• Aucun résultat trouvé

Donsker-Type Theorem for BSDEs: Rate of Convergence

N/A
N/A
Protected

Academic year: 2021

Partager "Donsker-Type Theorem for BSDEs: Rate of Convergence"

Copied!
28
0
0

Texte intégral

(1)

HAL Id: hal-02165750

https://hal.archives-ouvertes.fr/hal-02165750

Preprint submitted on 26 Jun 2019

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.

Donsker-Type Theorem for BSDEs: Rate of Convergence

Philippe Briand, Christel Geiss, Stefan Geiss, Celine Labart

To cite this version:

Philippe Briand, Christel Geiss, Stefan Geiss, Celine Labart. Donsker-Type Theorem for BSDEs:

Rate of Convergence. 2019. �hal-02165750�

(2)

Donsker-Type Theorem for BSDEs:

Rate of Convergence

Philippe Briand ∗† Christel Geiss Stefan Geiss Céline Labart This version: 2019-06-24

Abstract

In this paper, we study in the Markovian case the rate of convergence in the Wasserstein distance of an approximation of the solution to a BSDE given by a BSDE which is driven by a scaled random walk as introduced in Briand, Delyon and Mémin (Electron. Comm. Pro- bab. 6(2001), 1–14).

1. Introduction.

In this paper, we are concerned with the discretization of solutions to BSDEs of the form Yt=G(B) +

Z T

t

f(Bs, Ys, Zs)dsZ T

t

ZsdBs, 0≤tT,

where B is a standard Brownian motion. These equations have been introduced by Jean- Michel Bismut for linear generators in [2] and by Étienne Pardoux and Shige Peng for Lipschitz generators in [14].

In one of the first studies on this topic, in the case where the generator f may depend on z as well, Philippe Briand, Bernard Delyon and Jean Mémin [5] proposed an approximation based on Donsker’s theorem. They showed that the solution (Y, Z) to the previous BSDE can be approximated by the solution (Yn, Zn) to the BSDE

Ytn=G(Bn) + Z

]t,T]

f(Bns−, Ys−n, Zsn)dhBnis+ Z

]t,T]

ZsndBns, 0≤tT, where Bn is the scaled random walk

Bnt = q

T /n

[nt/T]

X

k=1

ξk, 0≤tT,

and (ξk)k≥1 is an i.i.d. sequence of symmetric Bernoulli random variables. They proved, in full generality, meaning that G(B) is only required to be a square integrable random variable, that (Yn, Zn) converges to (Y, Z). However, the question of the rate of convergence was left open. Right now it seems to be hopeless to get a result in this direction for such a general path- dependent terminal condition G(B).But in the Markovian case, meaning that G(B) = g(BT),

In memory of Jean Mémin from whom I learned lots of mathematics.

Many thanks to Pierre Baras for very fruitful discussions about the heat equation.

(3)

this problem seems to be tractable, in particular due to the PDE structure behind. Indeed, (Y, Z) is related to the semilinear heat equation

tu(t, x) +1

2∆u(t, x) +f(t, x, u(t, x),∇u(t, x)) = 0, (t, x)∈[0, T[×R, u(T,·) =g, (1) where, under certain regularity conditions, we can choose

Ys=u(s, Bs) and Zs=∇u(s, Bs).

In the case where Bn is the discretized Brownian motion, the link to PDEs was exploited in [18,3] to get the rate of convergence, in the Markovian case, of the classical scheme for BSDEs.

The convergence of this scheme was already proved in [6, Proposition 13] for a general terminal condition and a generator that is Lipschitz in its spatial coordinates but without any rate of convergence.

Even though the link with PDEs was pointed out in [5], the rate of convergence of the approximation of BSDEs given by scaled random walks was completely open. In two recent papers, Christel Geiss, Céline Labart and Antti Luoto [10,9] give a first answer to this question.

