This article appeared in a journal published by Elsevier. The attached copy is furnished to the author for internal non-commercial research and education use, including for instruction at the authors institution and sharing with colleagues. Other uses, including reproduction and distribution, or selling or licensing copies, or posting to personal, institutional or third party websites are prohibited. In most cases authors are permitted to post their version of the article (e.g. in Word or Tex form) to their personal website or institutional repository. Authors requiring further information regarding Elsevier’s archiving and manuscript policies are encouraged to visit: http://www.elsevier.com/copyright

Author's personal copy

Stochastic Processes and their Applications 121 (2011) 793–812 www.elsevier.com/locate/spa

Quantitative Breuer–Major theorems Ivan Nourdin a , Giovanni Peccati b , Mark Podolskij c,∗ a Laboratoire de Probabilit´es et Mod`eles Al´eatoires, Universit´e Pierre et Marie Curie, Boˆıte courrier 188, 4 Place

Jussieu, 75252 Paris Cedex 5, France

b Facult´e des Sciences, de la Technologie et de la Communication; UR en Math´ematiques. 6, rue Richard

Coudenhove-Kalergi, L-1359, Luxembourg

c Department of Mathematics, ETH Z¨urich, HG G32.2, 8092 Z¨urich, Switzerland

Received 12 July 2010; received in revised form 3 December 2010; accepted 15 December 2010 Available online 22 December 2010

Abstract � We consider sequences of random variables of the type Sn = n −1/2 nk=1 { f (X k ) − E[ f (X k )]}, n ≥ 1, where X = (X k )k∈Z is a d-dimensional Gaussian process and f : Rd → R is a measurable function. It is known that, under certain conditions on f and the covariance function r of X , Sn converges in distribution to a normal variable S. In the present paper we derive several explicit upper bounds for quantities of the type |E[h(Sn )] − E[h(S)]|, where h is a sufficiently smooth test function. Our methods are based on Malliavin calculus, on interpolation techniques and on the Stein’s method for normal approximation. The bounds deduced in our paper depend only on Var[ f (X 1 )] and on simple infinite series involving the components of r . In particular, our results generalize and refine some classic CLTs given by Breuer and Major, Giraitis and Surgailis, and Arcones, concerning the normal approximation of partial sums associated with Gaussiansubordinated time series. c 2010 Elsevier B.V. All rights reserved. � MSC: 60F05; 60H05; 60G15; 60H07 Keywords: Berry–Esseen bounds; Breuer–Major central limit theorems; Gaussian processes; Interpolation; Malliavin calculus; Stein’s method

∗ Corresponding author.

E-mail addresses: [email protected] (I. Nourdin), [email protected] (G. Peccati), [email protected], [email protected] (M. Podolskij). c 2010 Elsevier B.V. All rights reserved. 0304-4149/$ - see front matter � doi:10.1016/j.spa.2010.12.006

Author's personal copy 794

I. Nourdin et al. / Stochastic Processes and their Applications 121 (2011) 793–812

1. Introduction Fix d ≥ 1 and consider a d-dimensional centered stationary Gaussian process X = (X k )k∈Z , (1) (d) X k = (X k , . . . , X k ), defined on a probability space (Ω , F, P). For any 1 ≤ i, l ≤ d and j ∈ Z, we denote by (i)

(l)

r (i,l) ( j) = E[X 1 X 1+ j ],

(1.1)

the covariance of X 1 and X 1+ j . Let f : Rd → R be a measurable function and write (i)

(l)

n 1 � Sn = √ { f (X k ) − E[ f (X k )]}, n k=1

n ≥ 1,

(1.2)

to indicate the sequence of normalized partial sums associated with the subordinated process k �→ f (X k ). One crucial problem in Gaussian analysis is the following: Problem P. Find conditions on f and on the covariance r in order to have that, as n → ∞, Sn converges in distribution to a Gaussian random variable. Although easily stated, Problem P is in fact quite subtle. For instance, as observed e.g. in [8, p. 429], it is in general not possible to deduce a solution to Problem P by using standard central limit results for dependent random variables (for instance, by applying techniques based on mixing). More to the point, slight variations in the form of f and r may imply that either the normalization by the factor n 1/2 is inappropriate, or the limiting distribution is not Gaussian (or both apply): see Dobrushin and Major [15], Rosenblatt [31] and Taqqu [36,37] for several classic results connected with this phenomenon, as well as Breton and Nourdin [7] for recent developments. It turns out that an elegant solution to Problem P can be deduced by using the notion of Hermite rank. Recall that the function f is said to have Hermite rank equal to q with respect to X , where q ≥ 1 is an integer, if (a) E[( f (X ) − E[ f (X )]) pm (X )] = 0 for every polynomial pm (on Rd ) of degree m ≤ q − 1; and (b) there exists a polynomial pq of degree q such that E[( f (X ) − E[ f (X )]) pq (X )] �= 0. Then, one has the following well-known statement: Theorem 1.1 (Breuer–Major Theorem for Stationary Vectors). Let E[ f 2 (X 1 )] < ∞, and assume that the function f has Hermite rank equal to q ≥ 1. Suppose that � |r (i,l) ( j)|q < ∞, ∀i, l ∈ {1, . . . , d}. (1.3) j∈Z

Then σ 2 := Var[ f 2 (X 1 )] + 2 [0, ∞). Moreover, one has that d

Sn −→ S ∼ N (0, σ 2 ),

�∞

k=1 Cov[ f (X 1 ),

f (X 1+k )] is well-defined, and belongs to (1.4) d

where N (0, σ 2 ) indicates a centered Gaussian distribution with variance σ 2 , and −→ stands for convergence in distribution. In the case d = 1, Theorem 1.1 was first proved by Breuer and Major in [8], whereas Theorem 4 in Arcones [2] proves the statement for a general d (both proofs in [2,8] are based on the method of cumulants and diagram formulae — see e.g. [29,34]). The reader is referred

Author's personal copy I. Nourdin et al. / Stochastic Processes and their Applications 121 (2011) 793–812

795

to Sun [33] for an early statement in the case of a Hermite rank equal to 2, and to Giraitis and Surgailis [17] for some continuous-time analogues of Theorem 1.1. Note that any central limit result involving Hermite ranks and series of covariance coefficients is customarily called a ‘Breuer–Major theorem’, in honor of the seminal paper [8]. Theorem 1.1 and its variations have served as fundamental tools for Gaussian approximations in an impressive number of applications, of which we provide a representative (recent) sample: renormalization of fractional diffusions [1], power variations of Gaussian and Gaussian-related continuous-time processes [3,4,14,20,24], Gaussian fluctuations of heat-type equations [5], estimation of Hurst parameters of fractional processes [9,12,13], unit-root problems in econometrics [10], empirical processes of long-memory time series [22], level functionals of stationary Gaussian fields [19], variations of multifractal random walks [21], and stochastic programming [38]. See also Surgailis [34] for a survey of some earlier uses of Breuer–Major criteria. Despite this variety of applications, until recently the only available techniques for proving results such as Theorem 1.1 were those based on combinatorial cumulants/diagrams computations. These techniques are quite effective and flexible (see e.g. [23,32] for further instances of their applicability), but suffer from a fundamental drawback, namely they do not allow us to deduce Berry–Esseen relations of the type � � �E[h(Sn )] − E[h(S)]� ≤ ϕ(n), n ≥ 1, (1.5)

