• Tidak ada hasil yang ditemukan

getdoc4d61. 294KB Jun 04 2011 12:04:22 AM

N/A
N/A
Protected

Academic year: 2017

Membagikan "getdoc4d61. 294KB Jun 04 2011 12:04:22 AM"

Copied!
29
0
0

Teks penuh

(1)

El e c t ro n ic

Jo ur

n a l o

f P

r o

b a b il i t y

Vol. 14 (2009), Paper no. 94, pages 2691–2719. Journal URL

http://www.math.washington.edu/~ejpecp/

Series Representations of Fractional Gaussian Processes by

Trigonometric and Haar Systems

Antoine Ayache

Werner Linde

Abstract

The aim of the present paper is to investigate series representations of the Riemann–Liouville process,α >1/2, generated by classical orthonormal bases inL2[0, 1]. Those bases are, for

example, the trigonometric or the Haar system. We prove that the representation ofvia the trigonometric system possesses the optimal convergence rate if and only if 1/2< α 2. For the Haar system we have an optimal approximation rate if 1/2< α <3/2 while forα >3/2 a representation via the Haar system is not optimal. Estimates for the rate of convergence of the Haar series are given in the casesα >3/2 andα=3/2. However, in this latter case the question whether or not the series representation is optimal remains open.1.

Key words:Approximation of operators and processes, Rie-mann–Liouville operator, Riemann– Liouville process, Haar system, trigonometric system.

AMS 2000 Subject Classification:Primary 60G15; Secondary: 26A33, 47B06, 41A30. Submitted to EJP on December 07, 2008, final version accepted November 13, 2009.

UMR CNRS 8524, Laboratoire Paul Painlevé, Bat. M2, Université Lille 1, 59655 Villeneuve d’Ascq, France,

Antoine.Ayache@math.univ-lille1.fr

FSU Jena, Institut für Stochastik, Ernst–Abbe–Platz 2, 07743 Jena, Germany,lindew@minet.uni-jena.de

1Recently M. A. Lifshits answered this question (cf.[13]). Using a different approach he could show that in the case

(2)

1

Introduction

Let X = (X(t))tT be a centered Gaussian process over a compact metric space (T,d) possessing a.s. continuous paths. Then it admits a representation

X(t) =

X

k=1

εkψk(t), tT, (1.1)

with(εk)k≥1 i.i.d. standard (real) normal random variables and with continuous real–valued func-tionsψk on T. Moreover, the right hand sum converges a.s. uniformly on T. Since representation (1.1) is not unique, one may ask for optimal ones, i.e. those for which the error

Esup

tT

X

k=n

εkψk(t)

(1.2)

tends to zero, asn→ ∞, of the best possible order. During past years several optimal representations were found, e.g. for the fractional Brownian motion, the Lévy fractional motion or for the Riemann– Liouville process (cf.[7],[2],[5],[8],[1],[18]and[20]). Thereby the representing functionsψk were either constructed by suitable wavelets or by Bessel functions.

In spite of this progress a, to our opinion, natural question remained unanswered. Are the "clas-sical" representations also optimal ? To make this more precise, suppose that the process X has a.s. continuous paths and admits an integral representation

X(t) =

Z

I

K(t,x)dW(x), tT,

for some intervalI Rand with white noiseW onI. Given any ONBΦ = (ϕk)k1in L2(I)we set

ψk(t):=

Z

I

K(t,x)ϕk(x)dx.

By the Itô–Nisio–Theorem (cf.[10], Theorem 2.1.1) the sum in (1.1) converges a.s. uniformly on T, thus leading to a series representation ofX. For example, if I = [0, 1], then one may choose for Φnatural bases as e.g. the ONB of the trigonometric functionsT={1} ∪¦p2 cos(kπ·):k≥1©or that of the Haar functionsH(see (5.2)). There is no evidence that in some interesting cases these "classical" bases do not lead to optimal expansions as well.

The aim of the present paper is to investigate those questions for the Riemann–Liouville process defined by

(t):= 1 Γ(α)

Z t

0

(tx)α−1dW(x), 0≤t≤1 .

This process is known to have a.s. continuous paths wheneverα >1/2 (cf.[15]for further proper-ties of this process). Thus, for example, representation (1.1) ofby the basisTleads to

(t) =ε0 Γ(α+1)+

p 2 Γ(α)

X

k=1 εk

Z t

0

(3)

and in similar way it may be represented by the Haar systemHas

(t) =ε1 Γ(α+1)+

1 Γ(α)

X

j=0 2j1

X

k=0 εj,k

Z t

0

(tx)α−1hj,k(x)dx (1.4)

where thehj,k are the usual Haar functions. The basic question we investigate is whether or not representations (1.3) and (1.4) are optimal. The answer is quite surprising.

Theorem 1.1. If1/2< α≤2, then representation(1.3)is optimal while forα >2it is rearrangement non–optimal, i.e., it is not optimal with respect to any order chosen in T. If 1/2 < α < 3/2, then representation(1.4)is optimal and rearrangement non–optimal forα >3/2.

For the proof we refer to Theorems 4.1, 5.4, 5.5 and 5.11. As recently shown by M. A. Lifshits (oral communication) representation (1.4) is also not optimal for α= 3/2. Let us recall that the assertions forα >2 orα≥3/2, respectively, say that these bases are not only non–optimal in their natural order but also after any rearrangement of the bases.

Even in the cases where these representations are not optimal it might be of interest how fast (or slow) the error in (1.2) tends to zero as n→ ∞. Here we have lower and upper estimates which differ byplogn.

Another process, tightly related to, is the Weyl process which is stationary and 1–periodic. It may be defined, for example, by

(t) =p2 ∞

X

k=1 εk

cos(2kπtαπ/2)

(2πk)α +

p 2

X

l=1

εlsin(2lπtαπ/2)

(2πl)α . (1.5)

Here(εk)and(εl)are two independent sequences of i.i.d. standard (real) normal random variables. We refer to[3]or[14]for more information about this process.

In fact, (1.5) is already a series representation and we shall prove in Theorem 4.8 that it is optimal for allα >1/2. In comparison with Theorem 1.1 this is quite unexpected. Note that the processes and differ by a very smooth process (cf. [3]). Moreover, ifα >1/2 is an integer, then their difference is even a process of finite rank.