They showed that the error between (Yn, Zn) and (Y, Z) is of order n−ε/4 when g is assumed to be ε-Hölder continuous and f(t,·) Lipschitz continuous. One of the main arguments in these papers consists in constructing the random walk from the Brownian motion B using the Skorohod embedding (see [17]) together with generalizations of the pioneering work of Jin Ma and Jianfeng Zhang [13] on representation theorems for BSDEs. This approach allows to work with convergence in the L2-sense even if the problem naturally arises in the weak sense. The drawback is that the rate of convergence n−ε/4 obtained in these papers is not optimal as one can expect n−ε/2.

The objective of our study is to improve the previous rate of convergence by using a weak limit approach, the error being measured in the Wasserstein distance. Our starting point is a result of Emmanuel Rio [15] who proved that, whenT = 1, for allr≥1, there exists a constant Cr such that, for alln≥1,Wr(Bn1, G)Crn−1/2, whereWr is the Lr-Wasserstein distance and Ga standard normal random variable (see Section 3). Firstly, we generalize this result to cover the case wheref ≡0 which corresponds to the heat equation. Then, using the associated PDE, in particular representation formulas in the spirit of [13], we are able to prove that

Wr(Ytn, Yt)≤Crn−ε/2, and Wr(Ztn, Zt)≤ √Cr

Ttn−ε/2 for all t∈[0, T[,

whengandf(t,·, y, z) areε-Hölder continuous, which gives the correct rate of convergence when the error is measured in the Wasserstein-distance. We refer to Theorem 10 in Section5 for the precise statement.

2. Notation.

In all the sequel, T > 0 is a fixed positive real number. We work on a complete probability space (Ω,F,P) carrying a standard real Brownian motion{Bt}0≤t≤T,and{Ft}0≤t≤T stands for the augmented filtration of B which is right continuous and complete.

We consider the following BSDE Yt=g(BT) +

Z T t

f(s, Bs, Ys, Zs)dsZ T

t

ZsdBs, 0≤tT. (2) Throughout this article, we will assume for the function g defining the terminal condition and the generatorf the following:

(4)

Assumption (A1). There exist 0< ε≤1 and 0< α≤1 such that it holds:

(i) The functiong:R−→R isε-Hölder continuous: for all (x, x0)∈R2 one has

g(x)g x0≤ kgkε xx0

ε.

(ii) The function f : [0, T]×R×R×R −→ R is α-Hölder continuous in time, ε-Hölder continuous in space and Lipschitz continuous with respect to (y, z): for all (t, x, y, z) and (t0, x0, y0, z0) in [0, T]×R×R×R one has

f(t, x, y, z)−f t0, x0, y0, z0

≤ kftkα tt0

α+kfxkε xx0

ε+kfykLip yy0+kfzkLip zz0. (3) Most of the time, we do not need to distinguish between kfykLip and kfzkLip and we let kfkLip:= max (kfykLip,kfzkLip).

Convention: Later the phrase that a constant C >0 depends on (T, ε, f, g) stands for the fact that C can be expressed in terms of (T, ε,kfxkε,kfykLip,kfzkLip, Kf,kgkε, g(0)) where

Kf := sup

t∈[0,T]

|f(t,0,0,0)|.

Similarly, a dependence on (T, α, ε, f, g) means an additional dependence on (α,kftkα).

From [4, Theorem 4.2] it is known that under (A1), the BSDE (2) has a unique Lp-solution (Y, Z) for anyp∈]1,∞[.So for (t, x)∈[0, T[×Rwe let Yst,x, Zst,xs∈[t,T]be the square integrable solution to the BSDE

