• Tidak ada hasil yang ditemukan

getdoc6455. 196KB Jun 04 2011 12:04:26 AM

N/A
N/A
Protected

Academic year: 2017

Membagikan "getdoc6455. 196KB Jun 04 2011 12:04:26 AM"

Copied!
16
0
0

Teks penuh

(1)

ELECTRONIC COMMUNICATIONS in PROBABILITY

WEAK APPROXIMATION OF FRACTIONAL SDES: THE DONSKER

SETTING

XAVIER BARDINA1

Departament de Matemàtiques, Facultat de Ciències, Edifici C, Universitat Autònoma de Barcelona, 08193 Bellaterra, Spain

email:

❳❛✈✐❡r✳❇❛r❞✐♥❛❅✉❛❜✳❝❛t

CARLES ROVIRA2

Facultat de Matemàtiques, Universitat de Barcelona, Gran Via 585, 08007 Barcelona, Spain email:

❝❛r❧❡s✳r♦✈✐r❛❅✉❜✳❡❞✉

SAMY TINDEL

Institut Élie Cartan Nancy, B.P. 239, 54506 VandIJuvre-lès-Nancy Cedex, France email:

t✐♥❞❡❧❅✐❡❝♥✳✉✲♥❛♥❝②✳❢r

SubmittedNovember 17, 2009, accepted in final formJune 6, 2010 AMS 2000 Subject classification: 60H10, 60H05

Keywords: Weak approximation, Kac-Stroock type approximation, fractional Brownian motion, rough paths

Abstract

In this note, we take up the study of weak convergence for stochastic differential equations driven by a (Liouville) fractional Brownian motionB with Hurst parameterH (1/3, 1/2), initiated in [3]. In the current paper, we approximate thed-dimensional fBm by the convolution of a rescaled random walk with Liouville’s kernel. We then show that the corresponding differential equation converges in law to a fractional SDE driven byB.

1

Introduction

The current article can be seen as a companion paper to [3], to which we refer for a further introduction. Indeed, in the latter reference, the following equation on the interval [0, 1] was considered (the generalization to[0,T]being a matter of trivial considerations):

d yt=σ yt

d Bt+b yt

d t, y0=a∈Rn, (1)

whereσ:RnRn×d,b:RnRnare two bounded and smooth enough functions, andBstands for ad-dimensional fBm with Hurst parameterH>1/3.

1RESEARCH SUPPORTED BY THE MINISTERIO DE CIENCIA E INNOVACIÓN GRANT MTM2009-08869 2RESEARCH SUPPORTED BY THE MINISTERIO DE CIENCIA E INNOVACIÓN GRANT MTM2009-07203

(2)

Let us be more specific about the driving process for equation (1): we consider in the sequel the so-calledd-dimensional Liouville fBmB, with Hurst parameterH(1/3, 1/2). Namely,Bcan be written asB= (B1, . . . ,Bd), where theBi’s aredindependent centered Gaussian processes of the

form

Bit= Zt

0

(tr)H−12dWi

r, (2)

for a d-dimensional Wiener process W = (W1, . . . ,Wd). This process is very close to the usual fBm, in the sense that they only differ by a finite variation process (as pointed out in[1]), and we shall see that its simple expression (2) simplifies some of the computations throughout the paper. In any case, B falls into the scope of application of the rough paths theory, which means that equation (1) can be solved thanks to the semi-pathwise techniques contained in[9, 10, 13]. The natural question raised in[3]was then the following: is it possible to approximate equations like (1) in law by ordinary differential equations, thanks to a Wong-Zakai type approximation (see [12, 16, 17]for further references on the topic)?

Some positive answer to this question had already been given in[8], where some Gaussian se-quences approximations were considered in a general context. In [3], we focused on a natural and easily implementable (non Gaussian) scheme for B, based on Kac-Stroock’s approximation to white noise (see [11, 15]). However, another very natural way to approximate B relies on Donsker’s type scheme (see[14]for the caseH>1/2 and[4]for the Brownian case), involving a rescaled random walk. We have thus decided to investigate weak approximations to (1) based on this process.