2

Approximation by a Fixed Basis

LetH be a separable Hilbert space. ThenG(H,E)denotes the set of those (bounded) operatorsu fromH into a Banach spaceEfor which the sum

X

k=1

εku(ϕk)

converges a.s. in E for one (then for each) ONBΦ = (ϕk)k≥1 in H. As above(εk)k≥1 denotes an i.i.d. sequence of standard (real) normal random variables. Ifu∈ G(H,E), we set

l(u):=E

X

k=1

εku(ϕk)

(4)

which is independent of the special choice of the basis.

Foru∈ G(H,E)the sequence ofl–approximation numbers is then defined as follows:

ln(u):=infl(uuP):Porthogonal projection inH, rk(P)<n .

Note that, of course,l(u) =l1(u)≥l2(u)≥ · · · ≥0 andln(u)→0 as n→ ∞and that

ln(u) =inf

(

E

X

k=n

εku(ϕk)

:Φ = (ϕk)k≥1ONB inH )

.

We refer to[1],[9]or[19]for more information about these numbers.

For our purposes we need to specify the definition of thel–approximation numbers as follows. Let Φ = (ϕk)k≥1 be afixedONB in the Hilbert spaceH. Then we define thel–approximation numbers ofuwith respect toΦby

lΦn(u):=inf

(

E

X

k/N

εku(ϕk)

: N⊆N, #N<n )

.

Let us state some properties oflΦn(u)for later use.

Proposition 2.1. Let u∈ G(H,E)and letΦ = (ϕk)k≥1 be some ONB inH. Then the following are valid.

(1) We have

l(u) =lΦ1(u)l2Φ(u)≥ · · · →0 . (2) If u1,u2∈ G(H,E), then it follows that

lnΦ

1+n2−1(u1+u2)≤l Φ

n1(u1) +l Φ

n2(u2).

(3) If v is a (bounded) operator from E into another Banach space F , then it holds

lnΦ(vu)≤ kvklnΦ(u).

(4) We have

ln(u) =inf¦lnΦ(u):ΦONB inH© .

(5) If E=H, then it holds

lΦn(u)inf

  

X

k/N

u(ϕk)

2

H

!1/2

: N⊆N, #N<n

(5)

Proof. Properties (1), (2) and (4) follow directly from the definition of lΦn(u). Thus we omit their

which completes the proof of (3) by lettingǫ→0.

Property (5) is an easy consequence of

E

for any elements xk in a Hilbert spaceH. Note, moreover, that all moments of a Gaussian vector are equivalent by Fernique’s Theorem (cf.[6]).

Remark 2.1. It is worthwhile to mention that in general rk(u)<ndoes not implylΦn(u) =0. This is in contrast to the properties of the usuall–approximation numbers.

In order to define optimality of a given representation in its natural order we have to introduce a quantity tightly related tolnΦ(u). Foru∈ G(H,E)and an ONBΦ = (ϕk)k1 inH we set

The "o" in the notation indicates that lo,nΦ(u) depends on the order of the elements inΦ while, of course,lnΦ(u)does not depend on it.

Clearly lΦn(u)lno,Φ(u)and, moreover, it is not difficult to show (cf. Prop. 2.1 in[9]) thatlΦn(u) c1nα(logn)β for someα >0 and β∈R implieslo,Φ′

n (u)≤c2nα(logn)β whereΦ′ coincides with Φafter a suitable rearrangement of its elements.

We may now introduce the notion of optimality for a given basis (cf. also[9]and[1]).

Definition 2.1. An ONBΦis said to beoptimalfor u (in the given order ofΦ) provided there is some c>0such that

lno,Φ(u)c ln(u), n=1, 2, . . .

It isrearrangement non–optimalif

lim sup n→∞

lΦn(u)

ln(u) =∞.

(6)

For later purposes we state the following result. Since the proof is quite standard we omit it (cf. also

The following lemma is elementary, but very helpful to prove lower estimates forlΦn(u).

Lemma 2.3. Let u and Φ be as before. Suppose that there exists a subset M ⊆ N of cardinality m

possessing the following property: For someδ >0and some nm each subset LM with#L>mn

Taking on the left hand side of (2.1) the infimum over all subsets N with #N < n proves the assertion.

Let us shortly indicate how the preceding statements are related to the problem of finding optimal series expansions of Gaussian processes. If the processX = (X(t))tT is represented by a sequence

Ψ = (ψk)k≥1 as in (1.1), then we choose an arbitrary separable Hilbert space H and an ONB

Φ = (ϕk)k≥1 in it. By

u(ϕk):=ψk, k≥1 , (2.2)

a unique operatoru∈ G(H,C(T)) is defined. Of course, this construction implies lΦn(u) =lΨn(X) where the latter expression is defined by

(7)

As shown in[1]we haveln(u) =ln(X)where

ln(X):=inf

(

E

X

k=n εkρk

∞ :X =

X

k=1 εkρk

)

.

Hence representation (1.1) is optimal for X, i.e. there is some c > 0 such that for all n 1 we have lno,Ψ(X) c ln(X), if and only if this is so for Φ andu related to Ψ = (ψk)k≥1 via (2.2). In the same wayΦ is rearrangement non–optimal for u if and only if this is so for the representing functionsψk ofX. Consequently, all our results about series expansions may be formulated either in the language of ONB and operatorsu∈ G(H,C(T))or in that of series expansions of centered Gaussian processesX = (X(t))tT with a.s. continuous paths.

3

A General Approach

We start with a quite general result which is in fact the abstract version of Theorem 5.7 in [9]. Before let us recall the definition of the covering numbers of a compact metric space(T,d). Ifǫ >0, then we set

N(T,d,ǫ):=min

n≥1 :∃t1, . . . ,tnTs.t. min

1≤jnd(t,tj)< ǫ,tT

.

With these notation the following is valid.

Proposition 3.1. Let(T,d)be a compact metric space such that

N(T,d,ǫ)γ (3.1)

for someγ >0. Let u∈ G(H,C(T))and letΦ = (ϕk)k≥1 be some fixed ONB inH. Suppose there are β >0andα >1/2such that for all k1and all t,sT we have

(uϕk)(t)−(uϕk)(s)

c d(t,s)β (3.2)

and

u(ϕk)

∞≤c k

α. (3.3)

Then for each n≥1it follows that

E

2n+11 X

k=2n

εku(ϕk)

∞≤c n

1/22n(α−1/2). (3.4)

Proof. Let ǫn be a decreasing sequence of positive numbers which will be specified later on. Set Nn:=N(T,d,ǫn)and note that (3.1) implies

(8)

Then there are t1, . . . ,tN

In order to estimate the second term in (3.6) we need the following result (cf. Lemma 4.14 in[19]).