Yst,x =gBt,xT + Z T

s

fr, Brt,x, Yrt,x, Zrt,xdrZ T

s

Zrt,xdBr, tsT, (4) whereBrt,x:=x+Br−Bt, and set, as usual, forx∈R,u(T, x) :=g(x), and, for (t, x)∈[0, T[×R,

u(t, x) :=Ytt,x =E

"

g(BTt,x) + Z T

t

fr, Brt,x, Yrt,x, Zrt,xdr

# .

It is well known that the function u is continuous on [0, T]×R (see also Lemma 6 below) and under Lipschitz assumptions in (x, y, z) and for α12 it is the viscosity solution to (1), see [20, Theorem 5.5.8]. Moreover, in this Markovian setting, for (t, x) ∈ [0, T]×R, we have Yst,x =u(s, Bst,x) a.s. for alls∈[t, T]. In [19, Theorem 3.2], for a generator which is Lipschitz continuous in all space variables and a measurable g with polynomial growth, J. Zhang proved that u belongs to C0,1([0, T[×R) and that Zst,x = ∇u(s, Bst,x) a.e. on [t, T[×Ω. Moreover, the following representation holds

∇u(t, x) =E

"

g(BTt,x)BTBt Tt +

Z T t

fr, Brt,x, Yrt,x, Zrt,xBrBt rt dr

#

, (t, x)∈[0, T[×R.

If F is the function given by

F(s, x) :=f(s, x, u(s, x),∇u(s, x)) for (s, x)∈[0, T[×R, (5) we thus have

u(t, x) =E

"

g(BTt,x) + Z T

t

F(r, Bt,xr )dr

#

, (t, x)∈[0, T[×R, (6)

(5)

together with

∇u(t, x) =E

"

g(BTt,x)BTBt Tt +

Z T t

F(r, Brt,x)BrBt rt dr

#

, (t, x)∈[0, T[×R. (7) These formulas play an important role in the sequel.

In Section 4 and in the appendix, we extend these results to the case where f(t,·, y, z) is ε-Hölder continuous and make the regularity of uand ∇u precise.

As mentioned before, we are concerned with the approximation of the solution Yt,x, Zt,x to (4) by a solution to the BSDE driven by a scaled random walk. To do this, let us consider, on some probability space, not necessarily (Ω,F,P), an i.i.d. sequence (ξk)k≥1 of symmetric Bernoulli random variables. For n ∈ N := {1,2,3, ...} we set h := T /n and we consider the scaled random walk

Btn:=

h

[t/h]

X

k=1

ξk, 0≤tT,

where [x] := max{r∈Z:rx} for any real numberx.As we did for the Brownian motion, for x∈Rand 0≤tsT we put

Bsn,t,x:=x+BsnBtn.

Let us introduce some further notation. We denote the ceiling function bydxe := min{r ∈ Z:rx}for x∈R.Moreover, we set

nt:= [t/h], t:=h[t/h] =hnt and t:=hdt/he, t∈[0, T].

For n∈N let us consider the following BSDE driven byBn: Ytn=g(BTn) +

Z

]t,T]

f(s, Bsn, Ysn, Zsn)dhBnisZ

]t,T]

ZsndBsn, t∈[0, T].

It was shown in [5] that, as soon as hkfkLip < 1, this BSDE has a unique square integrable solution (Yn, Zn), Yn being adapted and Zn being predictable with respect to the filtration generated by Bn. By construction, Yn is a piecewise constant càdlàg process with Ytn =Ytn. The process Zn is defined as an element of L2(Ω×[0, T], dP⊗dhBni), where we start with a Zn defined only on the points {kh : k = 1, . . . , n} and extend it to ]0, T] as a càglàd process (Ztn)t∈]0,T] by setting Ztn = Ztn. The previous BSDE is actually a discrete BSDE that can be solved by hand since, for k= 0,· · · , n−1, we have

Ykhn =Y(k+1)hn +h f(k+ 1)h, Bkhn , Ykhn, Z(k+1)hn −√

h Z(k+1)hn ξk+1, Ynhn =g(BTn).

