πt) . (20)

Références

Documents relatifs

In this work, generalizing the techniques introduced by Jabir-Talay-Tomasevic [3], we prove the well- posedness and propagation of chaos for a stochastic particle system in

The idea is to prove in a first step the convergence for initial data which are bounded from above or from below by a solution of equation (5), and then to extend this result

We have presented two guaranteed a posteriori error estimators for the eddy current problems and proved the global upper bound for the error by the estima- tor based on the

2D Navier-Stokes equation, Stochastic particle systems, Propagation of Chaos, Fisher information, Entropy dissipation.. The first and third authors of this work were supported by

Theorem 1.8. This paper is some kind of adaptation of the work of Fournier- Hauray-Mischler in [9] where they show the propagation of chaos of some particle system for the 2D

For SNR ∈ [0dB; 2.19db]: the reduced trajectory converges to a very small sized volume of E BP that we can consider as a fixed point whereas λ is close to the null value, for SNR

We show that under suitable assumptions, there is creation of chaos then propagation of this chaos for a mean-field system of particles without assuming that the initial

Fouad M.A.,la prise en charge des infections de la main à propos de