Lemma 3.2. There is a constant c>0such that for any N ≥1and any centered Gaussian sequence Z1, . . . ,ZN one has,

Summing up, (3.6), (3.7) and (3.9) yield

E δthe first term in (3.10) is of lower order than the second one. This implies

E

(9)

Combining Propositions 2.2 and 3.1 gives the following.

Corollary 3.3. Suppose that(3.1), (3.2) and(3.3) hold for someγ,β >0andα >1/2. Then this implies

lno,Φ(u)c nα+1/2plogn for all n>1.

Remark 3.1. Results more or less similar to Proposition 3.1 and Corollary 3.3 were obtained in

[16; 17](see Proposition 4 in[17]and also the proof of Theorem 1 in[16]).

Let us formulate the preceding result in probabilistic language. We shall do so in a quite general way. LetX = (X(t))tT be an a.s. bounded centered Gaussian process on an arbitrary index set T (we do not suppose that there is a metric onT). Define the Dudley metricdX on T by

dX(t,s):=€E|X(t)X(s)|2Š1/2 , t,sT.

SinceX is a.s. bounded(T,dX)is known to be compact (cf.[6]). We assume that(T,dX)satisfies a certain degree of compactness, i.e., we assume that

N(T,dX,ǫ)≤γ (3.11)

for someγ >0. Suppose now that

X(t) =

X

k=1

εkψk(t), tT, (3.12)

where the right hand sum converges a.s. uniformly onT. Note that theψkare necessarily continu-ous with respect todX. This easily follows from

dX(t,s) =

X

k=1

ψk(t)ψk(s)

2

!1/2

. (3.13)

Then the following holds:

Proposition 3.4. Suppose(3.11)and(3.12)and that there is someα >1/2such that for all k1

sup tT

ψk(t)c kα. (3.14)

Then this implies

Esup

tT

X

k=n

εkψk(t)

c n

α+1/2plogn. (3.15)

(10)

(3.2) is satisfied. Yet this follows by the special choice of the metric on T. Indeed, for h∈ H and

Consequently, (3.2) holds withβ=1 and the assertion follows by (2.2) and by Corollary 3.3.

Remark 3.2. Let us demonstrate on a well–known example how Proposition 3.4 applies. If B = (B(t))0t1 denotes the Brownian Motion on[0, 1], then it admits the representation

Clearly, condition (3.11) holds withγ=2 and the representing functionsψksatisfy

ψk

∞≤c k

1. Consequently, Proposition 3.4 leads to the classical estimate

E sup

and representation (3.16) ofBis optimal. On the other hand, there exist optimal representations of Bwith functionsψkwhere

ψk

∞≈k

1/2(take the representation by the Faber–Schauder system). Thus, in general, neither (3.14) nor (3.3) are necessary for (3.15) or (3.4), respectively.

4

The Trigonometric System

The aim of this section is to investigate whether or not the ONB

T:={1} ∪¦p2 cos(kπ·):k

Equivalently, we may ask whether or not the representation of the Riemann–Liouville process= (Rα(t))0t1given by

is an optimal one.

(11)

Theorem 4.1. The trigonometric system T is optimal for Rα from L2[0, 1] to C[0, 1] in the case 1/2< α≤2. Ifα >2, thenTis rearrangement non–optimal.

Proof. In a first step we prove that T is optimal for Rα provided that 1/2< α ≤ 2. We want to apply Proposition 3.1 withT= [0, 1](here the metric onT is the Euclidean metric),H = L2[0, 1], ϕk(t) =

p

2 cos(kπt)and withu=. Of course, T = [0, 1]satisfies condition (3.1) withγ=1, hence it remains to prove that (3.2) and (3.3) are valid as well.

We claim that (3.2) holds for allα >1/2 withβ=α−1/2 ifα≤3/2 and withβ=1 forα >3/2. Indeed, if 1/2< α≤3/2, then the operatorRα:L2[0, 1]7→C[0, 1]is known to be(α−1/2)–Hölder (cf.[21], vol. II, p. 138), i.e. for allh∈ H andt,s∈[0, 1]we have

(Rαh)(t)(Rαh)(s)ckhk2 |ts|α−1/2 .

Clearly, this implies (3.2) withβ=α−1/2 provided that 1/2< α≤3/2.

Ifα >3/2, then for allhL2[0, 1]we have(Rαh)′(t) = (Rα−1h)(t), hence

(Rαh)(t)(Rαh)(s)=|ts|(Rα1h)(x) (4.2)

for a certain x (t,s). Since(Rα1h)(x) ≤ khk2 Rα1, where the last expression denotes the operator norm of−1:L2[0, 1]7→C[0, 1], by (4.2) we conclude that (3.2) holds withβ=1 in the remaining case.

In order to verify (3.3) we first mention the following general result:

If−1< α <0, then it follows that

sup x>0

Z x

0

sin(xs)ds

<∞and supx>0

Z x

0

cos(xs)ds

<∞. (4.3)

This is a direct consequence of the well–known fact thatR0cos(s)ds as well asR0sin(s)ds exist for thoseα.

The following lemma, which is more or less similar to Lemma 4 in[16], shows that (3.3) holds for α≤2.

Lemma 4.2. Suppose1/2< α≤2and let as beforeϕk(t) =

p

2 cos(kπt). Then it follows that

Rαϕk

∞≤c k

α. (4.4)

Proof. For 1/2 < α ≤ 1 this was proved in [9], Lemma 5.6. The case α = 2 follows by direct calculations. Thus it remains to treat the case 1< α <2. Ifα >1, then integrating by parts gives

(Rαϕk)(t) = 1

Γ(α)

Z t

0

(tx)α−1cos(kπx)dx

= k−1 1

πΓ(α−1)

Z t

0

(tx)α−2sin(kπx)dx

= kα π

α

Γ(α−1)

Z kπt

0

(12)

Moreover, if 1< α <2, then (4.3) and (4.5) imply

Rαϕk

∞=0supt1

(Rαϕk)(t)

c kα

as asserted. Observe that in that case1< α−2<0.