More precisely, as an approximating sequence ofB, we shall choose()ǫ>0, where,iis defined as follows fori=1, . . . ,d: consider a family of independent random variables{ηi

k;k≥1, 1≤i

d}, satisfying the

Hypothesis 1.1. The random variables{ηik;k1, 1id}are independent and share the same law as another random variableη. Furthermore,ηis assumed to satisfy E η

=0,E€η2Š=1and is almost surely bounded by a constant kη.

We then define,iin the following way:

Xi,ǫ(t) = Z t

0

(t+ǫ2r)H−12θi,ǫ(r)d r, (3) where

θi,ǫ(r):=1

ǫ

+∞

X

k=1

ηi kI[k−1,k)

 r

ǫ2 ‹

. (4)

Notice that is really a process given by the convolution of the rescaled random walkθǫ with Liouville’s kernel.

Let us then consider the process solution to equation (1) driven by, namely:

d yǫt =σ€ytǫŠd Xtǫ+b€tŠd t, y0ǫ=aRn, t[0,T]. (5) Our main result is as follows:

Theorem 1.2. Let()

ǫ>0be the family of processes defined by (5), and let1/3< γ <H, where H

(3)

Let us make a few comments about this theorem:

(i)We have presented our results for the Liouville fBmBbecause of some simplifications, apparent in[3], in the manipulations of some stochastic integrals. However, as mentioned in[1], the usual fractional Brownian motion Bˆ can be decomposed as Bˆ = B+V, where V is a finite variation process andBis the Liouville fBm. By writing the Lévy area ofBˆaccording to this decomposition, it is certainly possible to extend our results to fBm, which is in a sense a more natural process for its invariance properties. We did not explore this possibility for sake of conciseness.

(ii)We also have chosen the range(1/3, 1/2)for the coefficientH. Indeed, the caseH>1/2 can be deduced easily from[5]since in this case, the solution yto (1) is a continuous function ofB. The Brownian caseH =1/2 is proved easily following the methods in[6]. We could also have dealt with a coefficient H lying in(1/4, 1/3] according to[8], but this case is much harder for two reasons: (1) The approximation of third order integrals is also required in this case. (2) The assumption H >1/3 is needed for the convergence of certain terms in [3, Proof of Proposition 5.1](see our termB2,1,1there).

(iii)Notice that ift[2,(N+1)ǫ2)one has

Xi,ǫ(t) = 1 (H+1/2)ǫ

– N

X

k=1

ηi k

”

(t−(k−2)ǫ2)H+1/2(t(k1)ǫ2)H+1/

+ηi N+1

”

(t−(N−1)ǫ2)H+1/2ǫ2H+1— ™

.

Another possibility in order to approximate Xi would have been to define a process Xˆi,ǫ at any point of the form2by

ˆ

Xi,ǫ(2) = 1 (H+1/2)ǫ−2H

N X

k=1

ηik”(N(k2))H+1/2(N(k1))H+1/2—,

and then computeXˆi,ǫ at any other point by linear interpolation. This new approximation might be closer to the spirit of Donsker type results, since the process Xˆi,ǫ is piecewise linear. In any case, it should be easy (though technical) to show that Xˆi,ǫXi,ǫ vanishes whenǫ 0, as L2

-random variables taking values in an appropriate rough path space (including the Lévy area of both processes). Thus Theorem 1.2 certainly holds true whenXi,ǫis replaced byXˆi,ǫ. Here again, we did not explore this possibility for sake of conciseness.

Here is how our paper is structured: as the reader shall see, many of the techniques introduced in[3]are also useful in our context. In the end, as explained at Section 2, most of the technical differences between the two articles arise in the way to evaluate the moments of quantities like

R1 0 f(r)θ

i,ǫ(r)d rfor a given Hölder function f, and to compare them with the moments of Gaus-sian random variable. This is thus where we shall concentrate our efforts, and this essential point will be handled at Section 3.