Thus, ifY(k+1)hn is given, Z(k+1)hn =h−1/2E

Y(k+1)hn ξk+1| Fkhn , (8)

Ykhn =Y(k+1)hn +h f(k+ 1)h, Bkhn , Ykhn, Z(k+1)hn −√

h Z(k+1)hn ξk+1

=E

Y(k+1)hn | Fkhn +h f(k+ 1)h, Bkhn , Ykhn, Z(k+1)hn , (9) where the last equality follows by taking the conditional equation w.r.t. Fkhn of the second line.

(6)

Since we are in a Markovian setting, there is also an analog of the Feynman-Kac formula. If u is a given function we set

Dn+u(x) := 1 2

u(x+√

h) +u(x−√

h), Dnu(x) := 1 2

u(x+√

h)u(x−√ h), and

nu(x) :=h−1/2Dnu(x). (10)

Remark 1. From the definition ofDn+ and Dn, we get that ifu isε-Hölder,Dn+u and Dnu are also ε-Hölder with constantkukε.

LetUnbe the solution to the finite difference equation, where forx∈Randk= 0, . . . , n−1 we require

( Un(kh, x) =Dn+Un((k+ 1)h, x) +hf((k+ 1)h, x, Un(kh, x), h−1/2DnUn((k+ 1)h, x)), Un(nh, x) =g(x).

(11) Then, we obtain from (8) and (9) (cf [5, Proposition 5.1]) that

Ykhn =Un kh,

h

k

X

i=1

ξi

!

, k= 0, . . . , n, Zkhn =∇nUn kh,

h

k−1

X

i=1

ξi

!

, k= 1, . . . , n.

These formulas rewrite in continuous time to

Ytn=Ytn=Un(t, Btn), t∈[0, T], and Ztn=Ztn=∇nUn t, Bt−n , t∈]0, T].

If we set, for 0≤tT and x∈R,Un(t, x) :=Un(t, x), we haveYtn=Un(t, Btn).

More generally, for 0≤t < T, we define (Yn,t,x, Zn,t,x) as the solutionYn,t,x= (Ysn,t,x)s∈[t,T]

and Zn,t,x= (Zsn,t,x)s∈]t,T] to the BSDE Ysn,t,x=g(BTn,t,x) +

Z

]s,T]

f(r, Brn,t,x , Yrn,t,x , Zrn,t,x)dhBnirZ

]s,T]

Zrn,t,xdBnr, s∈[t, T]. (12) We set YTn,T ,x =g(x). Then,

Ysn,t,x=Ysn,t,x=Un(s, Bsn,t,x), 0≤tsT. (13) Let us observe that Zn,t,x is first defined at the pointst=kh, k=nt+ 1, . . . , n. As before we let Zsn,t,x:=Zsn,t,x fors∈]t, T]. We have

Zsn,t,x=Zsn,t,x=∇nUn(s, Bn,t,xs ) fors∈]t, T].

In particular,

Zsn,t,x=∇nUn(s+h, Bsn,t,x) whenever s∈]t, T]\{kh:k=nt+ 1, . . . , n}. (14)

(7)

Of course, we have

Un(t, x) =Ytn,t,x=Ytn,t,x fort∈[0, T].

Similarly, we define, for (t, x)∈[0, T[×R,

n(t, x) :=∇nUn(t+h, x) =Zt+hn,t,x. (15) With this notation, (14) rewrites as

Zsn,t,x= ∆n(s, Bsn,t,x) whenever s∈]t, T]\{kh:k=nt+ 1, . . . , n}. (16) It follows that

Un(t, x) =E

gBTn,t,x+h

n

X

k=nt+1

fkh, B(k−1)hn,t,x , Y(k−1)hn,t,x , Zkhn,t,x

=E

"

gBTn,t,x+ Z T

t

fs, Bsn,t,x, Ysn,t,x, Zsn,t,xds

# ,

which rewrites, taking into account (13) and (16), to Un(t, x) =E

"

gBTn,t,x+ Z T

t

fs, Bsn,t,x, Un(s, Bsn,t,x),∆n(s, Bsn,t,x)ds

#

=E

"

gBTn,t,x+ Z T

t

fs,Θn,t,xs ds

#

, (17)

where

Θn,t,xs :=Bsn,t,x, Un(s, Bsn,t,x),∆n(s, Bsn,t,x). (18) We will prove in Section 5 that (Un,n) converges to (u,∇u).

