• Aucun résultat trouvé

Propagation of chaos for stochastic particle systems with singular mean-field interaction of $L^q − L^p$ type

N/A
N/A
Protected

Academic year: 2021

Partager "Propagation of chaos for stochastic particle systems with singular mean-field interaction of $L^q − L^p$ type"

Copied!
11
0
0

Texte intégral

(1)

HAL Id: hal-03086253

https://hal.inria.fr/hal-03086253

Preprint submitted on 22 Dec 2020

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 for stochastic particle systems with singular mean-field interaction of L

q

− − L

p

type

Milica Tomasevic

To cite this version:

Milica Tomasevic. Propagation of chaos for stochastic particle systems with singular mean-field inter- action ofLq− −Lp type. 2020. �hal-03086253�

(2)

Propagation of chaos for stochastic particle systems with singular mean-field interaction of L

q

− L

p

type

Milica Tomaˇsevi´c

Abstract

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 mean-field interaction under the assumption that the interacting kernel belongs to a suitableLqt−Lpxspace. Contrarty to the large deviation principle approach by [2], the main ingredient of the proof here are the partial Girsanov transformations introduced in [3].

1 Introduction

In this work, we prove the propagation of chaos of the particle system ( dXti,N = N1 PN

j=1b(t, Xti,N, Xtj,N)dt+ 2dWti,

X0i,N i.i.d. and independent of W := (Wi,1iN), (1.1) towards the non-linear stochastic differential equation

( dXt=R

(b(t, Xt, y)ρt(y)dy dt+

2dWt, t >0,

ρs(y)dy:=L(Xs), X0ρ0(x)dx, (1.2)

under the assumption that

(Hb) |b(t, x, y)| ≤ht(xy) for somehLqloc(R+;Lp(Rd)), wherep, q(2,∞) satisfy dp+2q <1.

In (1.1) and (1.2) Wi’s and W are independent standard d dimensional Brownian motions defined on a probability space (Ω,F,P)

The propagation of chaos of (1.1) towards (1.2) has been very recently established in the literature using large deviation principle theory [2]. In this work, we propose a completely different approach relying on the techniques introduced in [3] (where the present author is one of the co-authors).

The problem treated in [3] is the case of probabilistic interpretation of the parabolic-parabolic Keller-Segel model in one-dimensional setting. The particles interaction is both non-Markovian and singular and as such does not enter in the framework of [2]. However, the propagation of chaos was established and it seemed important to adapt the techniques from [3] to the setting in (1.1) and make them available for other authors, rather than treating the problems with singular interactions in a case by case studies.

Despite the one-dimensional setting, the problem treated in [3] presents a notable difficulty due to the unusual interaction between the particles. Namely, instead of b(t, Xti,N, Xtj,N) in (2.1), one has Rt

0Kt−s(Xti,N Xsj,N)dswhereK is the singular kernelKt(x) = −x

2πt3/2ex2t2.

Inspired by [5], the authors in [3] first prove the weak well-posedness of the particle system by means of a Girsanov transformation exploting the integrability properties (in time and space) of the interaction kernel.

However, in order to get propagation of chaos, the above transformation is not helpful as the estimate on the exponential moment of the drift (appearing in the Novikov condition) explodes as N → ∞. This dif- ficulty is circumvented by an original idea to perform on the particle system the so-called partial Girsanov transformations. These partial transformations fix a finite number of particlesk < N and transform them to independent Brownian motions while they also take off their dependence of the restN kparticles. Their advantage is that the estimates on the respective exponential martingales do not explode with N.

When one proves by hand the propagation of chaos, these transformations naturally appear as the particle

(3)

system is exchangeable. Hence, as it was noted in Sznitman [8, p. 180], what one needs to control are functionals of finite number of particles (usually at most 4).

In this paper, we adapt the above techniques to the setting in (1.1). The fact the interaction is Markovian simplifies lightly the computations. However, the multidimensional setting and the general assumptions on b, complicate the computations.

We finish this section by noting that in [6, Theorem 1.1] strong well-posedness of (1.2) is established.

Also, a Gaussian estimate on the one dimensional time marginal densities is proved. In this work, as we show the convergence in law of the empirical measure of (1.1) towards the law of the solution to (1.2), we will only work with the martingale problem related to (1.2) that we formulate precisely in the next section.