where h is a suitable test function, and ϕ(n) → 0 as n → ∞. An upper bound such as (1.5) quantifies the error that one makes when replacing Sn with S for a fixed n. In [25, Section 4], the first two authors of this paper proved that one can combine Malliavin calculus (see e.g. [28]) and Stein’s method (see e.g. [11]) to obtain relations such as (1.5) (for some explicit ϕ(n)) in the case where: (i) d = 1, (ii) f = Hq is a Hermite polynomial of degree q ≥ 2 (and thus has Hermite rank equal to q), (iii) X is obtained from the increments of a fractional Brownian motion of Hurst index H < 1 − (2q)−1 , and (iv) h is either an indicator of a Borel set or a Lipschitz function. Since under (iii) one has that |r ( j)| ∼ j 2H −2 , these findings allow us to recover a very special case of Theorem 1.1 (see Example 2.5 below for more details on this point). The aim of the present work is to extend the techniques initiated in [25] in order to deduce several complete quantitative Breuer–Major theorems, that is, statements providing explicit upper bounds such as (1.5) for any choice of f and r satisfying the assumptions of Theorem 1.1. We stress now that we will not require that the functions f enjoy any additional smoothness property, so our results represent a genuine extension of the findings by Breuer and Major and by Arcones. As anticipated, our techniques are based on the use of Malliavin operators on a Gaussian space, that we combine both with Stein’s method and with an interpolation technique (already applied in [27,30]) which is reminiscent of the ‘smart path method’ in spin glasses — see e.g. Talagrand [35]. In particular, the use of Stein’s method allows us to deal with functions h that are either Lipschitz or indicators of the type h = 1(−∞,z] , whereas the use of interpolations requires test functions that are twice differentiable and with bounded derivatives. Note that this implies that the convergence (1.4) does indeed take place in the stronger topologies of the Kolmogorov and Wasserstein distances. The rest of the paper is organized as follows. Section 2 contains the statements of our main results, some examples and applications. Section 3 presents some notions and results that are needed to prove our main Theorem 2.1. Section 4 is devoted to proofs.

Author's personal copy 796

I. Nourdin et al. / Stochastic Processes and their Applications 121 (2011) 793–812

2. Statement of the main results We keep the assumptions and notation of the previous section. For the sake of notational simplicity, in the following we shall assume that E[ f (X 1 )] = 0. Also, we shall assume that X 1 ∼ (1) (d) Nd (0, Id ), implying that X k , . . . , X k are independent N (0, 1) random variables for all k ∈ Z. Note that this last assumption is not restrictive: indeed, by reduction of variables and at the cost of possibly decreasing the value of d, we may always assume that X k ∼ Nd (0, Σ ) for an invertible matrix Σ ∈ Rd×d , and by a linear transformation we can further restrict ourselves to the case X k ∼ Nd (0, Id ). Observe that, if f (x) has Hermite rank q with respect to X 1 , then f (L x) has the same Hermite rank q with respect to L −1 X 1 , for every invertible matrix L — see also [2, p. 2249]. Now, let Λ denote the set of all vectors α = (α1 , . .� . , αd ) with αi ∈ N�∪ {0}. For any d d multi-index α ∈ Λ, we introduce the notation |α| = i=1 αi and α! = i=1 αi !. When E[ f 2 (X 1 )] < ∞, the function f possesses the unique Hermite expansion � � d d � � � (i) f (x) = aα Hαi (xi ), aα = (α!)−1 E f (X 1 ) Hαi (X 1 ) , (2.6) i=1

α∈Λ

i=1

where (H j ) j≥0 is the sequence of Hermite polynomials, recursively defined as follows: H0 = 1, and H j = δ H j−1 ,

j ≥ 1,

where δ f (x) = x f (x) − f � (x) (for instance: H1 (x) = x, H2 (x) = x 2 − 1, and so on). When f has Hermite rank q ≥ 1 we introduce the decomposition ∞ �

f (x) =

m=q

f m (x),

f m (x) =



α∈Λ:|α|=m



d �

Hαi (xi ).

(2.7)

i=1

Notation. Fix a function f such that E[ f 2 (X 1 )] < ∞ and f has Hermite rank q ≥ 1. Our main results are expressed in terms of the following collection of coefficients (2.8)–(2.12): θ( j) = max |r (i,l) ( j)|

(2.8)

1≤i,l≤d

K = inf {θ ( j) ≤ d −1 , ∀| j| ≥ k} k∈N � θ= θ ( j)q ,

(with inf ∅ = ∞),

(2.9) (2.10)

j∈Z

σm2

=

E[ f m2 (X 1 )] + 2

γn,m,e =



2θ n −1

∞ � k=1



| j|≤n

E[ f m (X 1 ) f m (X 1+k )] (for m ≥ q),

θ ( j)e



| j|≤n

θ ( j)m−e

(for m ≥ q and 1 ≤ e ≤ m − 1).

(2.11) (2.12)

The coefficients θ ( j), K , θ , σm2 and γn,m,e will also be combined into the following expressions (2.13)–(2.17): � � �� � � E[ f 2 (X 1 )] 2K 2 | j| A1,n = + dq θ( j)q + θ( j)q (2.13) 2 n n | j|≤n | j|>n

Author's personal copy I. Nourdin et al. / Stochastic Processes and their Applications 121 (2011) 793–812

A2,N

� � ∞ � � q � 2 = 2(2K + d θ ) E[ f (X 1 )] E[ f m2 (X 1 )]

A3,n,N A4,n,N

A5,n,N

N E[ f 2 (X 1 )] � = 2 m=q



m=N +1

� � m �2 � d m m−1 ll! (2m − 2l)!γn,m,l mm! l=1 l

797

(2.14) �

E[ f 2 (X 1 )](2K + d q θ )1/2 = 2 � � � � s−1 � 1/2 s/2 p! p + s × d (s − p)!γn,s,s− p s! p p−1 q≤ p
(2.15)

(2.16)

� �� � p−1 � E[ f 2 (X 1 )] � p−1 s−1 � = ( p + s) (l − 1)! ( p + s − 2l)! √ l − 1 l − 1 2 2 q≤ p
Note that the coefficients K , θ and σm can in general be infinite, and also that, if E[ f (X 1 )2 ] < ∞, if f has Hermite rank q, and if (1.3) is in order, then 2

σ =

∞ �

m=q

σm2 < ∞,

(2.18)

where σ 2 is defined in Theorem 1.1. The next statement, which is the main result of the paper, asserts that the quantities defined above can be used to write explicit bounds of the type (1.5). Theorem 2.1 (Quantitative Breuer–Major Theorem). Let the notation and assumptions of this section prevail (in particular, E[ f (X 1 )] = 0 and X 1 ∼ Nd (0, Id )), and assume that the conditions of Theorem 1.1 are satisfied. Then, the coefficients appearing in formulae (2.8)– (2.17) are all finite. Moreover, the following three bounds are in order. (1) For any function h ∈ C 2 (R) (that is, h is twice continuously differentiable) with bounded second derivative, and for every n > K , |E[h(Sn )] − E[h(S)]| ≤ �h �� �∞ (A1,n + inf {A2,N + A3,n,N + A4,n,N + A5,n,N }). (2.19) N ≥q

(2) For any Lipschitz function h, and for every n > K ,     � �h �∞ 1 1  |E[h(Sn )] − E[h(S)]| ≤ A2,n  + � � � 2  σ q 2 (2K + d θ ) E f (X 1 )      A1,n + A3,n,N + A4,n,N + A5,n,N  + 4 inf . �  N ≥q N  �   σm2  m=q

(3) For any z ∈ R, and for every n > K ,

(2.20)

Author's personal copy 798

I. Nourdin et al. / Stochastic Processes and their Applications 121 (2011) 793–812 |[P(Sn ≤ z)] − P(S ≤ z)| ≤



2 σ

�   � � A + A3,n,N + A4,n,N + A5,n,N 1 1 �  + 4 inf 1,n × � A2,n  + � . � � � � σ N ≥q N � (2K + d q θ) E f 2 (X 1 ) � σm2

(2.21)

m=q

We will now demonstrate that Theorem 2.1 implies a stronger version of Theorem 1.1, namely that the convergence (1.4) takes place with respect to topologies that are stronger than the one of convergence in distribution. To prove this claim, we need to show in particular that, under the assumptions of Theorem 1.1, γn,m,e → 0 as n → ∞ for any choice of m ≥ q and 1 ≤ e ≤ m −1. This is a consequence of the next Lemma 2.2. In what follows, given positive sequences bn , cn , n ≥ 1, we shall write bn � cn whenever bn /cn is bounded, and bn ∼ cn if bn � cn and cn � bn . Lemma 2.2. Let (ak )k∈Z be a sequence of positive real numbers such that some m ∈ N. If 1 ≤ e ≤ m − 1, then e � n −1+ m ake → 0.



m k∈Z ak

< ∞ for

|k|≤n

� � �n Proof. Fix δ ∈ (0, 1), and decompose the sum as nk=1 = [nδ] + older k=[nδ]+1 . By the H¨ k=1 �∞ m inequality we obtain (recall that k=1 ak is finite) � �e/m [nδ] ∞ � � n −1+e/m ake ≤ n −1+e/m (nδ)1−e/m akm ≤ cδ 1−e/m , k=1

k=1

where c is some constant, as well as � n n � � −1+e/m e n ak ≤ k=[nδ]+1

k=[nδ]+1

akm

�e/m

.