From now on we assume thatnn0(T,kfkLip) wheren0(T,kfkLip)∈N is the integer given in Lemma12 in the appendix and which automatically implies also existence and uniqueness of solutions because n0 > TkfkLip.

3. Scaled random walk and Wasserstein distance.

One starting point of our paper is the following result of Emmanuel Rio [15] (Theorem 2.1); see also [16]. This result covers, up to a generalization, the case where the generator vanishes, i.e.

f ≡0.

Letψbe the convex function defined by ψ(x) =e|x|−1. The Orlicz norm associated to this functionψ of any real random variable X is given by

kXkψ := inf{a >0 :E[ψ(X/a)]≤1}, inf∅:= +∞.

Let us recall that, for anyr ≥1, sup

x>0

n xr ψ(x)

o

<+∞, kXkLr

sup

x>0

n xr ψ(x)

o1/r

kXkψ. (19) LetX andY be two random variables end let us denote byµthe law ofX and by ν the law of Y. With the usual abuse of notation, the Wasserstein distance associated to ψ is defined by

Wψ(µ, ν) =Wψ(X, Y) := inf{kX−Ykψ : law(X) =µ, law(Y) =ν}.

(8)

Let (Xk)k≥1 be an i.i.d. sequence of random variables withE[X] = 0, EX2= 1,and such that, for some σ > 0, E

heσ|X|i <+∞. Let G be a standard normal random variable. In [15, Theorem 2.1], Emmanuel Rio proved that there exists a constant C >0 such that, forn≥1,

Wψ

n−1/2Sn, GC n−1/2, where Sn=X1+. . .+Xn. As a byproduct, for anyr ≥1, there exists a constant cr>0 such that

Wrn−1/2Sn, Gcrn−1/2, where Wr stands for the Lr-Wasserstein distance

Wr(µ, ν) =Wr(X, Y) := infnE[|X−Y|r]1/r: law(X) =µ, law(Y) =νo. (20) We have also the result of Kantorovich-Rubinstein, i.e.

W1(µ, ν) =W1(X, Y) = sup{E[f(X)]−E[f(Y)] :kfkLip≤1}. (21) Remark 2. We could also consider the case where 0 < r < 1 by using the fact that, in this case, E(|X −Y|r) is a distance (see the arguments in [1, Section 7.1]). In general, we have Wp(µ, ν) =Wq(µ, ν) for 0< p < q <∞.

Let us start with a straightforward generalization of Rio’s result.

Proposition 3. There exists a C >0 such that, for all x∈R and all0≤tsT, WψBsn,t,x, Bt,xs C

T n

1/2

.

As a byproduct, taking into account (19), for anyr≥1, there exists a cr>0 such that, for all x∈R and all 0≤tsT,

Wr

Bsn,t,x, Bst,xcr

T n

1/2

. (22)

Proof of Proposition 3. We have, for anyx∈Rand all 0≤tsT, Wψ(x+BnsBtn, x+BsBt) =Wψ(BnsBtn, BsBt). If s=t, thenBtnBsn= 0,and we have

Wψ(BsnBtn, BsBt) =kBsBtkψ =√

stkGkψ ≤√

hkGkψ. Let us assume thatt < s and let us write

Wψ(BnsBtn, BsBt) =WψBsnBnt, BsBt

WψBsnBtn, BsBt+Wψ BsBt, BsBt. (23) Let us treat each term separately. For the first one, Rio’s result gives

Wψ

√ 1

nsnt

ns

X

k=nt+1

ξk, G

C (nsnt)−1/2,

(9)

and multiplying by√

nsnt

h, we get, sinceph(nsnt)Gis equal toBs−Btin distribution, Wψ BsnBtn, BsBtC

h.