2 Main Results

Let t > 0. As the interaction kernel b(t,·,·) has only integrability properties w.r.t. the second or third variable, it may be unbounded or not well defined in certain points. Let us denote byNb(t) the following set:

Nb(t) ={(x, y)Rd×Rd: lim

(x0,y0)→(x,y)|b(t, x0, y0)|= or lim

(x0,y0)→(x,y)b(t, x0, y0) does not exist}.

However, as|b(t, x, y)| ≤ht(xy) andhtLp(Rd), the setNb(t) is of Lebesgue’s measure zero inRd×Rd. Thus, in order to ensure that all the termes appearing in it are well defined, the particle system reads

( dXti,N =N1 PN

j=1,j6=ib(t, Xti,N, Xtj,N)1{(Xi,N

t ,Xtj,N)/∈Nb(t)} dt+ 2dWti,

X0i,N i.i.d. and independent ofW := (Wi,1iN). (2.1) Here Wi’s are N independent standard d-dimensional Brownian motions on a probability space (Ω,F,P).

Notice here that we assume that a particle does not interact with itself.

In practical examples, ht explodes (or it is not well defined) only in one point (zero) and thus, b(t,·,·) may explode (or not be well defined) on the line x=y. In order to keep the result as general as possible, we will rather work with the sets (Nb(t))t≥0. However, the reader should have in mind that this is just a technicality and that, in practice1{(Xi,N

t ,Xtj,N)/∈Nb(t)} becomes1{Xi,N

t 6=Xtj,N} (as in [3]).

It is reasonable to assume thatb(t,·,·) is continuos outside ofNb(t) as in practice one deals with interaction kernels in the form of convolutions that are well defined and continuos almost everywhere (like ±|x|xr).

Our first main result is weak well-posedness of (2.1).

Theorem 2.1. Assume (Hb). Given 0< T < andN N, there exists a weak solution(Ω,F,(Ft; 0 tT),QN, W, XN)to the N-interacting particle system (2.1)that satisfies, for any 1iN,

QN

Z T

0

1 N

N

X

j=1,j6=i

b(t, Xti,N, Xtj,N)1{(Xi,N

t ,Xj,Nt )/∈Nb(t)}

2

dt <

= 1. (2.2)

In view of Karatzas and Shreve [4, Chapter 5, Proposition 3.10], one has the following uniqueness result:

Corollary 2.2. Weak uniqueness holds in the class of weak solutions satisfying (2.2).

Before we state the propagation of chaos result, we formulate the martingale problem associated to (1.2).

Definition 2.3. Q∈ P(C[0, T];Rd) is a solution to (MP) if:

(i) Q0=µ0;

(ii) For anyt(0, T]and any r >1, the one dimensional time marginal Qt ofQ has a density ρt w.r.t.

Lebesgue measure onRd which belongs toLr(Rd) and satisfies

∃CT, 0< tT, tkLr(Rd) CT td2(1−r1);

(4)

(iii) Denoting by (x(t);t T) the canonical process of C([0, T];Rd), we have: For any f Cb2(Rd), the process defined by

Mt:=f(x(t))f(x(0)) Z t

0

∇f(x(s))·Z

b(s, x(s), y)ρs(y)dy

+4f(x(s))

ds is aQ-martingale w.r.t. the canonical filtration.

Remark 2.4. Under the assumption (Hb)and that the measure µ0 has a finite β-order moment for some β > 2, the martingale problem (MP) admits a unique solution according to [6, Thm. 1.1]. In fact, [6, Thm. 1.1] gives strong well-posedness of (1.2)and proves the Gaussian estimates punctually on the marginal densities. In the martingale formulation it is enough to impose such estimates in Lr-norms as along with (Hb)this will ensure that all the terms in the definition of the process(M)are well defined.

It is classical that a suitable notion of weak solution to (1.2) is equivalent to the notion of solution to (MP) (see e.g. [4]).

We are ready to state our second main result, the propagation of chaos of (2.1).

Theorem 2.5. In addition to (Hb), assume that for any t > 0, b(t,·,·) is continuous outside of the set Nb(t). Assume that theX0i,N’s are i.i.d. and that the initial distribution ofX01,N is the measureµ0 that for someβ >2has finiteβ-order moment . Then, the empirical measureµN = N1 PN

i=1δXi,N of (2.1)converges in the weak sense, when N → ∞, to the unique weak solution of (1.2).