The first term converges to 0 as δ goes to zero (because 1 ≤ e ≤ m − 1), and the second also converges to 0 for fixed δ and n → ∞. This proves the claim. � Now recall that, if X, Y are two real-valued random variables, then the Kolmogorov distance between the law of X and the law of Y is given by d K ol (X, Y ) = sup |P(X ≤ z) − P(Y ≤ z)|. z∈R

(2.22)

If E|X |, E|Y | < ∞, one can also meaningfully define the Wasserstein distance dW (X, Y ) =

sup |E[ f (X )] − E[ f (Y )]|,

f ∈Lip(1)

(2.23)

where Lip(1) indicates the collection of all Lipschitz functions with Lipschitz constant ≤1. Finally, if X, Y have finite second moments, for every constant C > 0 one can define the distance dC (X, Y ) = sup |E[ f (X )] − E[ f (Y )]|, f

2 ∈D C

(2.24)

Author's personal copy I. Nourdin et al. / Stochastic Processes and their Applications 121 (2011) 793–812

799

where DC2 stands for the class of all functions that are twice continuously differentiable and have a second derivative bounded by C. Note that the topologies induced by d K ol , dW and dC , on the probability measures on R, are strictly stronger than the topology of convergence in distribution (see e.g. [16, Ch. 11]). The next consequence of Theorem 2.1 provides the announced refinement of Theorem 1.1. Corollary 2.3 (Breuer–Major, Strong Version). Let the notation and assumptions of this section prevail (in particular, E[ f (X 1 )] = 0 and X 1 ∼ Nd (0, Id )), and assume that the conditions of Theorem 1.1 are satisfied. Then, the convergence (1.4) takes place with respect to the three distances d K ol , dW and dC (for all C > 0), namely lim d K ol (Sn , S) = lim dW (Sn , S) = lim dC (Sn , S) = 0.

n→∞

n→∞

n→∞

Proof. Under the assumptions of Theorem 1.1, one has that A1,n → 0 as n → ∞ (because � θ < ∞, | j|≤n θ ( j)q |nj| → 0 as n → ∞ by bounded convergence). On the other hand, because of (2.18) and since E[ f 2 (X 1 )] < ∞, one has that A2,N → 0 as N → ∞. Moreover, since γn,m,e → 0 for any m ≥ q and 1 ≤ e ≤ m −1 (due to Lemma 2.2), one has that A j,n,N → 0, j = 3, 4, 5, for any fixed N as n → ∞. We deduce that inf N ≥q {A2,N + A3,n,N + A4,n,N + A5,n,N } → 0 as n → ∞. To conclude the proof, it remains to apply (2.19)–(2.21). � Next, we present a simplified version of Theorem 2.1 for d = 1 and f = Hq , where Hq is the qth Hermite polynomial. Notice that in this case K = 0 and the constants A1,n and A3,n,N = A3,n are given by � � � � q! | j| A1,n = θ |r ( j)|q + |r ( j)|q , (2.25) 2 n | j|≤n | j|>n A3,n

q−1 1 � � q �2 � = ll! (2q − 2l)!γn,q,l , 2q l=1 l

(2.26)

with γn,q,l defined by (2.12).

Corollary 2.4 (Hermite Subordination). Assume that d = 1, f = Hq and and let A1,n , A3,n be given by (2.25)–(2.26).



(1) For any function h ∈ C 2 (R) with bounded second derivative it holds that |E[h(Sn )] − E[h(S)]| ≤ �h �� �∞ (A1,n + A3,n ).

q j∈Z |r ( j)|

< ∞,

(2.27)

(2) For any Lipschitz function h it holds that

2�h � �∞ |E[h(Sn )] − E[h(S)]| ≤ (A1,n + A3,n ). σ (3) For any z ∈ R it holds that

(2.28)

2�sz� �∞ (A1,n + A3,n ), (2.29) σ where sz is the solution of the Stein’s equation associated with the function h(x) = 1(−∞,z] (x), i.e. sz solves the differential equation |[P(Sn ≤ z)] − P(S ≤ z)| ≤

1(−∞,z] (x) − Φ(z) = sz� (x) − xsz (x),

x ∈ R,

Author's personal copy 800

I. Nourdin et al. / Stochastic Processes and their Applications 121 (2011) 793–812

with Φ being the distribution function of N (0, 1). Furthermore, we have that �sz� �∞ ≤ 1 for all z ∈ R.

Proof. From Theorem 3.1 in [25] and Theorem 3.2 in Section 3.3 we obtain the estimate |E[h(Sn )] − E[h(S)]| ≤ ch E|σ 2 − �D F, −DL −1 F�H |, ��



�s � �

where ch = �h 2�∞ in (1), ch = �hσ�∞ in (2) and ch = zσ ∞ in (3). We readily deduce the assertion since E|σ 2 − �D F, −DL −1 F�H | ≤ 2(A1,n + A3,n ), which follows from the proof of Theorem 2.1. � We remark that the upper bound in (2.29) of Corollary 2.4 is more efficient than the general upper bound obtained in (2.21) of Theorem 2.1 as the latter one converges to 0 more slowly. Next, we apply Corollary 2.4 to some particular classes of covariance functions r . Example 2.5 (Covariance Functions with Polynomial Decay). Assume that d = 1 and f = Hq with q ≥ 2, and consider a covariance function r which is regularly varying with parameter a < 0. That is, for all |k| ≥ 1, |r (k)| = |k|a l(|k|), where l is a slowly varying function. Recall that for any regularly varying function r with parameter α < 0, we have the following discrete version of Karamata’s theorem (see e.g. [6]):

α > −1 :

n �

k=1 n a+1l(n)

|r (k)|

→ 1/(α + 1),

α < −1 :

∞ �

|r (k)|

→ −1/(α + 1),

k=n n a+1l(n)

as n → ∞. Assume now that a < − q1 , which implies that the conditions of Theorem 2.1 are satisfied, and ae �= −1 for any e = 1, . . . , q − 1. By the aforementioned convergence results we immediately deduce the following estimates (1 ≤ e ≤ q − 1): � | j| |r ( j)|q � n −1 + n aq+1l(n), n | j|≤n � |r ( j)|q � n aq+1l(n), | j|>n

γn,q,e � n −1/2 + n a/2l(n) + n (aq+1)/2l 2 (n).