Let us deal with the second term of (23). Let β(s, t) := min(st, st). Then Wψ BsBt, BsBt

=Wψ(N(0, s−t),N(0, s−t))

=Wψ(N(0, β(s, t)),N(0, β(s, t))∗ N(0,|s−t−(s−t)|))

Wψ(0,N(0,|s−t−(s−t)|))

= q

|s−t−(s−t)| kGkψ. But |s−t−(s−t)| ≤h,and this concludes the proof.

Let us finish with a simple consequence of this result that we will use in the sequel.

Corollary 4. Let 0 < ε ≤ 1 and let g : R −→ R be an ε-Hölder continuous function. Then there exists a C >0 depending on T such that, for all x∈R and all0≤tsT,

E

hgBsn,t,xi−E

hgBst,xiCkgkεn−ε/2,

and, setting δ(t, s) := max (st, st),

E

hgBsn,t,x Bsn,t,xxi−E

hgBst,x Bst,xxiCkgkεδ(t, s)1/2n−ε/2.

Proof. Letx∈Rand 0≤tsT. For any coupling (X, Y) ofBsn,t,xand Bst,x, using Hölder’s inequality when 0< ε <1,

W1gBsn,t,x, gBst,x≤E[|g(X)−g(Y)|]≤ kgkεE[|X−Y|ε]≤ kgkεE[|X−Y|]ε. Thus, we have, by (22) for r= 1,

W1gBsn,t,x, gBst,x≤ kgkεW1Bsn,t,x, Bst,xε≤ kgkεcε1 Tnε/2. Choosing f(x) =x in (21), this implies the first result.

Let us prove the second assertion. We start by observing that, sinceBsBt and BsnBtn are centered random variables, we have, setting h(y) := (g(x+y)g(x))y,

E

hgBst,x Bst,xxi=E[(g(x+BsBt)−g(x)) (BsBt)] =E[h(BsBt)], E

h

gBsn,t,x Bn,t,xsxi=E[(g(x+BnsBtn)−g(x)) (BsnBtn)] =E[h(BsnBtn)]. Let us remark that, for any real numbers y and z,|h(y)| ≤ kgkε|y|1+ε, and using the fact that

|y−z|1−ε≤ |y|1−ε+|z|1−ε,

|h(y)−h(z)| ≤ |g(x+y)g(x)| |yz|+|(g(x+y)g(x))−(g(x+z)g(x))| |z|

≤ kgkε(|y−z| |y|ε+|y−z|ε|z|)

≤ kgkε|y−z|ε|y−z|1−ε|y|ε+|y−z|ε|z|

≤ kgkε|y−z|ε|y|+|y|ε|z|1−ε+|z|.

(10)

Young’s inequality,|y|ε|z|1−εε|y|+ (1−ε)|z| ≤ |y|+|z|, leads to

|h(y)−h(z)| ≤2kgkε|y−z|ε(|y|+|z|). (24) In the case where s=t we have

W1(h(BsnBnt), h(BsBt))≤E[|h(BsBt)|]≤ kgkεE

h|BtBs|(1+ε)i

≤ kgkε(s−t)(1+ε)/2E

h|G|(1+ε)i,

where law(G) =N(0,1). Sinces=timplies (s−t)1/2=δ(t, s)1/2 and (s−t)ε/2hε/2,we have W1(h(BsnBtn), h(BsBt))≤ kgkεδ(t, s)1/2(Tn)ε/2

using the fact that E

h|G|(1+ε)i≤1.