2

Reduction of the problem

(4)

First of all, we need to recall the definition of some Hölder spaces, in which our convergences take place. We call for instance Cj([0, 1];Rd) the space of continuous functions from[0, 1]j toRd, which will mainly be considered for j=1 or 2 variables. The Hölder norms on those spaces are then defined in the following way: forf ∈ C2([0, 1];Rd)let

kfkµ= sup

s,t∈[0,T] |fst|

|ts|µ, and C µ

2([0, 1];R d) =¦

f ∈ C2([0, 1];Rd);kfkµ<©.

The usual Hölder spacesC1µ([0, 1];Rd)are then determined by settingkgkµ=kδgkµfor a contin-uous functiong∈ C1([0, 1];Rd), whereδg∈ C2([0, 1];Rd)is defined byδg

st=gtgs. We then

say thatg∈ C1µ([0, 1];Rd)iffkgkµis finite. Note thatk · kµis only a semi-norm onC1([0, 1];Rd), but we will work in general on spaces of the type

C1,aµ ([0, 1];R d) =¦

g:[0,T]V;g0=a,kgkµ<

©

, (6)

for a givenaV, on whichkgkµ is a norm.

The second crucial point one has to recall is the natural definition of a Lévy area for Liouville’s fBm. To this purpose, considerE the set of step-functions on[0,T]with values inRd. LetH be the Hilbert spaceH defined as the closure ofE with respect to the scalar product induced by

¬

(1[0,t1], . . . ,1[0,td]),(1[0,s1], . . . ,1[0,sd]) ¶

H =

d X

i=1

R(ti,si), si,ti∈[0,T], i=1, . . . ,d,

whereR(t,s):=E[Bi

tBsi]. Then a natural representation of the inner product inH is given via the

operatorK, defined fromE toL2([0,T]), by:

Kϕ(t) = (Tt)H−12ϕ(t)−

1 2−H

Z T

t

[ϕ(r)ϕ(t)](rt)H−32d r,

and it can be checked thatK can be extended as an isometry betweenH and the Hilbert space L2([0,T];Rd). Thus the inner product inH can be defined as:

ϕ,ψ

H ¬

Kϕ,Kψ

L2([0,T];Rd).

The mapping(1[0,t1], . . . ,1[0,td])7→ Pd

i=1B i

ti can also be extended into an isometry between H

and the first Gaussian chaosH1(B)associated withB= (B1, . . . ,Bd). We denote this isometry by

ϕ 7→ B(ϕ), and B(ϕ)is called the Wiener-Itô integral of ϕ. It is shown in [7, page 284] that

C1γ(Rd)⊂ H wheneverγ >1/2−H, which allows to defineB(ϕ)for such kind of functions.

Proposition 2.1. Let B be a d-dimensional Liouville fBm, and suppose that its Hurst parameter satisfies H(1/3, 1/2). Then

(1)B is almost surely aγ-Hölder path for any1/3< γ <H. (2)A Lévy area based on B can be defined by setting

B2st= Z t

s

d Bu⊗ Zu

s

d Bv, i. e. B2st(i,j) = Z t

s

d Bui

Zu

s

(5)

for0s<tT . Here, the stochastic integrals are defined as Wiener-Itô integrals when i6= j, while, when i= j, they are simply given by

Z t

s

d Bui

Zu

s

d Biv=1

2

€

BitBsiŠ2.

(3)The processB2is almost surely an element ofC22γ([0, 1];Rd×d), and satisfies the algebraic rela-tion

B2

stB

2

suB

2

ut= BuBs

BtBu

, for all0sut1.