Thus, for all three cases of Corollary 2.4 we conclude that  −1/2 n : a <� −1   �   1  a/2 n l(n) : a ∈ −1, − |E[h(Sn )] − E[h(S)]| � q −1 � �   aq+1 1 1   ,− . n 2 l 2 (n) : a ∈ − q −1 q

Clearly, the same estimates hold for d(Sn , S), where d = d K ol , d = dW or d = dC . Example 2.6 (The Fractional Brownian Motion Case). Let d = 1, f = Hq with q ≥ 2 and H , where B H is a fractional Brownian consider the fractional Gaussian noise X i = BiH − Bi−1

Author's personal copy I. Nourdin et al. / Stochastic Processes and their Applications 121 (2011) 793–812

801

motion with parameter H ∈ (0, 1). Recall that B H = (BtH )t≥0 is a centered Gaussian process (with stationary increments) with covariance structure given by 1 2H (|t| + |s|2H − |t + s|2H ). 2 It is well-known that the correlation function r of the fractional Brownian noise has the following form: E[Bt Bs ] =

|r (k)| = |k|2H −2l(|k|),

k ≥ 1,

with l(|k|) → 2H |2H − 1| as |k| → ∞ when H �= 12 , and l(|k|) = 0 for |k| ≥ 1 when H = 12 . As in the previous example we immediately deduce that � �  1 −1/2  n : H ∈ 0,    2   � �  1 2q − 3 H −1/2 |E[h(Sn )] − E[h(S)]| � n : a∈ ,  2 2q − 2   � �   2q − 3 2q − 1  (2q H −2q+1)/2 n : a∈ , 2q − 2 2q

and the same estimates hold for d(Sn , S) with d = d K ol , d = dW or d = dC . Let us remark that these upper bounds coincide with those derived in Theorem 4.1 in [25]. We finally remark that the rate n −1/2 for H ∈ (0, 12 ] has been proved to be optimal in [26], i.e. |E[h(Sn )] − E[h(S)]| ∼ n −1/2 . For the other two cases the optimality question is still an open problem. 3. The toolbox 3.1. Malliavin calculus on a Gaussian space We shall now provide a short introduction to the tools of Malliavin calculus that are needed in the proof of our main Theorem 2.1. The reader is referred to [28] for any unexplained definition or result. Let H be a real separable Hilbert space. We denote by W = {W (h) : h ∈ H} an isonormal Gaussian process over H, that is, W is a centered Gaussian family indexed by the elements of H and such that, for every g1 , g2 ∈ H, � � E W (g1 )W (g2 ) = �g1 , g2 �H . (3.30) In what follows, we shall use the notation L 2 (W ) = L 2 (Ω , σ (W ), P). For every q ≥ 1, we write H⊗q to indicate the qth tensor power √ of H; the symbol H⊙q indicates the qth symmetric tensor power of H, equipped with the norm q!� · �H⊗q . We denote by Iq the isometry between H⊙q and the qth Wiener chaos of X . It is well-known (see again [28, Ch. 1]) that any random variable F belonging to L 2 (W ) admits the chaotic expansion: F=

∞ �

Iq ( f q ),

(3.31)

q=0

where I0 ( f 0 ) := E[F], the series converges in L 2 and the kernels f q ∈ H⊙q , q ≥ 1, are uniquely determined by F. In the particular case where H = L 2 (A, A , µ), with (A, A ) a measurable

Author's personal copy 802

I. Nourdin et al. / Stochastic Processes and their Applications 121 (2011) 793–812

space and µ a σ -finite and non-atomic measure, one has that H⊙q = L 2s (Aq , A ⊗q , µ⊗q ) is the space of symmetric and square integrable functions on Aq . Moreover, for every f ∈ H⊙q , Iq ( f ) coincides with the multiple Wiener–Itˆo integral (of order q) of f with respect to W (see [28, Ch. 1]). It is well-known that a random variable of the type Iq ( f ), f ∈ H⊙q , has finite moments of all orders (see [18, Ch. VI]). For every q ≥ 0, we write Jq to indicate the orthogonal projection operator on the qth Wiener chaos associated with W , so, if F ∈ L 2 (W ) is as in (3.31), then Jq F = Iq ( f q ) for every q ≥ 0. Let {ek , k ≥ 1} be a complete orthonormal system in H. Given f ∈ H⊙ p and g ∈ H⊙q , for every r = 0, . . . , p ∧ q, the r th contraction of f and g is the element of H⊗( p+q−2r ) defined as f ⊗r g =

∞ �

i 1 ,...,ir =1

� f, ei1 ⊗ · · · ⊗ eir �H⊗r ⊗ �g, ei1 ⊗ · · · ⊗ eir �H⊗r .

(3.32)

In the particular case where H = L 2 (A, A , µ) (with µ non-atomic), one has that � f ⊗r g = f (t1 , . . . , t p−r , s1 , . . . , sr ) Ar

× g(t p−r +1 , . . . , t p+q−2r , s1 , . . . , sr )dµ(s1 ) . . . dµ(sr ).

Moreover, f ⊗0 g = f ⊗ g equals the tensor product of f and g while, for p = q, f ⊗ p g = � f, g�H⊗ p . Note that, in general, the contraction f ⊗r g is not a symmetric element �r g. The following of H⊗( p+q−2r ) . The canonical symmetrization of f ⊗r g is written as f ⊗ ⊙ p ⊙q multiplication formula is also very useful: if f ∈ H and g ∈ H , then I p ( f )Iq (g) =

p∧q � r =0

r!

� p � �q � r

r

�r g). I p+q−2r ( f ⊗

(3.33)

Let S be the set of all smooth cylindrical random variables of the form � � F = g W (φ1 ), . . . , W (φn ) ,

where n ≥ 1, g : Rn → R is a smooth function with compact support and φi ∈ H. The Malliavin derivative of F with respect to W is the element of L 2 (Ω , H) defined as DF =

n � � ∂g � W (φ1 ), . . . , W (φn ) φi . ∂ xi i=1

Also, DW (φ) = φ for every φ ∈ H. As usual, D1,2 denotes the closure of S with respect to the norm � · �1,2 , defined by the relation � � �F�21,2 = E[F 2 ] + E �D F�2H .

Note that, if F is equal to a finite sum of multiple Wiener–Itˆo integrals, then F ∈ D1,2 . The Malliavin derivative D verifies the following chain rule: if ϕ : Rn → R is in Cb1 (that is, the collection of continuously differentiable functions with bounded partial derivatives) and if {Fi }i=1,...,n is a vector of elements of D1,2 , then ϕ(F1 , . . . , Fn ) ∈ D1,2 and

Author's personal copy I. Nourdin et al. / Stochastic Processes and their Applications 121 (2011) 793–812

803

n � ∂ϕ Dϕ(F1 , . . . , Fn ) = (F1 , . . . , Fn )D Fi . ∂ xi i=1

We denote by δ the adjoint of the operator D, also called the divergence operator. A random element u ∈ L 2 (Ω , H) belongs to the domain of δ, denoted as Domδ, if and only if it verifies |E�D F, u�H | ≤ cu �F� L 2

for any F ∈ S ,

where cu is a constant depending only on u. If u ∈ Domδ, then the random variable δ(u) is defined by the duality relationship (sometimes called the ‘integration by parts formula’) E[Fδ(u)] = E�D F, u�H ,

(3.34)

which holds for every F ∈ D1,2 . The operator L, acting on square integrable random �∞ variables of the type (3.31), is defined through the projection operators {Jq }q≥0 as L = q=0 −q Jq , and is called the infinitesimal generator of the Ornstein–Uhlenbeck semigroup. It verifies the following crucial property: a random variable F is an element of DomL (=D2,2 ) if, and only if, F ∈ Domδ D (i.e. F ∈ D1,2 and D F ∈ Domδ), and in this case, δ D F = −L F. Note that a random variable F as in (3.31) is in D1,2 if and only if ∞ � (q + 1)!� f q �2H⊗q < ∞, q=1

� � � and also E �D F�2H = q≥1 qq!� f q �2H⊗q . If H = L 2 (A, A , µ) (with µ non-atomic), then the derivative of a random variable F as in (3.31) can be identified with the element of L 2 (A × Ω ) given by Da F =

∞ � q=1

� � q Iq−1 f q (·, a) ,

a ∈ A.