3 Proof of Theorem 2.1

We start from a probability space (Ω,F,(Ft; 0 t T),W) on which d - dimensional Brownian motions (W1, . . . , WN) and the random variables X0i,N(see (2.1)) are defined. Set ¯Xti,N :=X0i,N +Wti (tT) and X¯ := ( ¯Xi,N,1iN). Denote the drift terms in (2.1) bybi,Nt (x),xC([0, T];Rd)N, and the vector of all the drifts asBtN(x) = (b1,Nt (x), . . . , bN,Nt (x)). For a fixedN N, consider

ZTN := exp (Z T

0

BtN( ¯X)·dWt1 2

Z T 0

BNt ( ¯X)

2dt )

.

To prove Theorem2.1, it suffices to prove the following Novikov condition holds true (see e.g. [4, Chapter 3, Proposition 5.13]): For anyT >0,N 1,κ >0, there existsC(T, N, κ) such that

EW exp (

κ Z T

0

|BNt ( ¯X)|2dt )!

C(T, N, κ). (3.1)

Drop the indexN for simplicity. Using the definition of (BtN) and Jensen’s inequality one has

EW

"

exp (

κ Z T

0

BtN( ¯X)

2 dt )#

EW

exp

1 N

N

X

i=1

1 N

N

X

j=1,j6=i

Z T 0

κN|b(t,X¯ti,X¯tj)|2dt

, from which we deduce

EW

"

exp (

κ Z T

0

BtN( ¯X)

2 dt )#

1 N

N

X

i=1

1 N

N

X

j=1,j6=i

EW

"

exp (

κN Z T

0

|b(t,X¯t i,X¯t

j)|2dt )#

.

Assume for a moment that fori, jN such thatj 6=ione has EW

"

exp (

κN Z T

0

|b(t,X¯t i,X¯t

j)|2dt )#

C(T, N). (3.2)

Then (3.1) is satisfied and the proof is finished.

The rest of the proof will be devoted to establishing (3.2). Actually, we will prove the following more general statement:

(5)

Proposition 3.1. Let T >0. Suppose the hypothesis (Hb). Letw:= (wt)be a (Gt)-Brownian motion with an arbitrary initial distribution µ0 on some probability space equipped with a probability measure P and a filtration (Gt). Suppose that the filtered probability space is rich enough to support a continuous process Y independent of (wt). For any α >0, one has

EP

"

exp (

α Z T

0

|b(t, wt, Yt)|2dt )#

C(T, α),

whereC(T, α)depends only on T andα, but does neither depend on the lawL(Y)nor ofµ0. To prove Proposition3.1, we need some preparation in form of two auxiliary Lemmas.

First, for (t, x)[0, T]×Rd, denote by gt(x) := (2πt)1d/2e|x|

2

2t . Note that for anyt >0 and anyp1, one has

kgtkp= Cp td2(1−1p)

. (3.3)

Lemma 3.2. Suppose the hypothesis(Hb). Letw:= (wt)be a(Gt)-Brownian motion with an arbitrary initial distribution µ0 on some probability space equipped with a probability measureP and a filtration (Gt). There exists a universal real number C0>0 such that

∀xC([0, T];R), ∀0t1t2T, Z t2

t1

EGPt1|b(t, wt, xt)|2dtC0(T)(t2t1)q−2q dp. Proof. Using the assumption (Hb), one has

I:=

Z t2

t1

EGPt1|b(t, wt, xt)|2dt Z t2

t1

Z

h2t(y+wt1xt)gt−t1(y)dy dt Applying H¨older inequality in space with p2 >1 and afterwards in time with q2>1, one has

I Z t2

t1

khtk2Lp(Rd)kgt−t1k

L

p

p−2(Rd)≤ khkLq((0,T);Lp(Rd))

Z t2

t1

kgt−t1k

L

p p−2(Rd)

q−2q dt

q−2 q

.

According to (3.3), we have

IC0(T) Z t2

t1

1 (tt1)dpq−2q

dt

!q−2q .

For the last integral to be finite, one needs to have dp < q−2q = 12q. This is exactly the constraint in the hypothesis (Hb). Thus, one obtains

IC0(T)tq−2q dp.