These algebraic and analytic properties of the fBm path allow to invoke the rough path machinery (see[9, 10, 13]) in order to solve equation (1):

Theorem 2.2. Let B be a Liouville fBm with Hurst parameter1/3<H<1/2, andσ:RnRn×d be a C2function, which is bounded together with its derivatives. Then

(1)Equation (1) admits a unique solution y ∈ C1γ(Rn)for any1/3< γ <H, with the additional structure of weakly controlled process introduced in[10].

(2)The mapping(a,B,B2)7→ y is continuous fromRn× Cγ

1(R

d)× C2γ

2 (R

d×d)toCγ

1(R n).

One of the nice aspects of rough paths theory is precisely the second point in Theorem 2.2, which allows to reduce immediately our weak convergence result for equation (1), namely Theorem 1.2, to the following result on the approximation of(B,B2):

Theorem 2.3. Recall that the random variablesηi

ksatisfy Hypothesis 1.1, and let X

ǫbe defined by (3). For anyǫ >0, let X2,ǫ = (X2st,ǫ(i,j))s,t≥0;i,j=1,...,d be the natural Lévy’s area associated to Xǫ,

given by

X2,ǫ

st (i,j) = Z t

s

(Xuj,ǫXsj,ǫ)d Xui,ǫ, (7) where the integral is understood in the usual Lebesgue-Stieltjes sense. Then, asǫ0,

(,X2,ǫ) −→Law (B,B2), (8) whereB2denotes the Lévy area defined in Proposition 2.1, and where the convergence in law holds in the spacesC1µ(Rd)× C2µ

2 (Rd×d), for anyµ <H.

The remainder of our work is thus devoted to the proof of Theorem 2.3.

As usual in the context of weak convergence of stochastic processes, we divide the proof into weak convergence for finite-dimensional distributions and a tightness type result. Furthermore, the tightness result in our case is easily deduced from the analogous result in[3]:

Proposition 2.4. The sequence(,X2,ǫ)ǫ>0defined at Theorem 2.3 is tight inC1µ(R

d)×C2µ

2 (R d×d).

(6)

With these preliminaries in hand, we can now turn to the finite dimensional distribution (f.d.d. in the sequel) convergence, which can be stated as:

Proposition 2.5. Under the assumption 1.1, let(,X2,ǫ)be the approximation process defined by (3) and (7). Then

f.d.d.−lim ǫ→0(X

ǫ

,X2,ǫ) = (B,B2), (9)

wheref.d.d.limstands for the convergence in law of the finite dimensional distributions. Otherwise stated, for any k1and any family{si,ti;ik, 0si<tiT}, we have

L −lim

ǫ→0(X

ǫ

t1,X 2,ǫ

s1t1, . . . ,X ǫ

tk,X

2,ǫ

sktk) = (Bt1,B 2

s1t1, . . . ,Btk,B

2

sktk). (10)

Proof. The structure of the proof follows again closely the steps of[3, Proposition 5.1], except that other kind of estimates will be needed in order to handle the Donsker case.

To be more specific, it should be observed that the first series of simplifications in the proof of[3, Proposition 5.1]can be repeated here. They allow to pass from a convergence of double iterated integrals to the convergence of some Wiener type integrals with respect to. Namely, fori=1, 2 and 0u<t1, set

Yi(u,t) = Z t

u

(BivBui)(vu)H−32d v,

and for 0u<t1 and(u1, . . . ,u6)in a neighborhood of 0 inR6, set also

Zu = u1+u2Bu2+u3Y2(u,t) +u4

Z t

u

(vu)H−12dW2

v

+u5 Zt

u

d w

Zw

u

(wv)H−3

2 (wu)H− 1

2−(vu) 1 2dW2

v

+u6 Zt

u

d w

Zu

0

(wv)H−32(wu)H− 1 2dW2

v.

Consider the analogous processesYi,ǫ,Zǫdefined by the same formulae, except that they are based on the approximationsθi,ǫ of white noise. We still need to recall a little more notation from[3]: for f L2([0, 1])andt[0, 1], we set

Φǫ(f) =E

ei

Rt

0f(u)θ

ǫ,1(u)du

, φǫf = Z1

0 Z1

0

f2(x)f2(y)I{|xy|2

}d x d y, (11) and

Φ(f) =E

ei

Rt

0f(u)dW 1

u

=e12

Rt

0f 2(u)du

.