(3.35)

We also define the operator L�−1 , which is the pseudo-inverse of L, as follows: for every F ∈ L 2 (W ), we set L −1 F = q≥1 q1 Jq (F). Note that L −1 is an operator with values in D2,2 and that L L −1 F = F − E[F] for all F ∈ L 2 (W ). 3.2. Assessing norms and scalar products The following statement plays a crucial role in the proof of Theorem 2.1. Lemma 3.1. Let F = I p (h) and G = Is (g) with h ∈ H⊙ p , g ∈ H⊙s and p < s ( p, s ≥ 1). Then � � s−1 � �4 1 1 � 2 2 2 s �l g�2 ⊗2s−2l , Var �DG�H = 2 l l! (2s − 2l)!�g ⊗ (3.36) H s l s l=1 and

E

��

1 �D F, DG�H s

�2 �

≤ p!



s−1 p−1

�2

(s − p)!E[F 2 ]�g ⊗s− p g�H⊗2 p

Author's personal copy 804

I. Nourdin et al. / Stochastic Processes and their Applications 121 (2011) 793–812

� �2 � � p−1 p2 � s−1 2 2 p−1 + (l − 1)! ( p + s − 2l)! 2 l=1 l −1 l −1 × (�h ⊗ p−l h�2H⊗2l + �g ⊗s−l g�2H⊗2l ).

(3.37)

Proof of 3.36. We have DG = s Is−1 (g), and so, by using (3.33), � s−1 � � 1 s−1 2 2 2 �l+1 g) �DG�H = s�Is−1 (g)�H = s l! I2s−2−2l (g ⊗ s l l=0 � � s � s−1 2 �l g) =s (l − 1)! I2s−2l (g ⊗ l − 1 l=1 � � s−1 � s−1 2 2 �l g) = s!�g�H⊗s + s (l − 1)! I2s−2l (g ⊗ l − 1 l=1 � �2 s−1 � s − 1 �l g). = E[G 2 ] + s (l − 1)! I2s−2l (g ⊗ l − 1 l=1 The orthogonality property of multiple integrals leads to (3.36).

Proof of 3.37. Thanks once again to (3.33), we can write � � �D F, DG�H = ps I p−1 (h), Is−1 (g) H p∧s−1 � � p − 1� �s − 1� �l+1 g) = ps l! I p+s−2−2l (h ⊗ l l l=0 � �� � p∧s � p−1 s−1 �l g). = ps (l − 1)! I p+s−2l (h ⊗ l − 1 l − 1 l=1 It follows that �� �2 � p � 1 2 �D F, DG�H E =p (l − 1)!2 s l=1 � �2 � �2 p−1 s−1 �l g�2 ⊗( p+s−2l) . × ( p + s − 2l)!�h ⊗ H l −1 l −1

(3.38)

If l < p, then

�l g�2 ⊗( p+s−2l) ≤ �h ⊗l g�2 ⊗( p+s−2l) = �h ⊗ p−l h, g ⊗s−l g�H⊗2l �h ⊗ H H ≤ �h ⊗ p−l h�H⊗2l �g ⊗s−l g�H⊗2l 1 ≤ (�h ⊗ p−l h�2H⊗2l + �g ⊗s−l g�2H⊗2l ). 2

If l = p, then

� p g�2 ⊗(s− p) ≤ �h ⊗ p g�2 ⊗(s− p) ≤ �h�2H⊗ p �g ⊗s− p g�H⊗2 p . �h ⊗ H H

By plugging these last expressions into (3.38), we deduce immediately (3.37).



Author's personal copy I. Nourdin et al. / Stochastic Processes and their Applications 121 (2011) 793–812

805

3.3. Estimates via interpolations and Stein’s method The forthcoming Theorem 3.2 contains two bounds on normal approximations, that are expressed in terms of Malliavin operators. As anticipated, the proof of Point (1) uses an interpolation technique already applied in [27,30], which is close to the ‘smart path method’ of spin glasses [35]. Point (2) uses estimates from [25]. Theorem 3.2. Let F be a centered element of D1,2 and let Z ∼ N (0, σ 2 ), σ > 0.

(1) Suppose that h : R → R is twice continuously differentiable and has a bounded second derivative. Then, �� � � �E[h(F)] − E[h(Z )]� ≤ �h �∞ E|σ 2 − �D F, −DL −1 F�H |. 2 (2) Suppose that h : R → R is Lipschitz. Then, � � � �E[h(F)] − E[h(Z )]� ≤ �h �∞ E|σ 2 − �D F, −DL −1 F�H |. σ

(3.39)

(3.40)

Proof. (1) Without loss of generality, we may assume that F and Z are independent and defined on the√same probability space. Fix h as in the statement, and define the function √ Ψ (t) = E[h( 1 − t F + t Z )], t ∈ [0, 1]. Standard results imply that Ψ is differentiable for every t ∈ (0, 1), and that � �√ � �√ √ � � √ � � 1 1 � � � Ψ (t) = √ E h 1 − tF + tZ Z − √ E h 1 − tF + tZ F . 2 t 2 1−t By using independence and integration by parts, we obtain immediately that � �√ √ � � σ 2 � �� �√ √ �� 1 1 − tF + tZ Z = E h 1 − tF + tZ . √ E h� 2 2 t

On the other hand, the relation F = L L −1 F = −δ DL −1 F and (3.34) imply that � �√ √ � � 1 E h� 1 − tF + tS F √ 2 1−t � �√ � √ � 1 � −1 = √ E h 1 − t F + t Z δ(−DL F) 2 1−t � √ � 1 � �√ = E h �� 1 − t F + t Z �D F, −DL −1 F�H . 2 The conclusion follows from the fact that � 1 � � � � � � � �E[h(F)] − E[h(Z )]� = �Ψ (1) − Ψ (0)� ≤ �Ψ (t)�dt. 0

(2) Here we follow the arguments contained in the proof of Theorem 3.1 in [25]. Define h σ (x) = h(σ x), Fσ = σ −1 F, and Z σ = σ −1 Z ∼ N (0, 1). Let s be the solution of the Stein equation associated with h σ , i.e. s solves the differential equation h σ (x) − E[h σ (Z σ )] = s � (x) − xs(x),

x ∈ R.

�x 1 It is well-known that such a solution is given by s(x) = ϕ(x) −∞ (h σ (t) − Φ(t))ϕ(t)dt, where ϕ and Φ are the density and the distribution function of N (0, 1), respectively, and �s � �∞ ≤ �h �σ �∞ .

Author's personal copy 806

I. Nourdin et al. / Stochastic Processes and their Applications 121 (2011) 793–812

Since Fσ is a centered element of D1,2 it holds that Fσ = L L −1 Fσ = −δ DL −1 Fσ . By the integration by parts formula (3.34) we deduce that � � � � �E[h(F)] − E[h(Z )]� = �E[h σ (Fσ )] − E[h σ (Z σ )]� � � = �E[s � (Fσ ) − Fσ s(Fσ )]� � � = �E[s � (Fσ )(1 − �D Fσ , −DL −1 Fσ �H )]� ≤ �s � �∞ E|1 − �D Fσ , −DL −1 Fσ �H |

≤ �h �σ �∞ E|1 − �D Fσ , −DL −1 Fσ �H |.

We conclude by using the relations �h �σ �∞ = σ �h � �∞ and �D Fσ , −DL −1 Fσ �H = σ −2 �D F, −DL −1 F�H .



When applied to the special case of a Gaussian random variable F, Theorem 3.2 yields the following neat estimates. Corollary 3.3. Let F ∼ N (0, γ 2 ) and Z ∼ N (0, σ 2 ), γ , σ > 0.