Summing up, andTsatisfy (3.2) and (3.3) of Proposition 3.1 provided that 1/2< α≤2. Thus, we get

E

2n+11 X

k=2n

εkRα(ϕk)

c n1/22−n(α−1/2),

for alln1 and Proposition 2.2 or Corollary 3.3 imply

lno,T(Rα)c nα+1/2plogn.

Recall (cf.[9]) thatln(Rα) nα+1/2plogn, thus the trigonometric systemT is optimal forRα in the case 1/2< α≤2.

To complete the proof of Theorem 4.1 we have to show that the basis T is rearrangement non– optimal for wheneverα >2. To verify this we need the following lemma.

Lemma 4.3. Ifα >2, then there is a c>0such that for k1

Rαϕk

2≥c k

2. (4.6)

Proof. We start with 2< α <3. Using (4.5) we get

(Rαϕk)(t) =cαkα

Z kπt

0

−2 sin(kπts)ds.

Another integration by parts gives

(Rαϕk)(t) =cαkα

¦

(t k)α−2+gk(t)

©

(4.7)

where

gk(t):=(α−2)

Z kπt

0

−3cos(kπts)ds.

Since−1< α−3<0, by (4.3) it follows that

sup k≥1

gk2sup k≥1

gk<∞. (4.8)

Consequently, (4.7) and (4.8) lead to

Rαϕk

2≥c1k

−2

cαkα gk2c1k−2c2kα

which proves our assertion in the case 2< α≤3 whereα=3 follows by direct calculations.

Suppose now 3< α <4. Another integration by parts in the integral defining gk gives gk =cα′ ˜gk with supk1˜gk<∞. Then the above arguments lead to (4.6) in this case as well.

(13)

Now we are in position to complete the proof of Theorem 4.1. This is done by using Lemma 2.3. In the notation of this lemma we setM :={1, . . . , 2n}for some given n≥1, hence we have m=2n. Let L be an arbitrary subset in M with #L> mn. Using that all moments of Gaussian sums are equivalent it follows that

E

where the last estimate follows by Lemma 4.3. Because of #L>nandLM we have

X

logn, thus (4.9) shows thatTis rearrangement non–optimal forα >2.

Remark 4.1. The optimality of the trigonometric system for 1/2 < α ≤ 2 is implicitly known. Indeed, as shown in[17], Proposition 4, estimate (4.4) implies optimality in the Riemann–Liouville setting.

Corollary 4.4. Ifα >2, then are constants c1,c2>0only depending onαsuch that

c1n−3/2lnT(Rα)c2n−3/2plogn. (4.10)

Proof. The left hand estimate in (4.10) was proved in (4.9). In order to verify the right hand one we use property (3) of Proposition 2.1. Ifα >2 this implies

lTn(Rα)≤

−2:C[0, 1]7→C[0, 1]

lnT(R2)c n−3/2plogn.

This completes the proof.

Remark 4.2. We conjecture that forα >2 the right hand side of (4.10) is the correct asymptotic of lnT(Rα).

Our next objective is to investigate the ONB

(14)

Theorem 4.5. Let˜Tbe the ONB defined above. ThenT˜is optimal for Rα provided that1/2< α≤1. Ifα >1it is rearrangement non–optimal.

Proof. We start with the case 1/2< α≤1. By using the same method as in the proof of Lemma 5.6

where as beforeϕkT˜are ordered in the natural way. Condition (3.2) holds by the same arguments as in the proof of Theorem 4.1. Thus Corollary 3.3 applies and proves that ˜Tis optimal.

To treat the caseα >1 we need the following lemma.

Lemma 4.6. Ifα >1, then it follows that

Proof. Suppose first 1< α <2 and write

(Rαsin(2·))(t) = 1 On the other hand,

completing the proof of the lemma in that case.

(15)

where we used (4.11) in the last step. The set LM was arbitrary with #L>n, hence, by Lemma 2.3 we obtain

lTn˜(Rα)cn−1/2. In view of ln(Rα) ≈ nα+1/2

p

logn this implies that ˜T is rearrangement non–optimal whenever α >1 completing the proof.

Finally we investigate series representations of the Weyl process. Recall the definition of the Weyl operator of fractional integration. Forα >1/2, the Weyl operator is given on exponential func-tions for all t [0, 1]by

Iα(e2πik·)(t):= e 2πikt

(2πik)α, k ∈Z\ {0},

where forα /∈N, the denominator has to be understood as

(2πik)α:=|2πk|−α·exp(πα

2 i sgn(k)).

By linearity and continuity, the definition of can be extended to the complex Hilbert space

L20[0, 1]:=

(

f L2[0, 1]:

Z 1

0

f(x)d x =0

)

,

thus, is a well–defined operator from L02[0, 1]into C[0, 1]. Note, that it maps real valued func-tions onto real ones.

Proposition 4.7. For allα >1/2it holds

ln(Iα)nα+1/2plogn.

Proof. Leten(u)denote then–th (dyadic) entropy number of an operatorufromH into a Banach space E (cf. [4]for more information about these numbers). As proved in [11], Proposition 2.1, whenever an operatoru∈ G(H,E)satisfies

en(u)c1na(logn)β for somea>1/2 andβ∈R, then this implies

ln(u)≤c2na+1/2(logn)β+1. Moreover, as shown in[3], for anyα >1/2 it follows that

en(Rα)≤c1e−c2n 1/3

.

In particular, we have en(RαIα) cγnγ for any γ > 0. Thus, by the above implication, for anyγ >0 holds ln(Rα)≤n

γ as well. Of course, since l

2n−1(Iα)≤ln() +ln(Rα)by

(16)

Remark 4.3. The upper estimate ln(Iα)≤c nα+1/2

p

lognmay also be derived from Theorem 4.8 below.

Before proceeding further, let us choose a suitable ONB in the real L02[0, 1]. We take T′:=T˜\ {1}=¦p2 cos(2·), p2 sin(2·):k,l ≥1©

and order it in the natural way. Then it holds

Theorem 4.8. For anyα >1/2the basisTis optimal for Iα.