Then it is shown in[3, Proposition 5.1]that one is reduced to prove that limǫ0vǫa=0, wherevǫa is given by

vǫa=E

Φǫ()eiw

Rt

0θ

ǫ,2(u)du

E

Φ()eiw

Rt

0θ

ǫ,2(u)du

, for an arbitrary real parameterwin a neighborhood of 0. Furthermore, boundingeiw

Rt

0θ

ǫ,2(u)du triv-ially by 1 and conditioning, it is easily shown thatva

(7)

Φ()|], for which Lemma 3.3 provides the bound

for anyα(0, 1). In order to reach our aim, it is thus sufficient to check the following inequalities: sup

However, these relations can be deduced, as (39), (40) and (41) in[3], from Lemma 3.1 (it should be noticed however that a one-parameter version of[2, Lemma 5.1]is needed for the adaptation of the latter result to our Donsker setting). The proof is thus finished once the lemmas below are proven.

3

Moments estimates in the Donsker setting

In order to deal with our technical estimates, let us first introduce a new notation: set ρ1 = (1−51/2)/2 andρ2= (1+51/2)/2. Then the moments of any integral of a deterministic kernel f with respect toθi,ǫ can be bounded as follows:

Lemma 3.1. Let m∈N, f L2([0, 1]), i∈ {1, 2}andǫ >0. Recall that the random variableηis assumed to be almost surely bounded by a constant kη. Then we have

f is the quantity defined at (11).

(8)

Step 1: Identification of some key iterated integrals.Notice that

Transforming the symmetric integral on[0, 1]2minto an integral on the simplex, and using

expres-sion (4) forθǫ, we can write the latter expression as:

(2m)!

Let us observe at this point that we have split our sum into T1

m and Tm2 because Tm1 represents

the dominant contribution to our moment estimate. This is simply due to the fact that T1 m is

obtained by assuming some pairwise equalities among the random variablesηi

k, whileT 2

(9)

on a higher number of constraints. In any case, both expressions will be analyzed through the introduction of some iterated integrals of the form

(k;v,w) = 1

ǫν

Z

[(k−1)ǫ2,kǫ2]ν

ν

Y

j=1

|f(rj)|I{wr1>···>rνv}d r1· · ·d rν,

defined forν,k1 and 0v<w1.

Step 2: Analysis of the integrals Kν. Those iterated integrals are treated in a slightly different way according to the parity ofν. Indeed, forν=2n, thanks to the elementary inequality 2a ba2+b2,

we obtain a bound of the form:

n(ǫ) X

k=1

K2n(k;v,w) (17)

n(ǫ) X

k=1

1

ǫ2n Z

[(k−1)ǫ2,kǫ2]2n n Y

i=1 ‚

f2(x

2i−1) +f2(x2i)

2

Œ

I{wx1≥···≥x2nv}d x1· · ·d x2n

n(ǫ) X

k=1

1

ǫ2n Z

[(k−1)ǫ2,kǫ2]2n

f2(x1)· · ·f2(xn)I{wx1≥···≥xnv}d x1· · ·d x2n

= n(ǫ) X

k=1 Z

[(k−1)ǫ2,kǫ2]n

f2(x1)· · ·f2(xn)I{wx1≥···≥xnv}d x1· · ·d xn

≤ Z

[0,1]n

f2(x1)· · ·f2(xn)I{x1−xn<ǫ2}I{wx1≥···≥xnv}d x1· · ·d xn.