Lemma 3.3. Same assumptions as in Lemma 3.2. Let C0(T) be as in Lemma 3.2. For any κ > 0, there existsC(T, κ)independent ofµ0 such that, for any0T1T2T satisfying T2T1<(C0(T)κ)

q−21

q d p,

∀xC([0, T];R), EGPT1

"

exp (

κ Z T2

T1

|b(t, wt, xt)|2dt )#

C(T, κ).

Proof. We adapt the proof of Khasminskii’s lemma in Simon [7]. Admit for a while we have shown that there exists a constantC(κ, T) such that for anyM N

M

X

k=1

κk k!EGPT1

Z T2 T1

|b(t, wt, xt)|2 dt

!k

C(T, κ), (3.4)

provided thatT2T1<(C0(T)κ)

q−21

q d

p. The desired result then follows from Fatou’s lemma.

(6)

We now prove (3.4).

By the tower property of conditional expectation,

EGPT1

Z T2

T1

|b(t, wt, xt)|2dt

!k

=k!

Z T2 T1

Z T2 t1

Z T2 t2

· · · Z T2

tk−2

Z T2 tk−1

EGPT1

h|b(t1, wt1, xt1)|2|b(t2, wt2, xt2)|2

× · · · × |b(tk−1, wtk−1, xtk−1)|2 E

Gtk−1

P |b(tk, wtk, xtk)|2 i

dtkdtk−1· · · dt2dt1. In view of Lemma3.2,

Z T2 tk−1

E

Gtk−1

P |b(tk, wtk, xtk)|2dtk C0(T)(T2tk−1)q−2q dp C0(T)(T2T1)q−2q dp. Therefore, by Fubini’s theorem,

EGPT1

Z T2

T1

|b(t, wt, xt)|2dt

!k

k!C0(T)(T2T1)q−2q dp Z T2

T1

Z T2

t1

Z T2

t2

· · · Z T2

tk−2

EGPT1

h|b(t1, wt1, xt1)|2|b(t2, wt2, xt2)|2

× · · · × |b(tk−1, wtk−1, xtk−1)|2i

dtk−1· · · dt2dt1.

Now we repeatedly condition with respect toGtk−i (i2) and combine Lemma3.2with Fubini’s theorem.

It comes:

EGPT1 Z T2

T1

|b(t, wt, xt)|2dt

!k

k!(C0(T)(T2−T1)q−2q dp)k−1 Z T2

T1

EGPT1|b(t1, wt1, xt1)|2dt1k!(C0(T)(T2−T1)q−2q dp)k.

Thus, (3.4) is satisfied provided thatT2T1<(C0(T)κ)

q−21

q d p.

Finally, the proof of Proposition3.1is identical to the proof of [3, Prop. 3.3] where F is replaced byb2 and the previous lemma is used instead of [3, Lemma 3.2].

4 Propagation of chaos

4.1 Girsanov transform for 1≤r < N particles

For any integer 1 r < N, proceeding as in the proof of Theorem 2.1 one gets the existence of a weak solution on [0, T] to

dXbtl,N =dWtl, 1lr, dXbti,N =n

1 N

PN

j=r+1b(t,Xbti,N,Xbtj,N)o

dt+dWti, r+ 1iN, Xb0i,N i.i.d. and independent of (W) := (Wi,1iN).

(4.1)

Below we set ˆX := ( ˆXi,N,1 i N) and we denote by Qr,N the probability measure under which ˆX is well defined. Notice that (Xbl,N,1 l r) is independent of (Xbi,N, r+ 1 i N). We now study the exponential local martingale associated to the change of drift between (2.1) and (4.1). ForxC([0, T];Rd)N set

βt(r)(x) :=

b1,Nt (x), . . . , br,Nt (x), 1 N

r

X

i=1

b(t, xr+1t , xit), . . . , 1 N

r

X

i=1

b(t, xNt , xit) . In the sequel we will need uniform w.r.tN bounds for moments of

ZT(r):= exp (

Z T

0

βt(r)(X)b ·dWt1 2

Z T 0

t(r)(X)|b 2dt )

. (4.2)

(7)

Proposition 4.1. For any T >0,γ >0andr1there existsN0r andC(T, γ, r)s.t.

∀N N0, EQr,Nexp (

γ Z T

0

(r)t (Xb)|2dt )

C(T, γ, r).