Proof. Direct calculations give

Iα(cos(2·))(t) = cos(2kπtαπ/2) (2πk)α as well as

(sin(2·))(t) =

sin(2lπtαπ/2)

(2πl)α .

Consequently, condition (3.3) in Proposition 3.4 holds for andT′. We claim now that for 1/2< α <3/2 it follows that

(Iαh)(t)(Iαh)(s)c khk2|ts|α−1/2 (4.13) for allhL02[0, 1]and t,s ∈[0, 1]. This is probably well–known, yet since it is easy to prove we shortly verify (4.13). It is a direct consequence of

X

k=1

1−e2πikǫ2

k2α =

X

k≤1

1−e2πikǫ2 k2α +

X

k>1

1−e2πikǫ2 k2αc1ǫ2 X

k≤1

k−2α+2+4 X k>1

k−2α−2α+1

for anyk≥1 andǫ >0 small enough. Clearly, (4.13) shows that (3.2) holds withβ =α−1/2 as long as 1/2< α <3/2. The identity1◦2=1+α2 implies the following: Suppose that

(Iαh)(t)(Iαh)(s)ckhk2 |ts|β (4.14) for someα >1/2 andβ∈(0, 1]. Then for anyα′≥α estimate (4.14) is also valid (with the same β). This, for example, follows from the fact that (4.14) is equivalent to

IαδtIαδs2c|ts|β

together with the semi–group property of , hence also of ∗. Here ∗ denotes the dual operator of Iα, mapping C[0, 1]∗ into L20[0, 1], and δt denotes the Dirac point measure concentrated at t∈[0, 1].

Summing up, we see that Iα and T′ satisfy (3.2) and (3.3). Clearly, (3.1) holds for T = [0, 1]. Thus Corollary 3.3 applies and completes the proof. Recall that by Proposition 4.7 we haveln(Iα)≈

nα+1/2plogn.

Remark 4.4. It may be a little bit surprising that for allα >1/2 the basisT′is optimal forIα while ˜

(17)

5

Haar Basis

5.1

Some useful notations and some preliminary results

Recall (cf. (1.4)) that for any parameterα >1/2, the Riemann-Liouville process can be written as

(t) =ε1

whereis the Riemann-Liouville operator and thehj,k’s are the usual Haar functions, i.e.

hj,k(x) =2j/2

and where the series converges almost surely uniformly in t (i.e. in the sense of the normk · k∞). For anyt[0, 1]andJ Nwe set,

In order to conveniently express the coefficients(Rαhj,k)(t), for any realsν andx we set

(x)ν+=

Let us now give some useful lemmas. The following lemma can be proved similarly to Lemma 1 in

[2], this is why we omit its proof.

Lemma 5.1. Let{εj,k : j∈N0, 0≤k≤2j−1}be a sequence of standard Gaussian variables. Then there exists a random variable C1 >0of finite moment of any order such that one has almost surely, for every j∈N0and0k<2j1,

|εj,k| ≤C1

p

(18)

Let us now give a lemma that allows to control the increments of the Riemann-Liouville process. This result is probably known however we will give its proof for the sake of completeness.

Lemma 5.2.

(i) For anyα ∈(1/2, 3/2), there is a random variable C2 > 0of finite moment of any order such that almost surely for every t1,t2[0, 1],

|(t1)(t2)| ≤C2|t1−t2|α−1/2

p

log(2+|t1−t2|−1). (5.8) (ii) For anyα >3/2, there is a random variable C3>0of finite moment of any order such that one

has almost surely for every t1,t2∈[0, 1],

|(t1)(t2)| ≤C3|t1t2|.

Proof. (of Lemma 5.2) Part (ii) is a straightforward consequence of the fact that the trajectories of are continuously differentiable functions when α > 3/2. Let us now prove part (i). First observe (see for instance relation (7.6) in[12]) that inequality (5.8) is satisfied when the Riemann-Liouville process is replaced by the fractional Brownian motion (fBm)(BH(t))0t1with Hurst index H:=α−1/2. Recall that for somecα>0 (againH andαare related viaH=α−1/2)

BH(t) =(t) +cα(t), (5.9)

where the process(Qα(t))t[0,1]is called the low-frequency part of fBm and up to a positive constant it is defined as

(t) =

Z 0

−∞

n

(tx)α−1(x)α−1odW(x).

Finally, it is well-known that the trajectories of the process(Qα(t))t[0,1] areC∞-functions.

There-fore, it follows from (5.9) that(Rα(t))0t1satisfies (5.8) as well.

Remark 5.1. It is very likely that (5.8) also holds forα=3/2. But in this case our approach does not apply. Observe thatα=3/2 corresponds toH=1 and (5.9) is no longer valid.

For any realsγ >0 and t∈[0, 1]and for any integers j∈N0and 0k≤2j1 we set

,j,k(t):=

t2k+2 2j+1

γ

+−2

t2k+1 2j+1

γ

++

t 2k 2j+1

γ

+. (5.10)

Observe that (5.10) and (5.7) imply that

(Rαhj,k)(t) = 2j/2

Γ(α+1),j,k(t). (5.11)

Furthermore, we denote byekj(t)the unique integer satisfying the following property:

ekj(t)

2jt<

ekj(t) +1

2j , (5.12)

with the convention thatekj(1) =2j1.

(19)

Lemma 5.3.

(i) For each k≥ekj(t) +1, one has Aγ,j,k(t) =0.

(ii) There is a constant c4>0, only depending onγ, such that the inequality

|Aγ,j,k(t)| ≤c42−1+ekj(t)−2 (5.13)

holds when0kekj(t).

(iii) Ifγ6=1, then there is a constant c5>0, only depending onγ, such that the inequality

|Aγ,j,k(t)| ≥c52−1+ekj(t)−2 (5.14)

holds when0kekj(t)2.

Proof. (of Lemma 5.3) Part(i)is a straightforward consequence of (5.6), (5.10) and (5.12), so we will focus on parts (ii) and(iii). Inequality (5.13) clearly holds whenγ= 1, this is why we will assume in all the sequel thatγ6=1. Let us first show that (5.13) is satisfied when

ekj(t)1kekj(t). (5.15)

Putting together (5.12) and (5.15) one has for anyl ∈ {0, 1, 2},

t2k+l 2j+1

γ

+≤

tekj(t)−1 2j

γ

≤2−(j−1)γ.