The caseν=2n+1 can be treated along the same lines, except for the fact that one has to cope with some expressions of the form

n(ǫ) X

k=1

K3(k;v,w) n(ǫ) X

k=1

1

ǫ

Z

[(k−1)ǫ2,kǫ2]2

|f(x1)|f2(x2)I{wx1≥x2≥v}d x1d x2

≤ 1

ǫ

Z

[0,1]2

|f(x1)|f2(x2)I{x1−x22

}I{wx1≥x2≥v}d x1d x2. (18)

Combining (18) and (17) we can state the following general formula: let ν 1, and define a couple(ν,νˆ)as: (i)ν=ν/2,νˆ=0 ifνis even, (ii)ν= (ν+1)/2,νˆ=1 ifνis odd. With this notation in hand, we have:

n(ǫ) X

k=1

(k;v,w)≤ 1

ǫνˆ Z

[0,1]ν

|f(x1)|2−νˆf2(x2)· · ·f2(∗)I{x1−2}

×I{wx1≥···≥∗≥v}d x1· · ·d xν∗. (19)

Step 3: Bound on T1

(10)

K2(k;w,v), for which one can apply (19). This yields

procedure in order to control this term. Namely, integrating with respect to the lastnsvariables,

one obtains that Plugging our bound (19) onKnsinto this expression, we get

1

In the end, sincePns=2m, the remaining singularity inǫis of the form Q

ǫ−ˆns. However, each

of the singularityǫ−ˆns comes with an integral that compensates the singularityǫnˆs (recall that ˆ

ns≤1). Hence, iterating the integrations with respect to the variablesr, we end up with a bound

of the form

is easily obtained. On the other hand, we will illustrate with an example how the bound can be obtained when all the terms are odd. Let us consider the casem=n1=n2=3. Following our

(11)

Step 5: Bound on T2

m. Owing to inequality (20), our bound on T 2

m can be reduced now to an

estimate of the number of terms in the sum overn1, . . . ,nsin formula (15). This boils down to the following question: given a natural numbern, how can we write it as a sum of natural numbers (larger than one)?

This is arguably a classical problem, and in order to recall its answer, let us take a simple example: forn=6, the possible decompositions can be written as{6; 2+2+2; 2+4; 4+2; 3+3}. Further-more, notice that the decompositions of 6 can be obtained by adding+2 to the decompositions of 4 or adding 1 to the last number of the decompositions of 5. Extrapolating to a general integern, it is easily seen that the number of decompositions can be expressed asun−1, where(un)n≥1stands

for the Fibonacci sequence. We have thus found a number of decompositions of the form Nn=5−1/ρ2n−1ρ1n−1Š,

where the quantitiesρ1,ρ2appear in formula (12). Moreover, the number of terms inT2is given

byN2m−1, the−1 part corresponding to the termT1.

Putting together this expression with (20) and the result of Step 3, our claim (12) is now easily obtained.

Step 6: Proof of (13).The proof of (13) follows the same arguments as for (12). We briefly sketch the main difference between these two proofs, lying in the analysis of the term Un1,...,ns. Indeed,

since we are now dealing with an odd power 2m+1, the equivalent of (20) is an upper bound of the form

1

ǫ

Z1

0 Z1

0

|f(y1)|f2(y2)I{|y1−y2|2}d y1d y2

Z

[0,1]m−1

m−1 Y

j=1

f2(xj)I{x1≥···≥xm−1}d x1· · ·d xm1. (21)

Furthermore, applying Hölder’s inequality twice, we obtain 1

ǫ

Z1

0 Z1

0

|f(y1)|f2(y2)I{|y1−y2|2}d y1d y2= 1

ǫ

Z1

0

f2(y2)

Z1∧(y2+ǫ2)

0∨(y2−ǫ2)

|f(y1)|d y1d y2

≤ Z 1

0 Z1

0

f2(y1)f2(y2)I{|y1−y2|2

}d y1d y2 !1

2

kfkL2= (φǫ

f)

1 2kfkL2,

and thus we can bound (21) by ( 1

m−1)!(φ

ǫ

f)

1

2kfk2m−1

L2 , which ends the proof.