(1) For every h twice continuously differentiable and with a bounded second derivative, �� � � �E[h(F)] − E[h(Z )]� ≤ �h �∞ |σ 2 − γ 2 |. 2 (2) For every Lipschitz function h,

(3.41)

� � � �E[h(F)] − E[h(Z )]� ≤ �h �∞ |σ 2 − γ 2 |. σ ∨γ

(3.42)

4. Proof of Theorem 2.1 4.1. Preparation

First, let us remark that the process X = (X k )k∈Z can always be regarded as a subset of an isonormal Gaussian process {W (u) : u ∈ H}, where H is a separable Hilbert space with scalar product �·, ·�H . More precisely, we shall assume (without loss of generality) that, for every k ∈ Z and every 1 ≤ l ≤ d, there exists u k,l ∈ H such that (l)

X k = W (u k,l ),



and consequently �u k,l , u k � ,l � �H = r (l,l ) (k − k � ),

for every k, k � ∈ Z and every 1 ≤ l, l � ≤ d. Observe also that H can be taken of the form H = L 2 (A, A , µ), where µ is σ -finite and non-atomic. Using the Hermite expansion (2.6) of the function f we obtain the Wiener chaos representation ∞ � n n Sn = Im (gm ), gm ∈ H⊙m , m=q

n have the form where the kernels gm

n � 1 � n gm =√ bt u k,t1 ⊗ · · · ⊗ u k,tm n k=1 t∈{1,...,d}m

(4.43)

Author's personal copy I. Nourdin et al. / Stochastic Processes and their Applications 121 (2011) 793–812

807

for certain coefficients bt such that the mapping t �→ bt is symmetric on {1, . . . , d}m . One also has the identities �

E[ f m2 (X 1 )] = m!

t∈{1,...,d}m

bt2 ,

m ≥ q,

E[ f 2 (X 1 )] =

∞ �

m!

m=q



t∈{1,...,d}m

bt2 .

Here is a useful preliminary result. n defined in (4.43) and any 1 ≤ e ≤ m−1 we obtain the inequality, Lemma 4.1. For the kernels gm valid for every n, n �gm

n ⊗e gm �H⊗2(m−e)

dm ≤ E[ f m2 (X 1 )]γn,m,e , m!

(4.44)

where γn,m,e is defined by (2.12). Furthermore, we have that, for every n, n 2 m!�gm �H⊗m ≤ E[ f m2 (X 1 )](2K + d q θ),

(4.45)

where the constants K and θ are defined, respectively, in (2.9) and (2.10). Proof of 4.44. Fix 1 ≤ e ≤ m − 1. Observe that n gm

n ⊗e gm

n e � � 1 � = bt bs r (t j ,s j ) (k1 − k2 ) n k ,k =1 t,s∈{1,...,d}m j=1 1

2

× u k1 ,t p+1 ⊗ · · · ⊗ u k1 ,tm ⊗ u k2 ,s p+1 ⊗ · · · ⊗ u k2 ,sm .

We obtain n �gm

n 2 ⊗e gm �H⊗2(m−e)

× n −2

n �

k1 ,...,k4 =1





�2 dm 2 E[ f m (X 1 )] m!

θ (k1 − k2 )e θ (k3 − k4 )e θ (k1 − k3 )m−e θ(k2 − k4 )m−e ,

where θ ( j) is defined in (2.8). Since θ (k3 − k4 )e θ(k1 − k3 )m−e ≤ θ(k3 − k4 )m + θ(k1 − k3 )m we deduce that n −2

n �

k1 ,...,k4 =1

≤ 2n −1

θ (k1 − k2 )e θ (k3 − k4 )e θ (k1 − k3 )m−e θ(k2 − k4 )m−e �

θ (k)m

k∈Z

Hence, we obtain (4.44).



|k|≤n



θ (k)e



|k|≤n

2 θ (k)m−e ≤ γn,m,e .

Proof of 4.45. By the Cauchy–Schwarz inequality we have �� � � � � � � � � � m! �� bt u k,t1 ⊗ · · · ⊗ u k,tm , bt u k+l,t1 ⊗ · · · ⊗ u k+l,tm � m � t∈{1,...,d}m � t∈{1,...,d} ⊗m H ≤ E[ f m2 (X 1 )].

Author's personal copy 808

I. Nourdin et al. / Stochastic Processes and their Applications 121 (2011) 793–812

We deduce, for any m ≥ q, n 2 m!�gm �H⊗m

n m � � m! � = bt bs r (t j ,s j ) (k1 − k2 ) n k ,k =1 t,s∈{1,...,d}m j=1 1 2  � �2  � � ≤ 2H E[ f m2 (X 1 )] + m! θ(k)m |bt |  |k|≥K



≤ E[ f m2 (X 1 )] 2K + which implies (4.45).





(dθ(k))m

|k|≥K



t∈{1,...,d}m

≤ E[ f m2 (X 1 )](2K + d q θ),

The proofs of Point 1 and Point 2 in Theorem 2.1 are similar, and are detailed in the subsequent two sections. 4.2. Proof of Theorem 2.1-(1) � First of all, we remark that θ ( j) → 0 as | j| → ∞, because j∈Z θ( j)q < ∞. This implies that K < ∞, where the constant � K is defined in (2.9). Moreover, the asymptotic variance σ 2 is n 2 2 q finite. Indeed we have that σ 2 = ∞ m=q m!�gm �H⊗m ≤ E[ f (X 1 )](2K + d θ) due to (4.45). The main proof is composed of several steps. (a) Reduction to a finite chaos expansion. We start by approximating Sn by a finite sum of multiple integrals. Define Sn,N =

N �

m=q

n Im (gm ).

Now, let h ∈ C 2 (R) be a function with bounded second derivative. Since |h(x) − h(y) − h � (0)(x − y)| ≤

1 �� �h �∞ (y − x)2 + �h �� �∞ |x||y − x| 2

for all x, y ∈ R, we immediately obtain that

|E[h(Sn )] − E[h(Sn,N )]| ≤ �h �� �∞ � � 1 × �Sn − Sn,N �2L 2 (P) + �Sn � L 2 (P) �Sn − Sn,N � L 2 (P) . 2

By inequality (4.45) we deduce that

�Sn �2L 2 (P) ≤ (2K + d q θ )� f (X 1 )�2L 2 (P) and �Sn − Sn,N �2L 2 (P) ≤ (2K + d q θ )

∞ �

m=N +1

� f m (X 1 )�2L 2 (P)

≤ (2K + d q θ )� f (X 1 )� L 2 (P)



∞ �

m=N +1

� f m (X 1 )�2L 2 (P)

�1/2

,

Author's personal copy I. Nourdin et al. / Stochastic Processes and their Applications 121 (2011) 793–812

809

where θ is defined by (2.10). We conclude that |E[h(Sn )] − E[h(Sn,N )]| ≤

3(2K + d q θ) �� �h �∞ � f (X 1 )� L 2 (P) 2 � �1/2 ∞ � × � f m (X 1 )�2L 2 (P) ,

(4.46)

m=N +1

which completes the first step. � (b) Bounds based on the interpolation inequality (3.39). Let Z N be a centered Gaussian �N 2 random variable with variance m=q σm . By using (3.39) in the special case F = Sn,N , Z = Z N and by applying e.g. (3.35), we obtain � � N �� � 1 �� � � 2 −1 E[[h(Z N )] − E[h(Sn,N )]] ≤ �h �∞ � σm − �DSn,N , −DL Sn,N �H � �m=q � 2 2 L (P)

N � 1 �� ≤ �h �∞ �δ ps σ p2 − s −1 �D I p (g np ), D Is (gsn )�H � L 2 (P) , 2 p,s=q

(4.47)

where δ ps is the Kronecker symbol. This completes the second step. � (c) The final estimates. Here we give the approximation of the term on the right-hand side of (4.47). By (4.45) and the dominating convergence theorem we immediately deduce that E[s

−1

�D Is (gsn )�2H ]

=

s!�gsn �2H⊗s



σs2

= s!





bt bl

k∈Z t,l∈{1,...,d}s

m �

r (t j ,l j ) (k).

j=1

As in the proof of (4.45) we conclude that (recall that we assumed n > K ) � � m � |k| � �� � � � −1 n 2 2 (t j ,l j ) |E[s �D Is (gs )�H ] − σs | ≤ s! bt bl r (k)� � � � K |k|n Thus we have N � s=q

�σs2 − s −1 �D Is (gsn ) �2H � L 2 (P)

� N � � � −1 n 2 2 2 −1 n ≤ |E[s �D Is (gs )�H ] − σs | + Var[s �D Is (gs )�H ] s=q

Author's personal copy 810

I. Nourdin et al. / Stochastic Processes and their Applications 121 (2011) 793–812



2H 2 ≤ E[ f s2 (X 1 )] + dq n +

N � � s=q





� | j| θ ( j)q + θ( j)q n | j|≤n | j|>n

��

Var[s −1 �D Is (gsn )�2H ]

≤ 2(A1,n + A3,n,N ), where the last inequality follows directly from Lemmas 3.1 and 4.1. On the other hand, using the approximation (4.45) and Lemmas 3.1 and 4.1 once again we deduce that � � �s −1 �D I p (g np ), D Is (gsn )�H � L 2 (P) + �s −1 �D I p (g np ), D Is (gsn )�H � L 2 (P) q≤ p
=



q≤ p
q≤s< p≤N

p + s −1 �s �D I p (g np ), D Is (gsn )�H � L 2 (P) ≤ 2(A4,n,N + A5,n,N ). p

Thus, we conclude that N � � �� � �� � � σm Ym − h(Sn,N ) � ≤ �h �� �∞ (A1,n + A3,n,N + A4,n,N + A5,n,N ), �E h

(4.48)

m=q

which finishes this step. � (d) Putting things together. We have the inequality

|E[h(Sn )] − E[h(S)]| ≤ |E[h(Sn )] − E[h(Sn,N )]| + |E[h(Sn,N )] − E[h(Z N )]| + |E[h(Z N )] − E[h(S)]|.