Therefore, it follows from (5.10) and (5.15) that

|,j,k(t)| ≤22−(j−1)γc62−

1+ekj(t)k

γ−2 ,

where the constantc6=22+γmax{1, 22−γ}. Let us now show that the inequalities (5.13) and (5.14) are verified when

0kekj(t)2. (5.16)

We denote by ,j,t the function defined for every real xk/2j as

fγ,j,t(x) =tx2−tx2−j−1γ.

By applying the Mean Value Theorem to ,j,t on the interval [2k2j−+11, k

2j], it follows that there exists a1(2k2j−+11,

k

2j)such that

Aγ,j,k(t) = fγ,j,t k 2j

fγ,j,t2k−1 2j+1

=2−j−1fγ,j,t(a1)

= γ2−j−1

ta12−−1ta12−j−1γ−1

(20)

Next, by applying the Mean Value Theorem to the function

y 7→ta1−1

on the interval[2−j−1, 2−j], it follows that there existsa2(2−j−1, 2−j)such that

ta12−−1ta12−j−1γ−1

=(γ−1)2−j−1ta1a2γ−2. (5.18)

Observe that the inequalities (5.12), 2kj <a1+a2< k2+j1 and (5.16) imply that 1

2j

ekj(t)k−1

2j <ta1−a2 <

ekj(t)k−1+2

2j

≤ 3(ekj(t)−k−1)

2j . (5.19)

Next settingc4 = c6+γ|γ−1|max{1, 3γ−2} andc5 =γ|γ−1|min{1, 3γ−2} and combining (5.17) with (5.18) and (5.19), it follows that the inequalities (5.13) and (5.14) are verified when (5.16) holds.

5.2

Optimality when

1

/

2

< α <

1

The goal of this section is to prove the following theorem.

Theorem 5.4. Suppose1/2< α <1. Then there is a random variable C7>0of finite moments of any order such that one has almost surely, for every JN,

kJk∞C72−(α−1/2)Jp1+J.

In particular, this implies that in this case representation (5.1) possesses the optimal approximation rate.

Proof. (of Theorem 5.4) Putting together (5.1), (5.3), Lemma 5.1, (5.11) and (5.13), one obtains that almost surely, for every t[0, 1], and every integerJN,

|(t)J(t)| ≤

X

j=J 2Xj−1

k=0

|εj,k||(Rαhj,k)(t)|

C1Γ(α+1)−1c4

X

j=J

2−j(α−1/2)pj+1

ekj(t) X

k=0

1+ekj(t)k

α−2

C72−J(α−1/2)

p

J+1.

Observe that the condition 1/2α <1 plays a crucial role in the proof of Theorem 5.4. Indeed,

one hasPekj(t) k=0

(21)

5.3

Optimality when

1

< α <

3

/

2

The goal of this subsection is to show that the following theorem holds.

Theorem 5.5. Suppose1< α <3/2. Then there is a constant c8>0such that for every J∈None has

EkRαRα

Jk∞≤c82−J(α−1/2)

p

J+1.

In particular, this implies that also in this case representation(5.1)possesses the optimal approximation rate.

First we need to prove some preliminary results.

Proposition 5.6. If1/2< α <3/2, there exists a constant c9>0such that one has for any JN,

e

σ2Jc292−J(2α−1).

Proof. (of Proposition 5.6) It follows from (5.4), (5.11) and parts(i)and(ii)of Lemma 5.3, that

e

σ2J(t)c42Γ(α+1)−2

X

j=J

2−j(2α−1)

ekj(t) X

k=0

1+ekj(t)−k

−2(2−α)

c922−J(2α−1),

where the constantc92=c42Γ(α+1)−2(12−(2α−1))P∞

l=1l−2

(2−α)<

∞.

Lemma 5.7. For anyα ∈(1, 3/2), there exists a random variable C10 >0 of finite moment of any order such that one has almost surely for any real t[0, 1]and any integer JN,

J(t)JekJ(t)2−JC102−J(α−1/2)pJ+1.

We refer to(5.12)for the definition of the integerekJ(t).

In order to be able to prove Lemma 5.7 we need the following lemma.

Lemma 5.8. For any realα >1, there exists a constant c11> 0such that for all t [0, 1], J N,

jN0 and kN0satisfying

0 jJ and0k≤ekj(t),

one has

(Rαhj,k)(t)−(Rαhj,k)

ekJ(t)2−Jc112(3/2−α)jJ1+ekj(t)−3. (5.20)

Proof. (Proof of Lemma 5.8) It is clear that (5.20) holds whent=ekJ(t)2−J, so we will assume that t 6=ekJ(t)2−J. By applying the Mean Value Theorem, it follows that there existsa (ekJ(t)2−J,t) such that

(Rαhj,k)(t)−(Rαhj,k)

ekJ(t)2−J (5.21)

= α2

j/2−J Γ(α)

a2k+2 2j+1

α−1

+ −2

a2k+1 2j+1

α−1

+ +

a 2k 2j+1

α−1

+ .

Observe that one hasekj(a) =ekj(t) since a ekJ2(Jt),t

⊂ ekj(t) 2j ,

ekj(t)+1 2j

(22)

We are now in position to prove Lemma 5.7.

Proof. (of Lemma 5.7) By using Lemma 5.1 and the fact that

0ekJ(t)2−α2−J,

On the other hand, it follows from Lemma 5.8 that

J−1

Lemma 5.9. There is a random variable C13 > 0 of finite moments of any order such that one has almost surely for every JN

sup

Using the triangular inequality and Lemmas 5.2(i)and 5.7, it follows that

(23)

Proof. (of Theorem 5.5) Putting together Lemma 5.9, Lemma 3.2, the fact that σe2J ≥ sup0K<2J,KN

0σe 2 J(2−

JK)and Proposition 5.6 one obtains the theorem.

5.4

The case

α

=

3

/

2

Recall (cf.[9]) thatln(R3/2)≈n−1

p

logn; this clearly implies that

lnH(R3/2)cn−1plogn.

The goal of this subsection is to show that a slightly stronger result holds, namely the following theorem.