Let us turn to the case t < s. For any coupling (X, Y) of BsnBtn andBsBt, using (20) and (24),

W1(h(BsnBtn), h(BsBt))≤E[|h(X)−h(Y)|]

≤2kgkεE[|X−Y|ε(|X|+|Y|)], and, by Hölder’s inequality with p= 2/εand q = 2/(2−ε),

W1(h(BsnBtn), h(BsBt))≤2kgkεE

h|X−Y|2iε/2 E[|X|2/(2−ε)]1−ε/2+E[|Y|2/(2−ε)]1−ε/2

≤2kgkεE

h|X−Y|2iε/2(s−t)1/2+ (s−t)1/2. From (20) it follows that

W1(h(BsnBtn), h(BsBt))≤2kgkεW2(BsnBtn, BsBt)ε (s−t)1/2+ (s−t)1/2

≤4 cε2kgkε T

n ε/2

δ(t, s)1/2,

where we have used (22) forr = 2.

Thus, for 0≤tsT,

W1(h(BsnBnt), h(BsBt))≤Ckgkεn−ε/2δ(t, s)1/2, and the result follows as before by choosing f(x) =x in (21).

4. Regularity results on u, Un, ∇u andn.

Let us start by known regularity properties of the function u that follow from classical a priori estimates for BSDEs.

Lemma 5. Under Assumption(A1) there exists a constantC >0depending on (T, ε, f, g)such that, for all (t, x)∈[0, T]×R,

|u(t, x)| ≤C(1 +|x|)ε, ku(t,·)kεC, ku(·, x)kε/2C(1 +|x|)ε.

(11)

Proof. The first two results follow directly from classical a priori estimates for BSDEs, see e.g.

[8, Proposition 2.1]. The last one ensues from the following upper bound: for any realx and for 0≤rtT,

|u(r, x)−u(t, x)|=|E[Yrr,x]−E[Ytt,x]|

≤ |E[Yrr,xYtr,x]|+|E[Ytr,xYtt,x]|

Z t

r E[|f(s, Bsr,x, Ysr,x, Zsr,x)|]ds+E[|Ytr,xYtt,x|]

C Z t

r E[1 +|Br,xs |ε+|Ysr,x|+|Zsr,x|]ds+E[|Ytr,xYtt,x|].

Since the norm inS2× H2 of (Yr,x, Zr,x) is of order (1 +|x|)ε, we use Cauchy-Schwarz inequality to bound the first term and a priori estimates enable (similarly as in the proof of [8, Proposition 4.1]) to bound the second term.

Next we extend [19, Theorem 3.2] to the case wheref(t,·, y, z) is Hölder continuous.

Lemma 6. Recall the notation (5) and let Assumption (A1) hold.

(a) The function u belongs to C0,1([0, T[×R) and, for all (t, x)∈[0, T[×R, we have,

Zst,x =∇u(s, Bst,x) for a.e.(s, ω)∈[t, T[×Ω, (25) as well as (7) i.e.

∇u(t, x) =E

g(BTt,x)BTBt

Tt

+E Z T

t

F(s, Bst,x)BsBt

st ds

! .

(b) Moreover, there exists a constant C >0 depending on (T, ε, f, g) such that, (i) k∇u(t,·)kεC

T−t for all t∈[0, T[, (ii) |∇u(t, x)| ≤ C

(T−t)(1−ε)/2 for all (t, x)∈[0, T[×R. Consequently, forEr :=E[· |Fr],

∇u(r, Brt,x) =Er

g(BTt,x)BTBr Tr

+Er

Z T r

F(s, Bst,x)BsBr sr ds

!

a.s. for r∈[t, T[.

(26) Proof of Lemma 6. The proof is divided into two steps.