(4.49)

We are left with the derivation of a bound for the third term. By using (3.41) in the special case F = Z N , we deduce that ∞ � � � 1 �� � � σm2 �E[h(Z N )] − E[h(S)]� ≤ �h �∞ 2 m=N +1

∞ � 2H + d q θ �� ≤ �h �∞ E[ f m2 (X 1 )], 2 m=N +1

(4.50)

where the last inequality is deduced by Lemma 4.1. The latter is smaller than 14 �h �� �∞ A2,N , which together with (4.46)–(4.48) completes the proof of Theorem 2.1-(1). � 4.3. Proof of Theorem 2.1-(2) Take h Lipschitz and write the inequality (4.49). Like for (4.46), standard computations yield |E[h(Sn )] − E[h(Sn,N )]| ≤ �h � �∞ �Sn − Sn,N � L 2 (P) ≤ �h � �∞ (2K + d q θ)1/2 � � ∞ � � ×� � f m (X 1 )�2L 2 (P) . m=N +1

Author's personal copy I. Nourdin et al. / Stochastic Processes and their Applications 121 (2011) 793–812

811

By applying inequality (3.40) in the case F = Sn,N , Z = Z N , one has that �h � �∞ |E[h(Sn,N )] − E[h(Z N )]| ≤ � �1/2 N � σm2 m=q

N �

× ≤ �

�δ ps σ p2 − s −1 �D I p (g np ), D Is (gsn )�H � L 2 (P)

p,s=q �h � �∞

N �

m=q

σm2

�1/2 2(A1,n + A3,n,N + A4,n,N + A5,n,N ),

where the last inequality is obtained by reasoning as in Part (c) of Section 4.2. Finally, an application of inequality (3.42) in the case F = Z N yields (since θ ≥ 1) �h � �∞ |E[h(Z N )] − E[h(S)]| ≤ σ

∞ �

m=N +1

σm2 ≤

�h � �∞ A2,N . 2σ

Putting the above estimates together yields the desired conclusion.



4.4. Proof of Theorem 2.1-(3) From [11, Theorem 3.1], one can deduce that � d K ol (Sn , S) ≤ 2 dW (Sn /σ, S/σ ).

Hence, we get the desired conclusion by combining this inequality with (2.20).



Acknowledgement We warmly thank Arnaud Guillin for pointing out to us an inaccuracy in a previous version. References [1] V.V. Anh, N.N. Leonenko, Renormalization and homogenization of fractional diffusion equations with random data, Probab. Theory Related Fields 124 (3) (2002) 381–408. [2] M.A. Arcones, Limit theorems for nonlinear functionals of a stationary Gaussian sequence of vectors, Ann. Probab. 22 (4) (1994) 2242–2274. [3] O.E. Barndorff-Nielsen, J.M. Corcuera, M. Podolskij, Power variation for Gaussian processes with stationary increments, Stochastic Process Appl. 119 (2009) 1845–1865. [4] O.E. Barndorff-Nielsen, J.M. Corcuera, M. Podolskij, Multipower variation for Brownian semi-stationary processes, Bernoulli (2009) (in press). [5] L. Beghin, V.P. Knopova, N.N. Leonenko, E. Orsingher, Gaussian limiting behavior of the rescaled solution to the linear Korteweg–de Vries equation with random initial conditions, J. Stat. Phys. 99 (3–4) (2000) 769–781. [6] N.H. Bingham, C.M. Goldie, J.L. Teugels, Regular Variation, Cambridge University Press, 1987. [7] J.-C. Breton, I. Nourdin, Error bounds on the non-normal approximation of Hermite power variations of fractional Brownian motion, Electron. Comm. Probab. 13 (2008) 482–493 (electronic). [8] P. Breuer, P. Major, Central limit theorems for nonlinear functionals of Gaussian fields, J. Multivariate Anal. 13 (3) (1983) 425–441. [9] C. Berzin, J. Le´on, Estimating the Hurst parameter, Stat. Inference Stoch. Process. 10 (1) (2007) 49–73.

Author's personal copy 812

I. Nourdin et al. / Stochastic Processes and their Applications 121 (2011) 793–812

[10] B. Buchmann, N.-H. Chan, Integrated functionals of normal and fractional processes, Ann. Appl. Probab. 19 (1) (2009) 49–70. [11] L. Chen, Q.-M. Shao, Stein’s method for normal approximation, in: An Introduction to Stein’s Method, in: Lect. Notes Ser. Inst. Math. Sci. Natl. Univ. Singap., vol. 4, Singapore Univ. Press, Singapore, 2005, pp. 1–59. [12] J.-F. Coeurjolly, Identification of multifractional Brownian motion, Bernoulli 11 (6) (2005) 987–1008. [13] J.-F. Coeurjolly, Hurst exponent estimation of locally self-similar Gaussian processes using sample quantiles, Ann. Statist. 36 (3) (2008) 1404–1434. [14] J.M. Corcuera, D. Nualart, J.H.C. Woerner, Power variation of some integral fractional processes, Bernoulli 12 (4) (2006) 713–735. [15] R.L. Dobrushin, P. Major, Non-central limit theorems for nonlinear functionals of Gaussian fields, Z. Wahrscheinlichkeitstheor. Verwandte Geb. 50 (1) (1979) 27–52. [16] R.M. Dudley, Real Analysis and Probability, 2nd ed., Cambridge University Press, Cambridge, 2003. [17] L. Giraitis, D. Surgailis, CLT and other limit theorems for functionals of Gaussian processes, Z. Wahrscheinlichkeitstheor. Verwandte Geb. 70 (1985) 191–212. [18] S. Janson, Gaussian Hilbert Spaces, Cambridge University Press, Cambridge, 1997. [19] M.F. Kratz, J.R. Le´on, Central limit theorems for level functionals of stationary Gaussian processes and fields, J. Theoret. Probab. 14 (3) (2001) 639–672. [20] J. Le´on, C. Lude˜na, Limits for weighted p-variations and likewise functionals of fractional diffusions with drift, Stochastic Process. Appl. 117 (3) (2007) 271–296. [21] C. Lude˜na, L p -variations for multifractal fractional random walks, Ann. Appl. Probab. 18 (3) (2008) 1138–1163. [22] D. Marinucci, The empirical process for bivariate sequences with long memory, Stat. Inference Stoch. Process. 8 (2) (2002) 205–223. [23] D. Marinucci, A Central limit theorem and higher order results for the angular bispectrum, Probab. Theory Related Fields 141 (2007) 389–409. [24] I. Nourdin, D. Nualart, C.A. Tudor, Central and non-central limit theorems for weighted power variations of fractional Brownian motion, Ann. Inst. H. Poincar´e Probab. Statist. 46 (4) (2008) 1055–1079. [25] I. Nourdin, G. Peccati, Stein’s method on Wiener chaos, Probab. Theory Related Fields 145 (1) (2009) 75–118. [26] I. Nourdin, G. Peccati, Stein’s method and exact Berry–Esseen asymptotics for functionals of Gaussian fields, Ann. Probab. 37 (6) (2010) 2231–2261. [27] I. Nourdin, G. Peccati, G. Reinert, Invariance principles for homogeneous sums: universality of Gaussian Wiener chaos, Ann. Probab. 38 (5) (2010) 1947–1985. [28] D. Nualart, The Malliavin Calculus and Related Topics, 2nd ed., Springer, Berlin, 2006. [29] G. Peccati, M.S. Taqqu, Wiener Chaos: Moments, Cumulants and Diagrams, Springer, Berlin, 2010. [30] G. Peccati, C. Zheng, Multi-dimensional Gaussian fluctuations on the Poisson space, Electron. J. Probab. 15 (48) (2010) 1487–1527. [31] M. Rosenblatt, Independence and dependence, in: Proc. 4th Berkeley Sympos. Math. Statist. and Prob., vol. II, Univ. California Press, Berkeley, Calif, 1961, pp. 431–443. [32] M. Sodin, B. Tsirelson, Random complex zeroes. I. Asymptotic normality, Israel J. Math. 144 (2004) 125–149. [33] T.-C. Sun, Some further results on central limit theorems for nonlinear functions of a normal stationary process, J. Math. Mech. 14 (1965) 71–85. [34] D. Surgailis, CLTs for polynomials of linear sequences: diagram formulae with applications, in: Long Range Dependence, Birkh¨auser, Basel, 2000, pp. 111–128. [35] M. Talagrand, Spin Glasses: a Challenge for Mathematicians. Cavity and Mean Fields, Springer, Berlin, 2003. [36] M.S. Taqqu, Weak convergence to fractional Brownian motion and to the Rosenblatt process, Z. Wahrscheinlichkeitstheor. Verwandte Geb. 31 (1975) 287–302. [37] M.S. Taqqu, Convergence of integrated processes of arbitrary Hermite rank, Z. Wahrscheinlichkeitstheor. Verwandte Geb. 50 (1) (1979) 53–83. [38] L. Wang, Asymptotics of statistical estimates in stochastic programming problems with long-range dependent samples, Math. Methods Oper. Res. 55 (1) (2002) 37–54.