Theorem 5.10. There exists a constant c14 >0such that, for any n ∈Nand any set N ⊆ {(j,k)

N2 : 0k2j1}satisfying#N<n one has

2−Jn 2Jn1

X

l=0

X

(j,k)∈/N

(R3/2hj,k)(l/2Jn)

21/2

c14n−1

p

logn,

where Jn2is the unique integer such that2Jn−2n<2Jn−1.

Remark 5.2. A straightforward consequence of (5.4), (5.5) and Theorem 5.10 is that for allJ∈N0

one has

e

σJc142−J

p

J+1.

In fact by using the same technics as before one can prove that 2−JpJ+1 is the right order of

e

σJ i.e. one has for some constant c15 > 0 and all J ∈N0, σeJc152−J

p

J+1. But observe that

e

σJ ≈2−J

p

J+1 does, unfortunately, not answer the question whether or not representation (5.1) is optimal in the caseα=3/2.

Proof. (of Theorem 5.10) Let us set

M :=nkN : 0k2Jn−11 and(Jn,k)/N

o

.

Clearly,

#M ≥2Jn#N>2Jn−1 (5.24) and

2−Jn 2Jn1

X

l=0

X

(j,k)∈/N

(R3/2hj,k)(l/2Jn)

2

≥2−Jn 2Jn1

X

l=0

X

kM

(R3/2hJn,k)(l/2 Jn)

2

. (5.25)

(24)

thatekJ n(l/2

Jn) =l for any integerl satisfying 0l 2Jn1 and (5.24), it follows that

2−Jn 2XJn−1

l=0

X

kM

(R3/2hJn,k)(l/2Jn)

2

c25Γ(α+1)−22−3Jn

X

kM 2XJn−1

l=k+2

lk+1

−1

c25Γ(α+1)−22−3Jn 2XJn−3

k=2Jn−11 2JnXk−2

p=1

(p+2)−1

c25Γ(α+1)−22−3Jn 2XJn−2

n=1

2Jn−1n

(n+2)−1

c25Γ(α+1)−22−2Jn−2 2Jn−2

X

n=1

(n+2)−1

c25Γ(α+1)−22−2Jn−2

Z 2Jn−2+1

1

(x+2)−1d x

c214n−2logn, (5.26)

with the convention thatP2l=Jnk+12. . .=P2pJn=1k−2=· · ·=0 whenever we have 2Jn2k2Jn1. Finally combining (5.25) with (5.26) we obtain the theorem

5.5

Non–optimality of the Haar basis for

α >

3

/

2

The goal of this subsection is to prove the following theorem.

Theorem 5.11. Ifα >3/2, then we have:

(i) For any t0 ∈ (0, 1] there exists a constant c16 > 0 such that, for each n ∈ N and each set N⊆ {(j,k)N2 : 0k2j1}, satisfying#N<n one has

X

(j,k)∈/N

(Rαhj,k)(t0)

21/2

c16n−1. (5.27)

(ii) There exists a constant c17>0such that for each JNone has

EkRαRα

Jk∞≤c172−J

p

J+1.

A straightforward consequence of Theorem 5.11 is that the Haar basis H is rearrangement non– optimal for whenα >3/2. More precisely:

Corollary 5.12. Ifα >3/2, then there are two constant0<cc, only depending onα, such that for each n2one has

(25)

Remark 5.3. We conjecture that

It is clear that

X

Moreover, it follows from (5.29), (5.28) and (5.12) that

#L≥ekJn(t0)−1−#N ≥ekJn(t0)−2 Jn−3t

0>3·2Jn−2t0. (5.31) On the other hand (5.11) and (5.14) imply that

(26)

where c19> 0 is a constant only depending on t0 andα. Finally, putting together, (5.30), (5.32), (5.33), (5.34) and (5.28) one obtains (5.27).

In order to be able to prove part(ii)of Theorem 5.11 we need some preliminary results.

Proposition 5.13. Ifα >3/2, there exists a constant c17>0such that one has for any J∈N,

e

σ2Jc172 2−2J.

Proof. (of Proposition 5.13) It follows from (5.4), (5.11) and parts(i)and(ii)of Lemma 5.3, that

e

σ2J(t)c42Γ(α+1)−2

X

j=J

2−j(2α−1)

ekXj(t)

k=0

1+ekj(t)k−2(2−α) (5.35)

Let us assume that 3/2< α <2; then the fact that

x 7→1+ekJn(t0)−k

2(α−2)

is an increasing function on[0,ekJ

n(t0) +1)implies

ekj(t) X

k=0

1+ekj(t)−k

−2(2−α)

Z ekj(t)+1

0

1+ekj(t)−x

−2(2−α)

d x

c2022α−3, (5.36)

wherec20>0 is a constant only depending onα. Next let us assume thatα≥2; then the fact that x 7→1+ekJn(t0)−k

2(α−2)

is an nonincreasing function on[1,ekJn(t0)]entails that

ekj(t) X

k=0

1+ekj(t)−k

−2(2−α)

Zekj(t)

−1

1+ekj(t)−x

−2(2−α)

d x

c2122α−3, (5.37)

wherec21>0 is a constant only depending onα. Finally putting together (5.35), (5.36) and (5.37) one obtains the proposition.

Lemma 5.14. For anyα >3/2, there exists a random variable C22>0of finite moment of any order such that one has almost surely for any real t∈[0, 1]and any integer J∈N,

J(t)JekJ(t)2−JC222−JpJ+1.

We refer to(5.12)for the definition of the integerekJ(t).

Proof. (of Lemma 5.14) By using Lemma 5.1 and the fact that

(27)

one gets that

J(t)J

2−JekJ(t)≤ |ε−1| α2−J Γ(α+1)

+ C1pJ+1 J−1

X

j=0 2j1

X

k=0

(Rαhj,k)(t)(Rαhj,k)ekJ(t)2−J. (5.38)

On the other hand, it follows from Lemma 5.8 that

J−1

X

j=0 2j1

X

k=0

(Rαhj,k)(t)(Rαhj,k)ekJ(t)2−J

c11 J−1

X

j=0

2(3/2−α)jJ

ekXj(t)

k=0

1+ekj(t)−3. (5.39)

Let us assume that 3/2< α <2; then one has

J−1

X

j=0

2(3/2−α)jJ

ekXj(t)

k=0

1+ekj(t)−3c232−J, (5.40)

where c23=c11

1−23/2−α

−1P l=1l

α−3<. Next let us assume thatα2. By using the same technics as in the proof of Proposition 5.13 one can show that there is a constant c24 > 0, only depending onα, such that for each jN,

ekXj(t)

k=0

1+ekj(t)−3c242j(α−2)+j+1.