Step 1. We assume in addition thatf is Lipschitz continuous w.r.t. x. Then according to [19], we have only the second point to prove and we know that, for some constant C,

|∇u(t, x)| ≤ C(1 +|x|)

Tt on [0, T[×R. (27)

(bi) The representation (26) yields to k∇u(r, Brt,x)− ∇u(r, Bt,yr )kL2

Er

g(BTt,x)−g(BTt,y)BTBr Tr

L2 +

Z T r

Er

F(s, Bst,x)−F(s, Bst,y)BsBr sr

L2

ds. (28)

(12)

Since g isε-Hölder continuous we get

Er

(g(BTt,x)−g(BTt,y))BTBr

Tr

≤ kgkε|x−y|ε 1

Tr. Similarly, we obtain by the conditional Cauchy-Schwarz inequality the estimate

Er

(F(s, Bst,x)−F(s, Bst,y))BsBr

sr

L2

F(s, Bst,x)−F(s, Bst,y)

L2

√ 1 sr. Using (3) for f, we have

|F(s, Bst,x)−F(s, Bst,y)|

≤ kfxkε|x−y|ε+kfykLip|u(s, Bst,x)−u(s, Bt,ys )|+kfzkLip|∇u(s, Bst,x)− ∇u(s, Bst,y)|, and the Hölder continuity ofu stated in Lemma 5 yields

|F(s, Bst,x)−F(s, Bst,y)|

≤(kfxkε+CkfykLip)|x−y|ε+kfzkLip|∇u(s, Bst,x)− ∇u(s, Bst,y)|. (29) By combining the above estimates we conclude from (28) that

k∇u(r, Brt,x)− ∇u(r, Bt,yr )kL2 ≤ kgkε|x−y|ε

Tr + 2 (kfxkε+CkfykLip)|x−y|εTr

+ Z T

r

∇u(s, Bst,x)− ∇u(s, Bst,y)

L2

kfzkLip

srds

C1|x−y|ε

Tr + Z T

r

∇u(s, Bst,x)− ∇u(s, Bt,ys )

L2

kfzkLip

srds forC1 :=kgkε+ 2T(kfxkε+CkfykLip).Because of (27) we have

∇u(s, Bst,x)− ∇u(s, Bst,y)

L2C2

1 +|x|+|y|

(T−s)1/2

with C2 =C2(C, T)>0. Hence we may apply Gronwall’s lemma (Lemma14) and get k∇u(r, Brt,x)− ∇u(r, Bt,yr )kL2C1c0|x−y|ε

Tr, for somec0 =c0(T,kfzkLip)>0.Especially, for r=tthis implies

|∇u(t, x)− ∇u(t, y)| ≤C|x−y|ε

Tt for someC =C(T, ε, f, g)>0.

(bii) We first notice that for any ε-Hölder continuous function k and for all 0≤t < sT we have

E

k(Bst,x)BsBt

st

= E

(k(Bt,xs )−k(x))BsBt

st

≤ kkkε

(s−t)(1−ε)/2. (30) Therefore, we obtain from (7) that

|∇u(t, x)| ≤ E

g(BTt,x)BTBt Tt

+ E

Z T

t

F(s, Bst,x)BsBt st ds

!

≤ kgkε

(T−t)(1−ε)/2 + Z T

t

kF(s,·)kε (s−t)(1−ε)/2ds.

Références

Documents relatifs

More specifically, studying the renewal in R (see [Car83] and the beginning of sec- tion 2), we see that the rate of convergence depends on a diophantine condition on the measure and

Keywords: Directed polymers, random environment, weak disorder, rate of convergence, regular martingale, central limit theorem for martingales, stable and mixing convergence.. AMS

Keywords : Directed polymers, random environment, weak disorder, rate of conver- gence, regular martingale, central limit theorem for martingales, stable and mixing con- vergence..

We extend the results of [11] about the large time behaviour of the solution (dispersion faster than usual with an additional logarithmic factor, convergence of the rescaled modulus

We extend to Lipschitz continuous functionals either of the true paths or of the Euler scheme with decreasing step of a wide class of Brownian ergodic diffusions, the Central

We prove a rate of convergence for the N-particle approximation of a second-order partial differential equation in the space of probability measures, like the Master equation or

Abstract: We established the rate of convergence in the central limit theorem for stopped sums of a class of martingale difference sequences.. Sur la vitesse de convergence dans le

We established the rate of convergence in the central limit theorem for stopped sums of a class of martingale difference sequences..  2004