Our next technical lemma compares the moments of a Wiener type integral with respect toθǫand with respect to the white noise.

Lemma 3.2. Let m∈N, f Cα([0, 1]), i∈ {1, 2},ǫ >0and for m≥1, set

Jm=

1

(2m)!E

 

Z1

0

f(r)θi,ǫ(r)d r

!2m

−

1 2mm!

Z

[0,1]m

f2(s1)· · ·f2(sm)ds1· · ·dsm

.

Then

(1) We have J1ǫ2α

(12)

(2) For any m>1, the following inequality holds true, where we recall thatρ1,ρ2have been defined just before Lemma 3.1:

Jm

Proof. We divide again this proof into several steps.

Step 1: Variance estimates.We prove here the first of our assertions: Notice that 1

On the other hand 1

and hence this quantity can be bounded as follows:

J1 ≤ which is the first claim of our lemma.

(13)

Step 3: Study ofT˜1

m: We analyze ˜T 1

min a slightly different way as in Lemma 3.1. Namely, we first

write another expression of the same type, calledTˆm1and defined by

ˆ

satisfyingkl=kl+1. However, this latter term can be bounded as in (20), yielding

|T˜m3| ≤m−1

Step 4: Conclusion. Putting together the decompositions we have obtained so far, we end up with

(14)

Invoking our estimates (22) and (24) on ˜T2 mand ˜T

3

m, our bound onJmeasily reduced to check that

1 2mm!ǫ2m

Z

[0,1]2m

f(r1)· · ·f(r2m)−f2(r1)f2(r3)· · ·f2(r2m

−1)

×I{|r1−r2|2}· · ·I{|r

2m−1−r2m|2}d r1· · ·d r2m≤

1

(m1)!ǫ

2αkfk

αkfk2mL2−1.

The latter inequality can now be obtained from the decomposition

f(r

1)· · ·f(r2m)−f2(r1)f2(r3)· · ·f2(r2m−1)

=f(r1)(f(r2)−f(r1))f(r3)f(r4)· · ·f(r2m) +f2(r1)f(r3)(f(r4)f(r3))f(r5)f(r6)· · ·f(r2m) +· · ·

+f2(r1)f2(r3)...f2(r2m−3)f(r2m−1)(f(r2m)−f(r2m−1)) ,

the inequalities

1 2ǫ2

Z1

0 Z1

0

f2(r1)I|r1−r2|2d r2d r1 ≤ kfk2

L2, 1

2ǫ2 Z1

0 Z1

0

f(r1)f(r2)I|r1−r2|2d r2d r1 ≤ kfk2

L2,

and from the estimate we have already obtained forJ1. This finishes the proof.

Finally, the characteristic function of a Wiener type integral of the formR1

0 f(r)θ

k,ǫ(r)d rcan be compared to its expected limitR1

0 f(r)dW k

r in the following way:

Lemma 3.3. Let f ∈ Cα([0, 1])for a certainα(0, 1), k∈ {1, . . . ,d}andǫ >0. For any uR, we have:

E

eiu

R1 0f(r)θ

k,ǫ(r)d rE

eiu

R1 0f(r)dW

k

r

≤4(1/5)1/2u3(φǫ

f)

1

2kfkL2k3ηexp(4u2kη2kfk2

L2) +u2ǫ2 α

kfkαkfkL2exp(u2kfk2

L2)

+8 5−1/2u4(φǫf)12kfk2

L2k

2

ηexp(4u

2k2

ηkfk

2

L2) + (1/2)u

4φǫ

fexp(u 2

kfk2L2).

(15)

fact that the odd moments of a Gaussian random variable are null, we get

In order to control the real part of the difference, we will use Lemma 3.2. This yields:

The latter quantity can be bounded by

u2ǫ2αkfkαkfkL2exp(u2kfk2 which ends the proof.