One has therefore,

J−1

X

j=0

2(3/2−α)jJ

ekj(t) X

k=0

1+ekj(t)−k

α−3

c252−J, (5.41)

wherec25=c24P∞j=02(3/2−α)j2j(α−2)+j+1<∞. Finally putting together (5.38), (5.39), (5.40) and (5.41) one obtains the lemma.

Lemma 5.15. There is a random variable C26 >0of finite moments of any order such that one has almost surely for every J∈N

sup t∈[0,1]

(t)J(t)

≤ sup

0≤K<2J,KN 0

(K2−J)J(K2−J)

+C262−JpJ+1.

(28)

We are now in position to prove part(ii)of Theorem 5.11.

Proof. (of Theorem 5.11 part (ii)) Putting together Lemma 5.15, Lemma 3.2, the fact that σe2J ≥ sup0K<2J,KN

0σe 2

J(2−JK)and Proposition 5.13 one obtains the theorem.

Acknowledgments

The authors are very grateful to Helga Wendel (formerly Helga Schack) for several helpful discus-sions about the subject of this paper.

References

[1] A. Ayache and W. Linde, Approximation of Gaussian random fields: general results and optimal wavelet representation of the Lévy fractional motion. J. Theoret. Probab. 21 (2008), 69–96. MR2384473

[2] A. Ayache and M. S. Taqqu,Rate optimality of wavelet series approximations of fractional Brow-nian motion.J. Fourier Anal. Appl.9(2003), 451–471. MR2027888

[3] E. Belinsky and W. Linde, Small ball probabilities of fractional Brownian sheets via fractional integration operators.J. Theoret. Probab.15(2002), 589–612. MR1922439

[4] B. Carl and I. Stephani,Entropy, Compactness and Approximation of Operators.Cambridge Univ. Press, Cambridge, 1990. MR1098497

[5] K. Dzhaparidze and H. van Zanten, Optimality of an explicit series expansion of the fractional Brownian sheet.Statist. Probab. Lett.71(2005), 295–301. MR2145497

[6] X. Fernique,Fonctions aléatoires gaussiennes, vecteurs aléatoires gaus-siens.Montréal: Les Pub-lications CRM, 1997. MR1472975

[7] H. Gilsing and T. Sottinen, Power series expansions for fractional Brownian motions. Theory Stoch. Process.9(2003), 38–49. MR2306058

[8] E. Iglói, A rate–optimal trigonometric series expansion of the fractional Brownian motion. Elec-tron. J. Probab.10(2005), 1381–1397. MR2183006

[9] Th. Kühn and W. Linde,Optimal series representation of fractional Brownian sheets.Bernoulli8 (2002), 669–696. MR1935652

[10] S. Kwapie´n and W. A. Woyczy´nski,Random Series and Stochastic Integrals: Single and Multiple. Birkhäuser, Boston, Basel, Berlin, 1992. MR1167198

[11] W. V. Li and W. Linde, Approximation, metric entropy and small ball estimates for Gaussian measures.Ann. Probab.27(1999), 1556–1578. MR1733160

[12] W. V. Li and Q. M. Shao, Gaussian processes: inequalities, small ball probabilities and applica-tions.Stochastic Processes: Theory and Methods, Handbook of Statistics.19(2001), 533–597. MR1861734

(29)

[14] M. A. Lifshits, W. Linde and Z. Shi,Small deviations for Riemann-Liouville processes in Lq-spaces with respect to fractal measures.Proc. London Math. Soc.92(2006), 224–250. MR2192391

[15] M. Lifshits and T. Simon.Small deviations for fractional stable processes.Ann. Inst. H. Poincaré Probab. Statist.41(2005), 725–752. MR2144231

[16] H. Luschgy and G. Pages. High–resolution product quantization for Gaussian processes under sup–norm distortion.Bernoulli13(2007), 653–671. MR2348745

[17] H. Luschgy and G. Pages. Expansions for Gaussian processes and Parseval frames. Electron. J. Probab.14(2009), 1198–1221. MR2511282

[18] A. Malyarenko, An optimal series expansion of the multiparameter fractional Brownian motion. J. Theoret. Probab.21(2008), 459–475. MR2391256

[19] G. Pisier, The Volume of Convex Bodies and Banach Space Geometry. Cambridge Univ. Press, Cambridge, 1989. MR1036275

[20] H. Schack, An optimal wavelet series expansion of the Riemann–Liouville process. To appear J. Theoret. Probab.

Referensi

Dokumen terkait

Paket pengadaan ini terbuka untuk penyedia barang/jasa yang teregistrasi pada Layanan Pengadaan Secara Elektronik dan memenuhi persyaratan sebagaimana tercantum dalam

effeminate terlihat mulai dari bagaimana ibu- ibu rumah tangga muda tersebut dalam melakukan praktek menonton program- program musik yang dipandu para pre- senter effeminate

terima pekerjaan Jenis, kapasitas, komposisi dan jumlah peralatan minimal yang disediakan Spesifikasi teknis dan brosur yang ditawarkan Tenaga Teknis/terampil sesuai dengan

cenderung imaging yang tidak sepenuhnya menggambarkan realitas bisa nampak seperti sebuah kebenaran yang tak terbantahkan, 2) media memiliki kekuatan untuk menganggap biasa

Sehubungan dengan hal tersebut di atas, kepada peserta yang berkeberatan atas pengumuman hasil Pelelangan Sederhana dengan Pascakualifikasi kegiatan Penyelenggaraan

Hipotesis didefinisikan sebagai dugaan atas jawaban sementara mengenai sesuatu masalah yang masih perlu diuji secara empiris, untuk mengetahui apakah pernyataan

Ia mendefinisikan ruang publik ( public sphere ) sebagai area dalam kehidupan sosial, sebuah tempat dimana orang akan berkumpul dan mengidentifikasi masalah mereka secara sosial

Pengadaan Pemerintah Kota Banjarbaru, telah dilaksanakan Pembukaan Penawaran. Pelelangan Umum dengan Pascakualifikasi Pekerjaan Pembangunan