This article appeared in a journal published by Elsevier. The ... - ORBi lu

Stochastic Processes and their Applications 121 (2011) 793–812 ...... C. Zheng, Multi-dimensional Gaussian fluctuations on the Poisson space, Electron.

515KB Sizes 1 Downloads 165 Views

Recommend Documents

This article appeared in a journal published by Elsevier. The ... - ORBi lu
Authors requiring further information regarding ... Available online 22 December 2010 ..... (with stationary increments) with covariance structure given by. E[Bt Bs] = 1. 2(|t| ... Let {ek,k ≥ 1} be a complete orthonormal system in H. Given f ∈ H

This article appeared in a journal published by ... - Kenny Coventry
contrasts – e.g., here is this room vs. here in this galaxy”. (Kemmerer, 2006 .... played in Table 14. A 4 (Distance) x 3 (Actor) x 2 (Tool use) x 2 (Addressee.

This article appeared in a journal published by ...
Apr 7, 2009 - Most networks had a linear ND–CC relationship with a minimum CC value of 0.41. Hence .... study site); (C) Arctic animals, Greenland (Olesen, J.M., Elberling, H., ..... Centrality in social networks, conceptual clarification. Soc.

This article appeared in a journal published by ...
Feb 6, 2010 - The ESR (electron spin reso- nance) dating method is ... remains, limiting the application of ESR dating to the calcretes. (assuming that they are ...

This article appeared in a journal published by ...
[8] was time hopping with data modulation accomplished by additional pulse position modulation at the ... many pulse per data symbol. ..... Networks, Inc; 2000.

This article appeared in a journal published by ...
institutional investors, in particular foreign institutional investors .... In 2008, the non-U.S. firms in our sample account for 71% of the world .... Singapore. 53. 59.

This article appeared in a journal published by Elsevier. The attached ...
18 Apr 2009 - of the fluorescence quenching is the resonance energy transfer through long-range dipole–dipole interaction rather than the sim- ple diffusion limited process between the excited donor and the ground state acceptor molecule. 4. Energy

This article appeared in a journal published by Elsevier. The attached ...
and education use, including for instruction at the authors institution ... of Electronic Science and Technology of China, Chengdu 610054, PR China. a r t i c l e.

This article appeared in a journal published by Elsevier. The attached ...
Nauru. 1968 Vanuatu. 1980. Macedonia. 1912† USA. Swaziland 1968 Djibouti. 1977. Libya. 1911† Palau. 1994. Yemen. 1967 Comoros. 1975. Germany. Philippines 1946. Barbados. 1966 Algeria. 1962† Burundi. 1918† Marshall Is. 1986. Botswana. 1966 Ben

This article appeared in a journal published by Elsevier. The attached ...
aDepartment of Computer Architecture and Electronics, University of Almeria, Almeria 04120, Spain. bDepartment of Computer Science, ... a 3D reconstruction can be obtained by means of tomographic reconstruction algorithms. .... whose price is very mi

This article appeared in a journal published by Elsevier. The attached ...
Oct 17, 2008 - Article history: .... [4,12]: a movie has been collected by a user if and only if the rating given is at least 3 (i.e. the user at least likes this movie).

This article appeared in a journal published by Elsevier. The attached ...
Available online 24 November 2009. Editor: M. Cvetic ..... predictions of these two classes of models at the LHC as can eas- ily be seen from these figures.

This article appeared in a journal published by Elsevier. The attached ...
Available online 16 September 2013. JEL classification: F42. E44. E52. G15 ...... London Business School. Conover, C. Mitchell, Jensen, Gerald R., & Johnson, ...

This article appeared in a journal published by Elsevier ...
Dec 8, 2011 - We thank BYK Chemie Company, Momentive per- formance materials, Kuraray America ... [15] Seo MK, Park SJ. Bull Korean Chem Soc 2009 ...

This article appeared in a journal published by Elsevier ...
requested token signatures according to the outcome of this monitoring. ...... casts an accusation packet including all digital signatures through the network to ...

This article appeared in a journal published by Elsevier ...
c Sony Computer Science Laboratories, 3-14-13 Higashigotanda, Shinagawa-ku, Tokyo, 141-0022, Japan d Statistics ... Available online 15 July 2011. Keywords .... of D. The stationarity is determined at a conventional significance level of 5%.

This article appeared in a journal published by Elsevier ...
and education use, including for instruction at the authors institution ... Article history: ... tive areas in the STI technology are sloped at 70°–85° in order to.

This article appeared in a journal published by Elsevier ...
Sep 2, 2008 - and education use, including for instruction at the authors institution and sharing ... b Norwegian University of Science and Technology, Norway.

This article appeared in a journal published by Elsevier ...
websites are prohibited. ... article (e.g. in Word or Tex form) to their personal website or institutional ... self in the distorted language, in the lack of social respon-.

This article appeared in a journal published by Elsevier ...
Sep 16, 2009 - article (e.g. in Word or Tex form) to their personal website or .... Note: Data are from the Annual Social and Economic Supplement (March.

This article appeared in a journal published by Elsevier.
and education use, including for instruction at the authors institution ... Interface Nano Technology (HINT), Department of Chemical Engineering, Sungkyunkwan ...

This article appeared in a journal published by Elsevier ...
article (e.g. in Word or Tex form) to their personal website or institutional repository. ... Certain combinatorial structures have been used to construct good codes.

This article appeared in a journal published by Elsevier ...
abstract. The benefits of multisensor fusion have motivated research in this area in recent years. ... bandelet, curvelet, and contourlet transform for image fusion.

This article appeared in a journal published by Elsevier ...
and education use, including for instruction at the authors institution and sharing with colleagues .... 2.4.1. Degree Centrality. Network nodes (actor) which directly linked to a specific node are in the neighborhood of that specific node. The numbe