Proof. ForxC([0, T];Rd)N, one has

t(r)(x)|2=

r

X

i=1

1 N

N

X

j=1

b(t, xit, xjt)

2

+ 1 N2

N−r

X

j=1 r

X

i=1

b(t, xr+jt , xit)

!2 .

By Jensen’s inequality,

t(r)|2 1 N

r

X

i=1 N

X

j=1

|b(t, xit, xjt)|2+ r N2

N−r

X

j=1 r

X

i=1

|b(t, xr+jt , xit)|2.

For simplicity we below writeE(respectively, ˆXi) instead ofEQr,N (respectively, ˆXi,N). Observe that Eexpn

γ Z T

0

t(r)(X)|b 2dto

EexpnXr

i=1

N

N

X

j=1

Z T 0

|b(t,Xbti,Xbtj)|2dto1/2

Eexpn2γr N2

N−r

X

j=1 r

X

i=1

Z T 0

|b(t,Xbtr+j,Xbti)|2dto1/2

Yr

i=1

1 N

N

X

j=1

Eexpn 2γr

Z T 0

|b(t,Xbti,Xbtj)|2dto2r1NY−r

j=1

1 r

r

X

i=1

Eexpn2γr2 N

Z T 0

|b(t,Xbtr+j,Xbti)dto2(N−r)1 .

In view of Proposition3.1, the proof is finished.

4.2 Tightness

We start with showing the tightness of N} and of an auxiliary empirical measure which is needed in the sequel.

Lemma 4.2. Let QN be as above. The sequence N} is tight under QN. In addition, let νN :=

1 N4

PN

i,j,k,l=1δXi,N

. ,X.j,N,X.k,N,X.l,N. The sequence N} is tight underQN.

Proof. The tightness of N}, respectively N}, results from the tightness of the intensity measure {EQNµN(·)}, respectively{EQNνN(·)}: See Sznitman [8, Prop. 2.2-ii]. By symmetry, in both cases it suffices to check the tightness of{Law(X1,N)}. We aim to prove

∃C >0,∀N N0, EQN[|Xt1,NXs1,N|4]CT|ts|2, 0s, tT, (4.3) where N0is as in Proposition4.1. LetZT(1) be as in (4.2). One has

EQN[|Xt1,NXs1,N|4] =EQ1,N[(ZT(1))−1|Xbt1,NXbs1,N|4].

AsXb1,N is a one dimensional Brownian motion underQ1,N,

EQN[|Xt1,NXs1,N|4](EQ1,N[(ZT(1))−2])1/2(EQ1,N[|Xbt1,N Xbs1,N|8])1/2(EQ1,N[(ZT(1))−2])1/2C|ts|2. Observe that, for a Brownian motion (W]) underQ1,N,

EQ1,N[(ZT(1))−2] =EQ1,Nexp (

2 Z T

0

βt(1)( ˆX)·dWt] Z T

0

t(1)( ˆX)|2dt )

.

Adding and subtracting 3RT

0 (1)t |2dtand applying again the Cauchy-Schwarz inequality, EQ1,N[(ZT(1))−2] EQ1,N exp

( 6

Z T 0

t(1)( ˆX)|2dt )!1/2

. Applying Proposition 4.1withk= 1 andγ= 6, we obtain the desired result.

Références

Documents relatifs

Keywords: stochastic particle systems, mean-field limit, propagation of chaos, stochas- tic partial differential equations, Cucker-Smale model, collective

Large Scale Behavior For this model, the state pro- cess of the numbers of cars and reserved parking spaces is an irreducible Markov process on a finite state space with no ex-

Mean field propagation of infinite dimensional Wigner measures with a singular two-body interaction potential.. Zied Ammari,

Mean-field approximation of stochastic population processes in games 4 time intervals, gives the expected net increase in each population share during the time interval, per time

As in [42], the core of the material presented here is the mean field limit of N -particle systems governed by the equations of classical mechanics in the case where the

In the present article, we complete this study with functional moderate deviations of mean field particle models for both the τ -topology on the space of signed and bounded measures

In the other hand, by using the convergence of McKean-Vlasov processes in long-time and the propagation of chaos, we prove that there exist traps such that the diffusion can not

Under suitable assumptions, if the system at time t = 0 is chaotic (that is to say the particles are independent), this chaos propa- gates as the number of particles goes to