References

[1] E. Alòs, O. Mazet and D. Nualart (2000):Stochastic calculus with respect to fractional Brown-ian motion with Hurst parameter lesser than 1/2.Stoch. Proc. Appl.86, 121-139. MR1741199 [2] X. Bardina, M. Jolis and L. Quer-Sardanyons (2009):Weak convergence for the stochastic heat

equation driven by Gaussian white noiseArxiv Preprint.

(16)

[4] E. Breuillard, P. Friz and M. Huesmann (2009): From random walks to rough paths.Proc. Amer. Math. Soc.137(10), 3487-3496. MR2515418

[5] R. Delgado and M. Jolis (2000): Weak approximation for a class of Gaussian processes. J. Appl. Probab.37(2), 400-407. MR1780999

[6] S.N. Ethier and T.G. Kurtz (1986):Markov processes. Characterization and convergence.Wiley Series in Probability and Mathematical Statistics: Probability and Mathematical Statistics. John Wiley & Sons, Inc., New York. MR0838085

[7] D. Feyel and A.de La Pradelle (1999): On fractional Brownian processes.Potential Anal.10, 273-288. MR1696137

[8] P. Friz and N. Victoir (2009): Differential equations driven by Gaussian signals (I).Ann. Inst. H. Poincaré Probab. Statist., to appear.

[9] P. Friz and N. Victoir (2010): Multidimensional stochastic processes as rough paths: theory and applications.Cambridge Studies in Advanced Mathematics, 120. Cambridge University Press, Cambridge. MR2604669

[10] M. Gubinelli (2004):Controlling rough paths.J. Funct. Anal.216, 86-140. MR2091358 [11] M. Kac (1974): A stochastic model related to the telegraphers equation. Rocky Moutain J.

Math.4497–509, reprinting of an article published in 1956. MR0510166

[12] T. G. Kurtz and P.E. Protter (1996): Weak convergence of stochastic integrals and differen-tial equations.Probabilistic models for nonlinear partial differential equations (Montecatini Terme, 1995), Lecture Notes in Math.1627, Springer, 1-41. MR1431298

[13] T. Lyons and Z. Qian (2002): System control and rough paths. Oxford University Press. MR2036784

[14] T. Sottinen (2001): Fractional Brownian motion, random walks and binary market models. Finance Stochast.5, 343-355. MR1849425

[15] D. Stroock (1982):Topics in Stochastic Differential Equations. Tata Institute of Fundamental Research/Springer Verlag, Bombay, 1982. MR0685758

[16] G. Tessitore and J. Zabczyk (2006):Wong-Zakai approximations of stochastic evolution equa-tions.J. Evol. Equ.6, no. 4, 621-655. MR2267702

Referensi

Dokumen terkait

[r]

Pada perancangan sistem ini akan dibuat suatu program yang dapat dijalankan pada PC maupun laptop, yang digunakan untuk simulasi pencarian rute terpendek, rute yang

Untuk menerima inputan dengan perintah ci n , maka operator yang akan digunakan adalah overloaded operator (&gt;&gt;) dan diikuti dengan variabel tempat menyimpan?.

Bagi Peserta yang keberatan dengan hasil Penetapan Pemenang di atas dapat mengajukan sanggahan secara tertulis kepada Pokja Pengadaan Pekerjaan Jasa Konstruksi

menggunakan logika kampanye yang sama, yakni, logika menjegal, “kalau tidak memilih, maka mereka akan menang.” Di sini, elit politik tinggal menata sedikit logika publik untuk

Beberapa contoh kecurangan misalnya dengan memasukkan transaksi-transaksi fiktif atau yang telah diubah ke dalam suatu sistem : menghilangkan transaksi yang sah;

Suatu sistem terdiri dari sejumlah komponen yang saling berinteraksi yangD. artinya saling bekerja sama dalam membentuk satu

Dalam kerangka konseptual ini, penulis mengaplikasikan teori yang digunakan sebagai landasan penelitian dengan keadaan yang ada di lapangan tentang Strategi komunikasi